You are on page 1of 77

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/24888.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Wood modifications for valued-added applications using nanotechnology-based approaches

Wood modifications for valued-added applications using


nanotechnology-based approaches
Cai, Xiaolin
Philosophi doctor (Ph.D.)
Universit Laval
Doctorat en sciences du bois
Directeur(trice) de recherche : Riedl, Bernard
Co directeur(trice) de recherche : Wan, Hui
Co directeur(trice) de recherche : Zhang, S.Y. Tony
2007-11
Xiaolin Cai, 2007

Table des matires

Rsum
Abstract
Avant-Propos
Acknowledgements
List of Abreviations
Chapter 1 Wood Quality Improvement Methods and Technologies, and Polymer/Layered Silicate Nanocomposites: General
Introduction and Literature Review
1.1 Introduction
1.2 Literature Review of Wood Quality Improvement Methods and Technologies
1.2.1 Feasibility of Wood Quality Improvement
1.2.2 Technical Processes
1.2.3 Polymers Used for Wood Quality Improvement
1.2.4 Polymer Location in the Wood
1.2.5 Curing Methods for Wood Polymer Composites (WPC) Preparation
1.2.6 Wood-Hardening Process
1.2.7 Effect of Wood Permeability on Properties of WPC
1.2.8 Summary of Wood Quality Improvement Methods and Technologies
1.3 Literature Review on Polymer/Layered Silicate Nanocomposites
1.3.1 Clay Structure and the Challenge of Processing
1.3.2 Structure and Properties of Organically Modified Layered Silicate
1.3.3 Techniques Used for the Characterization of Nanocomposites
1.3.4 Synthesis and Properties of Polymer-clay Nanocomposites
1.3.5 Summary of Polymer/Layered Silicate Nanocomposites
1.4 Motivation for the Study
Chapter 2 Experimental and Methodology
2.1 Survey of Experimental Design
2.2 Wood Sample Preparation
2.2.1 Wood Cutting Design
2.2.2 Lumber Drying and Sample Preparation
2.3 Synthesis of Low Viscosity Melamine-Urea-Formaldehyde (MUF) Resin and Mix of Nanoclay and MUF Resin
2.3.1 Synthesis of Low Viscosity Melamine-Urea-Formaldehyde (MUF) Resin
2.3.2 Mixture of MUF Resin and Nanoclay
2.3.3 Characterization Ball Mill Ground Layered Aluminium Silicate Nanoclays
2.3.4 Characterizing Nanoparticles of MUF/Nanoclay System
2.4 Color Measurement
2.5 Characterizing the Interphase and Morphology of Wood MUF Nanoclay with Electron Probe Micro-Analysis (EPMA)
Element Distribution
Chapter 3 Effects of Aluminium Silicate Clays on the Morphology and Curing Behaviour of Melamine-Urea-Formaldehyde Resin
Rsum
Abstract
3.1 Introduction
3.2 EXPERIMENTAL

08-Aug-15 00:13

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/24888.html

3.2.1 Ball-milling and Characterization of the Layered Silicate Nanofillers


3.2.2 Preparation of MUF Resin and Nanoclay/MUF Resin Mixtures
3.2.3 DSC Measurement
3.2.4 Dynamical Mechanical Properties Measurements
3.3 RESULTS AND DISCUSSION
3.3.1 Effect of Ball-milling on Surface Properties of Layered Silicate Nanofillers
3.3.2 Effect of Nanofillers on the Curing Behaviour of MUF Resin from Thermal Analysis
3.3.3 Effect of Nanofillers on the Network Properties of MUF Resin
3.4 Conclusions
Acknowledgments
Chapter 4 Formation and Properties of Nanocomposites Made up from Solid Aspen Wood, Melamine-Urea-Formaldehyde, and
Clay
Rsum
Abstract
4.1 Introduction
4.2 Experimental
4.2.1 Materials and Preparation of Impregnation Solutions
4.2.2 Impregnation Process
4.2.3 Physical and Mechanical Properties Measurements
4.2.4 Thermogravimetric Analysis (TGA)
4.2.5 Statistical Analysis
4.3 Results and discussion
4.3.1 Polymer Retention
4.3.2 Surface Hardness
4.3.3 Static Bending (MOE/MOR) Analysis
4.3.4 Wear Resistance
4.4 Conclusions
Acknowledgments
Chapter 5 Effects of Nanofillers on Water Resistance and Dimensional Stability of Solid Wood Modified by MelamineUrea-Formaldehyde Resin
Rsum
Abstract
5.1 Introduction
5.2 Experiment
5.2.1 Materials and Sample Preparation
5.2.2 Moisture Absorption
5.2.3 Water Absorption and Dimensional Stability Measurements
5.2.4 Anti-Swelling-Efficiency Measurements
5.2.5 Scanning Electron Microscopy (SEM) measurement
5.2.6 Electron Probe Micro-Analysis (EPMA)
5.2.7 Color Measurement
5.3 Results and Discussion
5.3.1 Microscopy Investigations
5.3.2 Moisture Absorption
5.3.3 Water Absorption
5.3.4 Dimensional Stability
5.3.5 Color Variations
5.4 Conclusions
Acknowledgments
Chapter 6 Impact of the nature of nanofillers on the Performances of Wood Polymer Nanocomposites
Rsum
Abstract
6.1 Introduction
6.2 Experiment
6.2.1 Materials
6.2.2 Water Absorption
6.2.3 Microscopic Analysis of the WPCs
6.2.4 Thermogravimetric and Dynamical Mechanical Properties Measurements
6.3 Results and Discussion
6.3.1 Ball-milling Treatment of Nanofillers
6.3.2 Enhanced Properties of the WPC Composites
6.3.3 Improvement in Water Absorption
6.3.4 Microscopy investigations
6.3.5 Dynamic Mechanical Property Investigations
6.4 Conclusions
Acknowledgements
Chapter

7 Montmorillonite Nanoparticles Distribution

and Morphology in

Melamine-Urea-Formaldehyde (MUF) Resin

08-Aug-15 00:13

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/24888.html

Impregnated Wood Nanocomposites


Rsum
Abstract
7.1 Introduction
7.2 Experiment and Materials
7.2.1 Electron Probe Micro-analysis (EPMA) Element Distribution, Transmission Electron Microscope (TEM) and
Atomic Force Microscopy (AFM) Observation
7.3 Results and discussion
7.3.1 Montmorillonite Nanoparticles Coverage Rate on the Wood Transverse Surface of Modified Wood
7.3.2 Montmorillonite Nanoparticles Distribution on the Treated Wood Transverse Surface
7.3.3 Adhesion between Nanoparticles, MUF Resin and Wood
Acknowledgements
Conclusions and Recommendations
Recommendations for Future Work
Bibliographie
APENDIX I Scanning Electron Microscope Observation Imagines of Pure Aspen Wood, MUF Treated Aspen Wood, and
Montmorillonite Clay/MUF/Wood Nanocomposites.
APENDIX II Electron Probe Micro-Analysis (EPMA) Images of Aluminium Distribution on Different Surface
APENDIX III Transmission Electron Microprobe (TEM) Observation Images of Montmorillonite Distribution in the Cell Wall of
Treated Wood Nanocomposites
Liste des illustrations
Figure 1.1 Idealized structure of a montmorillonite layer showing two tetrahedral-site sheets fused to an octahedral-site sheet
(2:1 type) [Outubuddin and Fu 2002, Ray and Okamoto 2003].
Figure 1.2 Processing challenge for exfoliated nanoclay [Dennis et al., Southern Clay Products].
Figure 1.3 Orientations of alkylammonium ions in the galleries of clay layers with different layer charge densities [Qutubuddin
and Fu 2002].
Figure 1.4 Schematic illustrations of three types of polymer-clay composites.
Figure 1.5 Four samples of X-ray diffraction [Dennis et al, Southern Clay Products].
Figure 1.6 TEM imagine of PA6/nanoclay by extruder process [Dennis et al, Southern Clay Products].
Figure 1.7 Dispersion mechanism of nanoclay to a polymer.
Figure 1.8 Model of alkylammonium exchanged clay swollen by monomer or polymer precursors such as styrene,
-caprolactam, and epoxide [Qutubuddin and Fu 2003].
Figure 1.9 Relationship between the dispersion and the mechanical properties [Dennis et al Southern Clay Products, Inc.].
Figure 1.10 Effect of relative humidity on O2 transmission at various clay loadings [Dennis et al, Southern Clay Products, Inc.].
Figure 1.11 Water vapour transmission vs. clay content [Dennis et al, Southern Clay Products, Inc.].
Figure 2.1 Survey of the experimental procedure.
Figure 2.2 (a)Aspen green log before cutting.
Figure 2.2 (b) Sketch showing method of cutting the bolt and marking the sticks. For example in lumber AS01E1 (stands for: A
aspen, S sapwood, 01 log lumber, E eastern direction, 1 - lumber number) and AH01-1 (Stands for: A aspen, H
heartwood, 01 log number, 1 heartwood lumber number), respectively.
Figure 2.2 Wood sample cutting design (a) Aspen green log before cutting, (b) Sketch showing method of cutting up the bolt
and marking the sticks, (c) Samples end match.
Figure 2.3 Forintek staff is preparing specimens with Wood-Mizer.
Figure 2.4 Preparation of kiln drying green aspen lumbers.
Figure 2.5 Equipment of melamine urea formaldehyde resin synthesis.
Figure 2.6 Ball mill equipment.
Figure 2.7 Desintegrator for high speed mixing MUF resin with nanoclay.
Figure 2.8 Zetarsizer Nano ZS.
Figure 2.9 Principle of color meaurement.
Figure 2.10 Calculation of color differences.
Figure 2.11 Color-guide BYK-Gardner GmbH equipment
Figure 2.12 Electron Microprobe Analyzer (CAMECA S100, France).
Figure 3.1 TGA results of received nanoclays, NF1 - Cloisite30B, NF2 - ClaytoneAPA, NF3 CloisiteNa+
Figure 3.2 X-Ray diffraction analysis of the pristine nanofiller before and after the ball-mill treatment.
Figure 3.3 SEM pictures of layered aluminium silicate (a) NF1, (b) ball-milled NF1.
Figure 3.4 X-Ray composition analysis of clay received and ball mill treated clay (a) NF2, and (b) milled NF2.
Figure 3.5 DSC thermograms using different heating rates for (a) MUF resin, (b) 1% NF2/MUF, and (c) 1% NF2-m/MUF resins.
Figure 3.6 Activation energy dependency on the degree of conversion for the MUF resin, and MUF resin mixed with different
nanoclays.
Figure 3.7 Activation energy dependency on the degree of conversion for the MUF resin, 1% NF2/MUF, and 1% NF2-m/MUF
resins.
Figure 3.8 Storage modulus of MUF, 1% NF1/MUF, 1% NF2/MUF and 1% NF3/MUF nanocomposites.
Figure 3.9 Loss tangent of MUF, 1% NF1/MUF, 1% NF2/MUF and 1% NF3/MUF nanocomposites.
Figure 3.10 Storage modulus of clay/MUF and ball mill ground clay/MUF nanocomposites.
Figure 3.11 Loss tangent of clay/ MUF and ball mill ground clay/MUF nanocomposites.
Figure 4.1 Density profiles of MUF and nanoclay/MUF treated and untreated aspen wood after hot-pressing: NF1, NF2 and NF3

08-Aug-15 00:13

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/24888.html

are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled nanofillers.
Figure 4.2 Mechanical properties of aspen and impregnated aspen with MUF and 1 wt % montmorillonite/MUF. a) Brinell
hardness, b) Wear resistance of MUF and nanoclay/MUF (100cycles/300cycles/500cycle) treated samples NF1, NF2 and NF3 are
montmorillo
Figure 4.3 Relationship between strength at rupture and elasticity modulus (MOR/MOE) of aspen and impregnated aspen with
MUF and montmorillonite/MUF. NF1, NF2 and NF3 are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled
nanofillers.
Figure 5.1 Scanning electron microscopy photographs of (a) control wood samples, (b) MUF-treated wood sample, containing
42.1% polymer, (c) NF1/MUF-treated wood sample, NF1/MUF content 40.9%, (d) NF2/MUF-treated wood sample, containing
40.6% NF2/MUF, and (e) NF3/MUF-treated wood sample, containing 35.6% NF3/MUF.
Figure 5.2 Electron Microprobe Analysis of NF2/MUF treated wood, (a) transversal face, and (b) Al elemental distribution.
Figure 5.3 Water repellency efficiency.
Figure 5.4 Antiswelling efficiency, (a) tangential direction and radial direction, (b) ASE in volume.
Figure 6.1 X-ray diffraction of the pristine nanofiller before and after the ball-mill treatment.
Figure 6.2 X-ray composition analysis of the hydrophobic nanoclay NF1, NF2 before (a,c) and after ball-milling treatment (b,d),
respectively.
Figure 6.3 SEM photographs of the fracture surface of (a) untreated, (b) MUF treated, (c) NF3/MUF treated, and (d) NF2/MUF
treated wood. (e) and (f) are trace of nanoclay in the nanoclay/MUF treated wood.
Figure 6.4 Electron microprobe analysis of aluminum distribution in the NF2/MUF treated wood: (a) transverse image, and (b)
Al distribution.
Figure 6.5 TEM photographs of (a) MUF and (b) NF2/MUF treated wood cell wall.
Figure 6.6 Storage modulus E profiles of pure control wood, and MUF, NF2/MUF and NF3/MUF impregnated wood.
Figure 6.7 tan profiles of pure control wood, MUF, NF2/MUF and NF3/MUF impregnated wood
Figure 7.1 Image of Wincell analysis window.
Figure 7.2 Ultrathin sections of the walls of aspen vessel with various wall layers: ML (middle lamella), M (compound middle
lamella), P (primary wall, S1 (the out layer of secondary wall), S2 (the centre layer of secondary wall), T (tertiary wall).
Figure 7.3 Distribution of montmorillionite nanoparticles along the transverse ultrastructure of aspen/MUF/nanocomposites.
Figure 7.4 Photographs of AFM analysis control sample and treated aspen samples, a) aspen wood, resolution with axis X 0.5
m/div and axis Z 200 nm/div.
Figure 7.4 b) MUF impregnated aspen wood, resolution with axis X 100 nm/div and axis Z 100 nm/div.
Figure 7.4 c) MUF_NF1/aspen wood, resolution axis X 2 m/div and axis Z 200 nm/div
Figure 7.4 d) MUF_NF1/aspen wood, resolution with axis X 0.5 m/div and axis Z 100 nm/div.
Figure 7.4 e) MUF_NF2/aspen wood, resolution with axis X 100 nm/div and axis Z 50 nm/div.
Figure 7.4 f) MUF_NF3/aspen wood, resolution with axis X 0.2 m/div and axis Z 50 nm/div
Figure I 1 SEM analysis (transverse section) of pure aspen wood.
Figure I 2 SEM analysis (transverse section) of MUF treated wood.
Figure I 3 SEM analysis (transverse section) of MUF/NF2/wood nanocomposites.
Figure I 4 SEM analysis (transverse section) of MUF/NF2/wood nanocomposites at higher magnification with rough surface
(microscale).
Figure I 5 SEM analysis (transverse section) of MUF/NF2_bm/wood nanocomposites at high magnification with fine surface
(microscale).
Figure II 1 EPMA observation (transverse section) with high magnification (microscale).
Figure II 2 EPMA observation of aluminium element distribution (transverse section, cell wall) with high magnification
(microscale).
Figure II 3 EPMA observation of aluminium element distribution (transverse section, cell wall, cell lumen).
Figure II 4 EPMA observation (tangential section) with high magnification (microscale).
Figure II 5 EPMA observation aluminium element distribution (tangential section) with high magnification, color picture
(microscale).
Figure II 6 EPMA observation of aluminium element distribution (tangential section) with high magnification, black and white
(microscale).
Figure II 7EPMA observation (radial section) with high magnification (microscale).
Figure II 8 EPMA observation of aluminium element distribution (radial section) with high magnification, color picture
(microscale).
Figure II 9 EPMA observation of aluminium element distribution (radial section) with high magnification, black and white
(microscale).
Figure II 10EPMA observation (radial section) with higher magnification (microscale).
Figure II 11 EPMA observation of aluminium element distribution (radial section) with higher magnification, color picture
(microscale).
Figure II 12 EPMA observation of aluminium element distribution (radial section) with higher magnification, black and white
(microscale).
Figure III 1 TEM observation of MUF/NF2/wood nanocomposite (transverse section, microscale).
Figure III 2 TEM observation of MUF/NF2/wood nanocomposite (transverse section, microscale).
Figure III 3 TEM observation of MUF/NF2/wood nanocomposite (transverse section, cell wall, nanoscale).
Figure III 4 TEM observation of MUF/NF2/wood nanocomposite (transverse section, cell wall, nanoscale).
Figure III 5 TEM observation of MUF/NF2/wood nanocomposite (transverse section, cell wall part, nanoscale).
Figure III 6 TEM observation of MUF/NF2/wood nanocomposite (transverse section part M, ML layers, microscale magnification).
Figure III 7 TEM observation of MUF/NF2/wood nanocomposite (transverse section, part M, ML layers, nanoscale).

08-Aug-15 00:13

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/24888.html

Figure III 8 TEM observation of MUF/NF2/wood nanocomposite (transverse section, part M, ML layers, nanoscale).
Figure III 9 TEM observation of MUF/NF2/wood nanocomposite (transverse section, part M, ML layers nanoscale).
Liste des tableaux
Table 1.1 Comparison of viscosities of liquids used in preservative treatmetn compared to those of several vinyl monomers [Loos
et al 1967].
Table 1.2 Properties of styrene-treated and untreated southern pine [Chao and Lee 2003].
Table 1.3 Summary of tangential hardness and air permeability data [Young and Meyer 1968].
Table 1.4 Comparison of the measured properties and the specific properties (Measured divided by density) from the literature
data vs. the data for balsa impregnated with EHMA-ST-pBDDA [Wright and Matias 1993].
Table 1.5 Effect of relative molecular size of phenolic resin and concentration on the dimensional stabilization of wood [Bryant
1966].
Table 1.6 The effect of phenolic resin impregnation on mechanical properties of wood [Bryant 1966].
Table 1.7 Compression set (C2) and set-recovery (R2) of compressively molded squares [Ito et al 1998a].
Table 1.8 Brinell hardness of CWPC at 50% compression set and WPC obtained by vinyl polymerization with and without boric
acid (BA) pretreatment. [Yalinkilic et al 1999]
Table 1.9 Chemical formula and characteristic parameters of commonly used 2:1 phyllosilicates [Ray and Okamoto 2003].
Table1.10 Mechanical and thermal properties of Nylon-6 and Nylon-6-Clay nanocomposites [Kurauchi 1991].
Table 3.1 Physical properties of nanoclays.
Table 3.2 Gel time test results of MUF resin and layered aluminium silicate nanoclay/MUF at different temperatures.
Table 4.1 Properties of nanoclays.
Table 4.2 T-test analysis of density and weight gain of MUF/nanoclay treated and untreated aspen1.
Table 4.3 t-test analysis of Brinell hardness of MUF/nanoclay treated and untreated aspen1.
Table 4.4 t-test analysis of modulus of elasticity (MOE) and modulus of rupture (MOR) of MUF/nanoclay treated and untreated
aspen1.
Table 4.5 Tukey analysis of wear resistance index of MUF/nanoclay treated and untreated aspen1.
Table 5.1 Moisture absorption and moisture content at equilibrium (at 21C, 65% MC for 6 weeks) of MUF and nanoclay/MUF
treated and untreated aspen1.
Table 5.2 Water absorption in one day and one week of MUF and nanoclay/MUF treated and untreated aspen1.
Table 5.3 Thickness swelling (radial) after 1 day and 1 week in water of MUF and nanoclay/MUF treated and untreated aspen1
samples.
Table 5.4 Swelling in width (tangential) (1 day and 1 week) for MUF and nanoclay/MUF treated and untreated aspen1 samples.
Table 5.5 Swelling in length (1 day and 1 week) for MUF and nanoclay/MUF treated and untreated aspen1 samples.
Table 5.6 Color variation values of the MUF- and nanofiller/MUF-treated wood samples as compared to the controls.
Table 6.1 Properties of nanoclays.
Table 6.2 Weight gains for the treated samples, and physical and mechanical properties of the control and the prepared wood
polymer composites
Table 7.1 Calculation of surface area of aluminium distribution on the wood surface via Wincell software.
Table 7.2 Layered silicate distribution with the change of concentration of aluminium distribution.
Xiaolin Cai, 2007

08-Aug-15 00:13

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/pr01.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Table des matires


Rsum
Abstract
Avant-Propos
Acknowledgements
List of Abreviations

Rsum
Le dveloppement de produits en bois transforms et ayant de la valeur ajoute prsente une occasion importante de bnficier des
ressources naturelles canadiennes pour l'industrie des produits en bois. Optimiser lemploi de la ressource de la fort par
l'amlioration des traitements est une bonne manire dajouter une valeur ajoute aux produits en bois. Nos objectifs ont vis le
march des produits d'application de plancher par un procd appropri de modification du bois de mauvaise qualit. La combinaison
de l'imprgnation, de la nanotechnologie et de lutilisation de produits chimiques est une approche prometteuse pour amliorer les
proprits du bois telles que la duret, la rsistance l'abrasion et la stabilit dimensionnelle. L'application de la nanotechnologie
dans ce projet a fourni une nouvelle mthodologie pour dvelopper des produits valeur ajoute pour l'industrie des produits en bois.
Les nanocomposites en bois et polymre ont t prpars partir du bois massif de peuplier (Populus tremuloides), de la rsine
hydrosoluble de type mlamine-ure-formaldhyde (MUF), et de nanoargiles de silicates d'aluminium, appeles montmorillonites.
Des nanoargiles hydrophiles et hydrophobes de montmorillonite ont t ajouts aux MUF. La dispersion des nanoargiles est
dimportance cruciale pour maximiser l'avantage de cet ajout de nanoparticules. cette fin, les nanoargiles de montmorillonite ont
t broyes et rectifis avec un moulin billes avant d'tre mlangs de la rsine MUF et imprgnes dans le bois. Le bois a t
imprgn de rsine MUF, qui a polymris in situ dans certaines conditions.
L'influence des nanoparticules de montmorillonite sur le comportement du bois suite au traitement avec de la rsine MUF et les
proprits de visco-lasticit ont t tudies par la calorimtrie balayage diffrentiel (DSC) et l'analyse de DMTA (analyse
thermique mcanique dynamique).
Des amliorations significatives des proprits physiques et mcaniques, telles que la duret, la rsistance l'abrasion et le module
d'lasticit (MOE) ont t obtenues pour les spcimens imprgns de mlanges de rsine MUF et de rsines nanoargile-MUF. Des
amliorations significatives de rsistance l'eau et de stabilit dimensionnelle ont t obtenues pour le bois trait par
nanorenfort/MUF. L'efficacit antigonflante (ASE) a t amliore de 63.3% 125.6% pour le bois trait par nanorenfort/MUF.
L'interactions linterface et la morphologie entre les nanorenforts, le MUF et le bois sur les proprits physiques et mcaniques des
nanocomposites en bois ont t tudies en utilisant la spectroscopie des lectrons de rayon-X, la microscopie lectronique balayage
(SEM), la microscopie lectronique de transmission (TEM), la microscopie force atomique (AFM) et les mthodes de microanalyse de
sonde lectronique (EPMA). Les proprits amliores des bois ont pu tre attribues aux proprits inhrentes aussi bien qu au
meilleures interactions linterface du bois, de la MUF et des nanoparticules. Le traitement de moulin billes a favoris la dispersion
des nanorenforts dans le bois, mais a dcompos les groupes fonctionnels sur la surface des nanoargiles hydrophobes, ce qui a t
nuisible pour la liaison entre le nanorenfort et la matrice de MUF.
Suite lacquisition dimages de la microsonde de fluorescence de rayons-X, les images de la distribution daluminium, lment
uniquement prsent dans largile, ont t analyses laide du logiciel Wincell. En comparant ces images et les micrographies de
microscopie conventionnelle rayons-X, il a t constat que les renforts de montmorillonite sont distribus de faon non-homogne
dans la lamelle moyenne, la lamelle moyenne compose, la paroi primaire et le mur secondaire. Ces lments fonctionnent comme
des filtres qui ont captur les nanoparticules de montmorillonite dans les rgions amorphes cest--dire non cristallines de la paroi
cellulaire. On a aussi observ l'adhrence entre la montmorillonite et la rsine MUF par microscopie AFM. On a constat que les
groupes fonctionnels la surface des montmorillonites hydrophobes jouent un rle important de comptabilisation entre la
montmorillonite nanoargile et la rsine MUF, ce qui exerce une influence forte sur les proprits physique et mcaniques des
nanocomposites en bois/ nanoargile/MUF.

Abstract
The development of value-added wood products from low-quality resource through innovative technology presents an excellent
opportunity to maximize the value from the forest resource and thus contributes to the global competitiveness of the wood industry
in Canada. To combine nanotechnology with chemical impregnation technique becomes particularly appealing to improve some
value-added wood attributes such as wood surface hardness, abrasion resistance and dimensional stability. The combination of
nanotechnology with the traditional impregnation technique has provided a new approach to improve the wood quality attributes of
critical importance to value-added applications.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/pr01.html

In this study, wood polymer nanocomposites were prepared from solid aspen (Populus tremuloides) wood, water-soluble melamineurea-formaldehyde (MUF) resin, and montmorillonite aluminium silicate nanoclays. Both hydrophilic and hydrophobic montmorillonite
nanoclays were introduced to the system. The dispersion of nanoclay is crucial for completely utilizing the concept of nanoparticles. To
do so, the montmorillonite nanoclays were ground with a ball-mill before being mixed with MUF resin and impregnated into the aspen
wood. The wood samples were impregnated with resin, which polymerized in situ under specific conditions.
The influence of the montmorillonite nanoparticles on the curing behaviour of MUF resin and visco-elasticity properties were
investigated using differential scanning calorimetry (DSC) and dynamical mechanical thermal analysis (DMTA).
Significant improvements in wood physical and mechanical properties, such as surface hardness, abrasion resistance, modulus of
elasticity (MOE) were observed for the specimens impregnated with MUF resin and nanoclay-MUF resin mixtures. Significant
improvements in water repellence and dimensional stabilities were also found for the nanofiller/MUF treated wood. The antiswelling
efficiency (ASE) was improved from 63.3% to 125.6% for the nanofiller/MUF treated wood.
This study also examined the influence of the interphase interactions and morphology between the nanofillers, MUF and wood on the
physical and mechanical properties of the resulting wood-polymer nanocomposites using X-ray electron spectroscopy, scanning
electron microscope (SEM), transmission electron microscope (TEM), atomic force microscope (AFM) and electron probe microanalysis (EPMA). The improved wood properties could be ascribed to inherent properties as well as better interphase interactions
between the wood, MUF and nanofillers. Ball-mill treatment favoured the dispersion of the nanofillers into the wood, but broke down
functional groups on the hydrophobic nanoclay surface, which was detrimental for the bonding between the nanofiller and MUF
matrix.
The montmorillonite nanoclay coverage rate on the nanofiller/MUF wood nanocomposite surface was further investigated using the
Wincell software analysis of the images of aluminium distribution. By duplicatine the image of aluminium distribution to the part was
observed, it was found that the distribution of montmorillonite nanoclay looks like a network along the layer of ML (middle lamella),
M (compound middle lamella), P (primary wall), S1 (secondary wall 1). These parts function like a sieve which captured the
montmorillonite nanoparticles in the amorphous substance. The adhesion between montmorillonite and MUF resin was observed with
AFM. It was confirmed that the functional groups of the organophilic montmorillonites play an important role on the compatibility
between montmorillonite nanoclay and MUF resin, have strong influence on the physical/mechanical properties of the nanoclay/MUF
wood nanocomposites.

Avant-Propos
Ce document est prsent sous la forme d'une thse de publications. Il a t conu selon les critres de prsentation adopts par le
comit du programme de 2me et 3me cycles en sciences du bois de l'Universit Laval, en avril 1988. On retrouve dans cet ouvrage
cinq articles crits en anglais prsents dans les chapitres 3, 4, 5, 6 et 7. Une description sommaire de ces articles est propose
ci-aprs :
Chapitre 3:
Cai X., Riedl B., Zhang S.Y., Wan H., Wang X. M. 2007. The effect of layered aluminum silicates on the morphology and curing
behaviors of melamine-urea-formaldehyde resin. Il sera soumis pour publication dans la revue Polymer.
Chapitre 4:
Cai X., Riedl B., Zhang S.Y., Wan H. 2007. Formation and properties of nanocomposites made up from solid aspen wood, melamineurea-formaldehyde, and clay. Holzforchung, 61: 148-154.
Chapitre 5:
Cai X., Riedl B., Zhang S.Y., Wan H. 2007. Effects of nanofillers on water resistance and dimensional stability of solid wood modified
by melamine-urea-formaldehyde resin. Wood and Fiber Sci. 39(2) :307-318.
Chapitre 6:
Cai X., Riedl B., Zhang S.Y., Wan H. 2007. The impact of the nature of nanofiller on the performance of wood polymer
nanocomposites. Composites Part A : Applied science and manufacturing, admis.
Chapitre 7:
Cai X., Riedl B., Zhang S.Y., Wan H. 2007. Montmorillonite nanoparticles distribution and morphology in melamine-urea-formaldehyde
(MUF) resin impregnated wood nanocomposites. Il sera soumis pour publication dans la revue Holzforchung.
Ces articles ont t tous rdigs par lauteure de cette thse, laquelle a fait la rdaction et dvelopp le contenu de la premire
version de ceux-ci, et par le directeur de cette thse, le professeur Bernard Riedl, et les deux co-directeurs Dr. S. Y. (Tony) Zhang et
Dr. Hui Wan chez FP Innovations div Forintek, lesquels ont ralis la rvision portant autant sur le fond que sur la forme de ces
articles.
Larticle prsent au Chapitre 3 a t crit en collaboration avec le Chercheur et chef de groupe Dr. Xiangming Wang, du Dpartement
composites chez FP Innovations div Forintek.
Les rsultats de ce travail ont galement t prsents aux confrences suivantes:
Wood Science and Engineering in the Third Millennium 20-22 juin 2007, Brasov, Romania. Larticle Characterization of wood polymer

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/pr01.html

nanocomposites made of melamine-urea-formaldehyde and layered silicate clays a t publi dans les comptes-rendus de cette
conference.
Colloque Nano 2007, 7 fvrier, Montral, QC, Canada. Le poster Effet de nanoparticule sur les proprits du bois/mlamine
ureformaldhyde nanocomposites a t publi dans les comptes-rendus de cette confrence.
NERM (American Chemistry Society, ACS) 2006, The 34th Northeast Regional Meeting, 5-7 octobre, Binghamton, NY, USA. Le poster
The impact of interphase between wood, melamine-urea-formaldehyde and layered silicate on the performance of wood polymer
nanocomposites a t publi dans les comptes-rendus de cette confrence.
BioPlastics 2006 International Symposium on Bioplastics and Natural Fiber Composites. 27-29 septembre, Montral, Canada. Larticle
Effect of nanofiller on the performance of wood/melamine-urea-formaldehyde nanocomposites a t publi dans les comptes-rendus
de cette confrence.
74e Congres de LAcfas 2006, 16 mai, Montral, QC, Canada. Larticle Effets de la combinaison nanoparticules/MUF sur la rsistance
leau et la stabilit dimensionnelle a t publi dans les comptes-rendus de cette confrence.
PPS (Polymer Processing Society) 2005 Americas Regional Meeting, 14-17 aot, Qubec City, Canada. Le poster Formation and
properties of aspen solid wood/MUF/clay nanocomposites a t prsent dans les comptes-rendus de cette confrence.
International Polymer Materials Engineering Conference (IPMEC), 18-21 septembre 2005, Shanghai, Chine, Larticle Preparation and
properties of wood/MUF/clay nanocomposites a t publi dans les comptes-rendus de cette confrence.
59me Congrs International de la Forest Product Society, qui a eu lieu Qubec, du 20 au 23 juin 2005; Larticle Formation and
properties of aspen solid wood/MUF/clay nanocomposites a t publi dans les comptes-rendus de cette confrence.
73e Congres de LAcfas, 10 mai 2005 Chicoutimi, QC, Canada. Larticle Fabrication et proprits de nanocomposites de bois de
peuplier MUF/argile a t publi dans les comptes-rendus de cette confrence.

Acknowledgements
I would like to express my sincere gratitude to my director, Prof. Bernard Riedl, whose wide knowledge and logical way of thinking
was invaluable on numbers of key points during the full course of my study. His understanding, encouraging and guidance have built
a solid foundation for the thesis. I wish to express my sincere appreciation to my co-directors, Dr. Hui Wan from FPInnovations
Forintek (Eastern Region) and Dr. S.Y. (Tony) Zhang from FPInnovations Forintek (Western Region)for their detailed and
constructive comments, their tremendous support and encouragement throughout this work.
My sincere thanks are due to the official referees, Prof. Alain Cloutier, Director de Centre de recherches sur le bois, Prof. Denis
Rodrigue from Department Chemical Engineering of Universit Laval and Prof. Marie-Pierre Laborie from Washington State University,
for their detailed review, constructive criticism and excellent advice during the preparation of this thesis.
Much of the research work from MUF resin synthesis, to wood polymer nanocomposites manufacturing and testing was carried out
and supported by FPInnovations Forintek (Eastern Region) and TEMBEC adhesive group. I would like to give my thanks to these
organizations. Acknowledgements are also due to Mr. Gilles Brunette, the manager of the composite products department, and Mr.
Grald Beaulieu, the manager of value added department of FPInnovations Forintek (Eastern Region), for their support and advice
toward trial arrangement, wood impregnation and testing.
My warm appreciations are due to Dr. Xiangming Wang, FPInnovations Forintek (Eastern Region) for his valuable advice and
friendly help. His extensive discussions about my work and interesting explorations in operations have been very helpful for this
study. I wish to express my thanks to Dr. Chuangming Liu for his kind help on statistic analysis of the experimental results.
Additionally, the following personnel are thankful for their technical assistances, Mr. Miro Luptovski, Mr. Marcel Lefebvre, Mr. Denis
Lachance, Ms. Hongliu Zhou, Ms. Francine Cot, Mr. Stphane Raymond, Mr. Yves Lavoie at FPInnovations Forintek (Eastern
Region).
I wish to extend my warmest acknowledgement to Mr. Marc Choquette, for his kind help to training me using the SEM and Electron
Probe Micro-Analysis at Laboratoire de microanalyse - dpartement de gnie des mines, de la mtallurgie et des matriaux of
Universit Laval. My thanks are also due to Mr. Richard Janvier, a specialist on microscopy and histology unit at Recherche en sciences
de la vie et de la sant of Universit Laval.
My sincere appreciations are due to Mr. Yves Bdard and Ms. Rodica Plesu, for their directions and help in my experimental work, Mr.
Ndanga Lionel Kamwenubusa, a trainee, for his hard work to prepare the samples and the DMTA measurement, Mr. Pierre-Louis Cyr
and Mr. Guillaume Giroud, for their help in AFM and Wincell measurements, respectively. I would like to express my thanks to all
members in dpartement de sciences du bois et de la fort, Centre de recherches sur le bois (CRB) and Centre de recherche en
science et ingnierie des macromolecules (CERSIM), for their help and cooperation. Many thanks also go to my friends, for their
encouragement and help.
The financial support of the Natural Sciences and Engineering Research Council of Canada (NSERC), the Fonds Qubcois pour la
Recherche en Sciences et en Technologie (FQRST), FPInnovations Forintek (Eastern Region) and Universit Laval Fonds de soutien
au doctorat for this project is gratefully acknowledged.
Finally, I am deeply grateful to my wonderful family, my parents, my brother, my husband Jingui Huang, and my son Eric Taiyi Huang.
Their tremendous love and support has given me the courage to follow a dream.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/pr01.html

List of Abreviations

AFM:

atomic force microscope

ASE :

anti-swelling efficiency

BA :

boric acid

CEC :

cation exchange capacity

CWPC

compressed-wood polymer composites

DMTA: dynamic mechanical thermal analysis


DSC:

differential scanning calorimetric

EHMA : ethyl-a-hydroxymethylacrylate
EPMA: electron probe micro-analysis
FA :

furfuryl alcohol-based

FTIR:

Fourier-transform infrared analysis

GMA : glycidyl methacrylate


HMA : hydroxymethylacrylate
LVL :

laminated veneer lumber

MA :

maleic anhydride

MF:

melamine formaldehyde

MMA : methyl methacrylate


MMT : montmorillonite
MOE : modulus of elasticity
MOR:

modulus of rupture

MUF :

melamine-urea-formaldehyde resin

MW :

molecular weight

PA6 :

polyamide 6

PEG :

polyethylene glycol

PF :

phenol-formaldehyde

PG :

polyglycerol

PGMA : poly(glycerol methacrylate)


RH :

relative humidity

SD:

standard deviation

SEM:

scanning electron microscopy

SG :

specific gravity

ST :

styrene

TEM:

transmission electron microscopy

TGA:

thermogravimetric analysis

UF :

urea-formaldehyde

WAXD: wide angle X-ray diffraction


WPC : wood polymer composites
WPNC wood polymer nanocomposites
XRD:

X-ray diffraction

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/pr01.html

Xiaolin Cai, 2007

Wood modifications for valued-added applications


using nanotechnology-based approaches

Chapter 1 Wood Quality Improvement Methods and


Technologies, and Polymer/Layered Silicate
Nanocomposites: General Introduction and
Literature Review

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Chapter 1 Wood Quality Improvement Methods and Technologies, and Polymer/Layered Silicate Nanocomposites:
General Introduction and Literature Review

Chapter 1 Wood Quality Improvement Methods and Technologies, and Polymer/Layered Silicate
Nanocomposites: General Introduction and Literature Review
Table des matires
1.1 Introduction
1.2 Literature Review of Wood Quality Improvement Methods and Technologies
1.2.1 Feasibility of Wood Quality Improvement
1.2.2 Technical Processes
1.2.3 Polymers Used for Wood Quality Improvement
1.2.4 Polymer Location in the Wood
1.2.5 Curing Methods for Wood Polymer Composites (WPC) Preparation
1.2.6 Wood-Hardening Process
1.2.7 Effect of Wood Permeability on Properties of WPC
1.2.8 Summary of Wood Quality Improvement Methods and Technologies
1.3 Literature Review on Polymer/Layered Silicate Nanocomposites
1.3.1 Clay Structure and the Challenge of Processing
1.3.2 Structure and Properties of Organically Modified Layered Silicate
1.3.3 Techniques Used for the Characterization of Nanocomposites
1.3.4 Synthesis and Properties of Polymer-clay Nanocomposites
1.3.5 Summary of Polymer/Layered Silicate Nanocomposites
1.4 Motivation for the Study

1.1 Introduction
Wood has many desirable characteristics required for a variety of end uses such as construction, furniture, and tools. A shortage of
high quality hardwoods has driven researchers and manufacturers to search for alternative resources, such as softwoods and some
low-density hardwoods, for value-added applications. To achieve this goal, suitable technologies are needed to improve wood quality
properties such as dimensional stability, durability, mechanical properties and hardness, to meet specific end-use requirements.
Wood has some features that may have negative impact if handlesd improperly, such as high moisture uptake, biodegradability, and
dimensional variations. The dimensional instability of wood due to moisture changes limits its use in certain applications, especially
for external uses. The various defects of wood could be attributed to the chemical structure and some anatomical characteristics
inherent in wood [Rowell 1983]. Chemically, wood contains numerous reactive sites. The most common reactive sites are the
hydroxyl groups, which are abundantly available in the three major chemical components of wood, i.e. cellulose, hemicellulose, and
lignin. These hydroxyl groups adsorb moisture from humid environments, which then enters the wood matrix. Moisture is held by
hydrogen bonding, and wood moisture changes caused by dynamic humidity conditions result in alternate swelling and shrinkage of
wood and in physical degradation, sometimes leading to mechanical failure [Kumar 1994]. On the other hand, wood is a porous
material. Anatomically, the porosity of both cell cavities (cell-lumen) and micropores in the cell walls exists. The cell cavities are
major paths for moisture movement in wood [Schneider 1994]. Blockage of these reactive sites with some reactive chemicals or
plugging the pores with a polymer could not only make the wood more resistant to changes in moisture content, but improve its
dimensional stability and physical and biodegradation properties as well.
People have made great efforts to reinforce solid wood with polymers in the past decades. Both thermoplastic and thermosetting
systems have been used [Schneider 1994, Ayer et al 2003]. It is observed that formation of wood polymer composites (WPC) is a
promising way to improve wood properties. The most common and inexpensive technique developed to prepare wood polymer
composites is by impregnating wood with a polymeric monomer or prepolymer, followed by in situ polymerization of the monomer or
curing the prepolymer to form a solid. Attempts to dimensionally stabilize wood by treating it with thermoplastic-related monomers
such as acrylates, or methacrylate monomers, have been made. The success, however, was limited. Most efforts to stabilize wood
using monomers have failed because the monomers did not penetrate the cell-wall or react with the wood. The polymers formed thus
only fill the empty lumens in the wood. This leads to a mixture of two materials of the wood and the polymer rather than a true
composite. Accordingly, the resulting products are still subject to dimensional changes with water uptake.
A desirable wood polymer composite should have the ability to transfer energy efficiently from the polymer matrix to the rigid
cellulose component. Instead of using polymers only as bulking agents, it is better to seek ways to generate wood-polymer interaction
and/or chemical bonding that will involve polymers performing an active role in the composite. This may result in increased
mechanical strength, dimensional stability, and resistance to degradation and deterioration of the wood. It has been observed that
wood impregnated with thermosetting resins such as water-soluble phenolic resins, urea-formaldehyde and melamine-formaldehyde
prepolymers, improves its compressive strength properties and moisture-related shrinking and swelling behaviors. The thermosetting
resins penetrate and bulk the cell wall, and may react with the hydroxyl groups of the wood components, preventing shrinkage of the

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

wood upon drying. But in most cases, the resulting wood-polymer composites are quite brittle, and the toughness of the original wood
are deteriorated [Deka and Saikia 2000].
Nanocomposite technology using organophilic layered silicates as in situ nano-reinforcement offers new opportunities for the
modification of thermoset micromechanics. Large improvements of physical and mechanical properties including tensile modulus and
strength, flexural modulus and strength, thermal stability, flame resistance, and barrier resistance have been observed for various
thermoplastic and thermoset nanocomposites at low silicate content [Ray and Okamoto 2003]. It should be possible to compensate
the matrix flexibilization of a thermoset via matrix reinforcement using organophilic layered silicates. Polymer-clay nanocomposites
were first reported in the literature as early as 1961 [Blumstein, 1961]. Polymer/layered silicate nanocomposites were first used in
industrial materials by researchers from Toyota based on the thermoplastic polyamide 6 [Usuki et al 1993a, b]. Since then, a great
deal of research has been carried out in the field of polymer nanocomposites with various thermoplastic and thermoset polymers over
the past decade [Qutubuddin and Fu 2002, Ray and Okamoto 2003].
Barrier properties, fire resistance and mechanical properties are of great importance for the application of a wood product. It would be
attractive to prepare a WPC product using nano technique and to improve the above properties of WPC products. Little work, however,
has been done to prepare a wood polymer composite through nano technology. This led us to investigate the feasibility to prepare a
wood polymer nanocomposite using nano technique approach.

1.2 Literature Review of Wood Quality Improvement Methods and Technologies


1.2.1 Feasibility of Wood Quality Improvement
Dimensional change and biological degradation of wood have greatly been ascribed to the presence of numerous hydroxyl groups in
wood components and various cavities in the wood [Kumar 1994]. These hydroxyl groups are considered to be reactive sites and the
cell cavities are main paths for moisture movement in wood. Thus wood quality may be changed if these reactive sites could be
substituted by some reagents and/or the pores in the wood were blocked. Blockage of these reactive sites with some reactive
chemicals or plugging the pores with a polymer will make the wood more resistant to changes in moisture content and improve its
dimensional stability, physical and biodegradation properties. Wood could be modified through different methods such as chemical,
thermal or enzymatic modification. The concept of chemical modification of wood was developed primarily to improve the dimensional
stability of wood when subjected to change of moisture. Treatments range from application of heat to impregnation with monomers or
prepolymers for in situ polymerization, or alteration of the chemical composition of wood by chemical reactions. The essential
requirements of chemical reactions are that the reacting chemicals should penetrate into the cell wall and react with the available
hydroxyl groups of the cell-wall polymer, preferably in neutral or mild alkaline conditions at temperatures below 120C [Kumar
1994]. Different technologies were attempted to make wood-polymer composites. Typical methods based on the process procedures
can be found such as impregnation, compression and compreg (the combination of impregnation and compression). The polymer types
used in wood quality improvement are thermoplastics such as vinyl monomer and similar oligomers, or thermosets such as phenol
formaldehyde, urea formaldehyde, melamine formaldehyde, and epoxy resins. The location of the added chemicals in the wood could
be found in the cell lumen, cell wall or the combination of both in cell lumen and cell wall.

1.2.2 Technical Processes


The basic technical processes for wood quality improvement can be classified into the following types [Ayer et al 2003]:
1) Impregnation of wood with reactive chemicals. This method, which is applicable to both lumber and veneer to harden wood, has
been developed since the mid-1960s. The wood impregnation process depends greatly upon diffusion and permeability of the gas or
liquid into the wood. Water and air must be removed from the wood before impregnation to provide space for the treating fluid. Low
or medium viscosity treating fluid is required, which can be forced into wood using vacuum or pressure. Vinyl monomers such as
styrene and methyl methacrylate, are examples of low viscosity fluids largely used in wood polymer composites preparation through
impregnation technique.
2) Compression of wood - The mechanical properties of wood such as stiffness, strength, surface hardness, and abrasion resistance
can be improved by compression perpendicular to the grain direction, resulting in higher wood density. However, most compressed
wood can expand almost to its initial uncompressed state under the influence of moisture. Therefore, permanent fixation of
deformation after compression is the key to the success of this approach [Ayer et al 2003].
3) Chemical impregnation and compression of wood [Fry and Harry 1976]. The chemical (such as phenol-formaldehyde resin)
impregnated dry wood is compressed in a hot press, and the resin is cured during the process at certain temperature. The
impregnated resin can be used to crosslink the veneers without the use of additional adhesive. High-density wood-polymer
composites were obtained by this approach.

1.2.3 Polymers Used for Wood Quality Improvement


The polymers used in the wood-polymer composites could be thermoplastic and thermosetting.
1) Thermoplastic, such as vinyl monomers or similar monomers, modified vinyl monomers and polar monomers (e.g. furfuryl alcohol).
Vinyl polymers have a large range of properties from soft rubber to hard, brittle solids depending upon the groups attached to the
carbon-carbon backbone. Some examples of vinyl monomers are: vinyl chloride, vinyl acetate, acrylonitrile, ethylene oxide, acrylates
(especially methyl methacrylate), t-butyl styrene, styrene and chlorostyrene [Meyer 1981]. Vinyl type monomers could be
polymerized into a solid polymer by means of heat, radiation, or heat-catalyst polymerization [Siau and Meyer 1966].Since most
vinyl monomers are nonpolar, little interaction between wood hydroxyl groups and the resulting vinyl polymers is expected. In

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

general, vinyl polymers fill the capillaries, vessels, and other void spaces in the wood structure [Timmons et al 1971]. Polar
monomers can swell cell wall to give high chemical resistance and dimensional stability in water. Modification of vinyl monomers by
introduction of polar groups has been attempted for some specific applications such as for cell wall swelling, fungicide or fire
retardancy [Schneider 1994].
2) Thermosetting polymerizable substances are in general prepolymers, which include phenol formaldehyde, urea formaldehyde,
melamine formaldehyde products, polyurethanes, epoxides, silicones, and unsaturated polyesters [Paavo and Markku 1994]. Phenolformaldehyde resin is located mostly in the cell wall of wood and yields dimensionally stable composites [Kumar 1994].

1.2.4 Polymer Location in the Wood


Monomer or prepolymer grafting can take place with reactive groups on wood components within the cell wall. Ungrafted bulk
polymers, on the other hand, can be formed in the wood voids. These two types of composite formation can give different properties
to the wood [Rowell et al 1982]. The polymer in the wood can be found broadly into three different locations [Schneider 1994, 2000,
Ayer et al 2003]:
1) In cell-lumen, eg. mostly vinyl monomers such as methyl methacrylate and styrene. The monomer or prepolymer introduced into a
dry wood does not cause swelling and remains in cell lumens. When a nonswelling monomer or prepolymer is polymerized, the
polymer will occupy the cell cavities but not the cell walls [Schneider 1994, 2000].
2) In cell-wall, eg. furfuryl alcohol monomer or phenol-formaldehyde prepolymer. The monomer or prepolymer swells and penetrates
the cell-wall, resulting in a WPC with macropores plugged by a polymer in the cell wall. Furuno et al [1978, 1992a, b] found that the
polymer existing in the cell wall of a WPC has a greater influence upon the dimensional stability of the WPC, and is more important
than the polymer existing in the voids like the cell lumen.
3) Combination of cell-lumen and cell-wall types, eg. modified vinyl monomers such as ethyl -hydroxylmethylacrylate, or the
InduriteTM process.

1.2.5 Curing Methods for Wood Polymer Composites (WPC) Preparation


Different curing approaches can be performed to convert a monomer or a prepolymer impregnated in wood to a solid wood. These
techniques can be detailed as follows:
1) Radiation curing - commercial production of wood-polymer materials began in the mid-1960s using the radiation process.
Examples of such radiation are atomic particles, neutrons, photons, gamma rays, X-rays and electrons [Betty and Warren 1976].
Gamma radiation treatments appear to be the cheapest and most practical means of penetrating large sections of wood to obtain a
uniform distribution of the polymer. In general, ionizing radiation of above 5 to 10 megarads degrades wood, but a complete
polymerization of most vinyl monomers requires under 5 megarads [Siau et al 1965]. There is some indication that ionizing radiation
creates radical sites on the cellulose chain from which branched vinyl polymers can grow [Ramalingam et al 1963]. Parquet flooring
was the primary product, where the increased hardness and abrasion resistance were used to advantage in high-traffic commercial
installations. The long life and ease of maintenance justified the increased cost over normal flooring [Meyer 1982]. The main
drawback of the radiation method is the safety and environment concerns. The cost of transportation of the various wood products to
and from the irradiation site would be prohibitive [Meyer 1965].
2) Catalyst-heat curing - Meyer [1965] developed the catalyst-heat techniques to treat wood-polymer composites. Until the early of
1980s, the volume of production using the catalyst-heat process for making wood-polymer composites was smaller than with the
radiation process but its use is much more widespread throughout the United States and the world. The simplicity of the catalyst-heat
process and the low initial cost to begin production are attractive for techniques of the process to be used particularly by small
companies who make high cost, small volume items [Meyer 1982].
3) Micro-wave heating curing - as the heat transfer from the surface into the material is very low for wood, the heating rate of a
wood material is limited. Microwave radiation has been found to be efficient for curing treatment of an impregnated wood because
bulk heating was observed to be independent from the thermal conductivity of material [Galperin et al 1995]. Galperin et al. have
investigated heating behavior of wood impregnated with urea and formaldehyde resin solutions. It was found that microwave
treatment of birch samples with transverse dimensions of 25 42 (mmmm) took 2.5 min to heat the wood up to 100C, while it
took 32 min in the case of convective heating process.
4) Radio-frequency heating [Beall et al 1966] - radio-frequency heating was provided by a 30-mega watts generator, using a parallelplaten configuration. Radio-frequency heating was used to raise the temperature of the sample quickly, which then polymerized while
exposed to room temperature. Some researchers also used UV light to heat the impregnated wood.

1.2.6 Wood-Hardening Process


Efforts to improve wood properties, either by physical or chemical methods, have been done in the past. Knowledge and some
techniques for wood quality improvement have been established, which significantly advanced development of high performance,
value added wood products for various applications. It has been found that wood can be hardened by impregnation with reactive
chemicals, by compression, or by a combination of chemical impregnation and compression. Greater dimensional stability,
compressive strength and hardness of the modified wood were observed than in the original wood [Schneider 1994; Ayer et al 2003].
1.2.6.1 Impregnation

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

The hardness of wood can be greatly enhanced by filling the empty spaces in the cell tissue with a solid. The filling can be done by
impregnating the wood in a closed space with a liquid that can be hardened inside the wood, for example an impregnant monomer
fluid such as styrene or MMA that becomes solid by polymerization. One challenging problem is how to let the liquid penetrate evenly
into the cell tissue of the wood, since the permeability of most species of wood is poor and varies significantly in various parts of the
wood. For this reason, poorly impregnable wood is treated with an impregnation liquid in a closed space by using vacuum and/or
pressure. Another problem is how to let the impregnating substances be polymerized inside the wood so that the wood is not burned
or split, or without deteriorating wood machine ability.
Generally, reactive substances such as monomers or prepolymers used for the impregnation of wood are polymerized by heatwith
addition of suitable polymerization initiations. The reaction is exothermal and usually results in wood materials and wood-based
materials harmful splitting, changes in the dimensions, or burning. The temperature within the wood rises to a very high level owing
to the reaction. Thus thermal plates, which serve to initiate the reaction and also to even out the heat of reaction and cool the
reaction, can be applied. This method, however, is space consuming and costly. Another method of controlling the polymerization
reactions is to use radiation to initiate the polymerization. Pieces of wood material are, in this case, treated, for example, by gamma
radiation. The commercial application of this method on an industrial scale is expensive, and its implementation for sheet-like pieces
of wood requires a larger apparatus configuration, thus increasing for example occupational safety risks [Paavo and Markku 1994].
The type of wood, its anatomic structure, composition, and density, and the viscosity, chemical structure, and polarity of the monomer
are important factors in the impregnation of monomers into wood.
1.2.6.1.1 Monomer and Polymer Treatments

When the cell lumens of wood are filled with a suitable liquid monomer or prepolymer, which is subsequently polymerized, the
resulting wood-polymer material normally retains most of the desirable characteristics of wood and displays many improved
properties. Most wood species can be readily impregnated with various vinyl monomers or prepolymers by using equipment and
techniques similar to those used in the conventional wood treating industry. There is, however, a significant difference in the
viscosities encountered; vinyl monomers have viscosity values considerably lower [Table 1.1, Loos et al 1967].
Vinyl Monomers
Treatment of wood with vinyl polymers is not only advantageous for preservation but also for mechanical properties and water
repellency improvement. Treatment with vinyl type monomer followed by curing significantly improved the moisture resistance,
hardness, etc. of wood [Meyer 1981]. Vinyl monomers can be polymerized by using catalysts [Meyer 1965] or radiation techniques
[Siau et al 1965]. Langwig et al [1968] polymerized t-butylstyrene, epoxy monomers, and methyl methacrylate in wood by in situ
impregnation. Hardness and static bending strength increased but mechanical properties showed little change. It is desirable that the
vinyl polymer would effectively stabilize the wood by bulking the cell wall (Timmons et al 1971), but vinyl monomers are, generally,
poor swelling agents for wood with the exception of acrylonitrile [Siau 1969, Loos and Robinson 1968]. Ellwood et al[1972]
investigated the dimensional stabilization of wood with vinyl monomers. Impregration with four different monomers followed by
radiation curing produced evidence of sufficient dimensional stabilization. Of the four monomers tested, namely styrene/acrylonitrile
(60/40), methyl methacrylate/phosgard (88/12) and ethyl acrylate/acrylonitrile (80/20), only styrene-acrylonitrile gave antiswell
efficiencies comparable with those obtained with originally water soluble polymers that bulk the cell walls [Ellwood et al 1972].
Because vinyl monomers enter the cell walls only to a limited extent without an additional swelling agent present, they are not
expected to impart any substantial dimensional stabilization to wood. The impregnation of wood with acrylic or vinyl type monomer
showed less dimensional stability in the presence of moisture. This was due to the confinement of the monomer in the cell lumen
instead of the cell wall [Rowell and Ellis 1978]. Chemically treating woods with styrene [Raff et al 1965, Langwig et al 1968, 1969,
Brebner et al 1988, Khan et al 1992a, b, Chao and Lee 2003], or methyl methacrylate is expected to be better at improving wood
properties.
Styrene (ST)
Wood impregnated with styrene has been largely investigated, and showed improved water repellency, compression and bending
strength [Baki et al 1993]. Siau et al[1965] studied wood-styrene combinations using radiation techniques. This investigation was
primarily concerned with bulking wood with styrene, followed by polymerization with gamma irradiation. The dimensional stability of
wood was improved. But an initial swelling of the wood by water or some other polar solvent was required and was considered to be
necessary. Impregnation was done by solvent exchange and high-vacuum methods. It was found that gamma irradiation produced
graft copolymers in the wood that were more effective than homopolymer in improving dimensional stability and tangential tensile
strength. Antishrink efficiencies as high as 80 percent were achieved with polymer retentions from 0.8 to 2.0 times wood weight
(oven-dry) [Siau et al 1965].

Table 1.1 Comparison of viscosities of liquids used in preservative treatmetn compared to those of several vinyl
monomers [Loos et al 1967].

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

Brebner et al [1985] investigated eastern white pine heartwood impregnated with a mixture of styrene monomer and cross-linking
agent. The wood-polymer composites was then produced using a heat-catalyst technique. The flexural properties of the combination
were measured in the oven-dry and air-dry moisture content conditions. Strength increases were greater than those predicted by the
rule of mixtures. Styrene uptake was dependent on the growth rate of the individual wood samples, but the strength did not suffer in
comparison with more uniformly-loaded, low density woods. At air-dry moisture content and equal density, the pine-polystyrene
combination was found to be mechanically inferior to sugar maple.
Devi et al[2003] investigated chemical modification of rubber wood (Hevea Brasiliensis) by impregnating the wood with styrene and in
combination with a crosslinker glycidyl methacrylate (GMA). Polymerization was carried out by catalyst heat treatment. The
dimensional stability and anti-shrink efficiency were determined and found to be improved by the treatment. Water absorption was
also found to be decreased considerably for the treated wood samples. Mechanical strength of the treated samples, in terms of
modulus of rupture and modulus of elasticity, was also found to be improved. An anti-shrink efficiency value of 53 % was obtained for
styrene-GMA treated samples compared to that of 23 % for styrene treated wood samples for 24 hours in water. Hardness was almost
the same for the styrene treated and styrene-GMA-treated samples and was increased by 33% over that of the untreated samples.
Chao and Lee [2003] investigated the improvement in hardness of southern pine wood by styrene impregnation. Regular diffusion
and vacuum impregnation were used to transport the styrene into the wood, and styrene was then polymerized in situ by a
catalyst-heat process. The hardness, density, weight gain and water absorption of treated Southern Pine samples were determined.
Table 1.2 gives the results of the properties of the styrene-treated and untreated pine. The density, water repellence, and hardness of
southern pine increased after impregnation with styrene (Table 1.2). The styrene gradually penetrated into the wood during the
soaking process and resulted in a linear increase of density following polymerization and a decrease of water absorption as the
treatment time increased from 0 to 8 days. The hardness of wood on both the R-L and the T-L surfaces did not show a significant
improvement after the soaking treatment. On the other hand, the vacuum process facilitated absorption of a greater amount of
styrene into the wood in a much shorter time. A greater density increase and less water absorption were observed on the treated
wood with only one minute of vacuum. The hardness of the wood improves significantly by more than two-fold after one minute of
the vacuum process.

Table 1.2 Properties of styrene-treated and untreated southern pine [Chao and Lee 2003].

Methyl Methacrylate (MMA)


Methyl methacrylate (MMA) is one of the most common vinyl monomers used to make WPC. MMA is an inexpensive, industrial, clearcolored, and commodity chemical, which can have dyes added. It is mainly used to produce clear and colored WPC. MMA swells wood
little [Loos and Robinson 1968], thus producing WPC with polymer-filled cell lumens but cell walls that are virtually unchanged. This
type of WPC has time-dependent dimensional stability improvement, in the long term it might be the same as that of untreated wood.
Over the years, MMA has been the most extensively used monomer in making wood-polymer composites [Siau and Meyer 1966,
Langwig et al 1968, 1969, Siau et al 1968, Duran and Meyer 1972, Siau et al 1978, Noah and Foudjet 1988, Yalinkilic et al
1998].Duran and Meyer [1972] investigated the amount of exothermic heat released during catalytic polymerization of basswoodmethyl methacrylate composites. It was found that increasing the concentration of Vazo (azobisisobutyronitrile, a free radical
catalyst) in the basswood-MMA impregnated samples reduces the time to reach the maximum of the exothermic peak and increases

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

the exothermic peak temperature, but decreases the conversion rate of the monomer to polymer. The use of a crosslinker rapidly
increases the initial viscosity of the system, and an auto-acceleration of the polymerization is observed, e.g. the crosslinker, trimethyl
propane trimethacrylate (TMPTMA), reduces the time to the exothermic peak and dramatically increases the exothermic peak
temperature. The effect of the crosslinking agent on the resulting wood mechanical properties is however uncertain.
Siau and Meyer [1966] compared the properties of heat and radiation cured wood-MMA polymer composites. No significant
differences were found between heat and radiation-polymerized yellow birch-polymethyl methacrylate combinations when tested for
compression perpendicular to grain direction, shear strength, permeability, diffusion coefficient, and antishrink efficiency. The surface
hardness of the radiation polymerized material was about 25 percent greater than that for the heat-polymerized combination,
possibly due to the preferential evaporation of monomer from near the surface during polymerization. The permeability of treated
and untreated wood samples was apparently increased when the cobalt 60 radiation exposure was increased from 3 to 10 megarads.
No apparent mechanical degradation was observed up to 10 megarads of exposure to gamma radiation. Siau et al[1968] developed a
geometrical model for wood polymer composites. Basswood was impregnated in the dry condition with MMA and styrene, which
were then polymerized in situ by the catalyst-heat technique. The composites and constituent materials were tested in compression
parallel to grain direction, static bending strength, dynamic modulus of elasticity, and dielectric constant. Theoretical equations for
these properties were derived from an assumed cellular model. A good agreement between theoretical and experimental values
indicates that the model may be a useful tool for the prediction of properties of a composite from those of the constituent materials.
The evidence of an interaction between vinyl monomers and the cell wall was found for the wood-polymer composite materials. When
wood is impregnated with MMA and polymerized by heat, dimensional distortion and volumetric changes are observed [Siau 1969].
The distortion of basswood is much more severe than that of maple. Siau [1969] also investigated the swelling behavior of basswood
impregnated with MMA and styrene. Antishrink efficiency of up to 40 percent indicates the entry of monomer into the cell wall before
polymerization. The rate of swelling by monomers increases with temperature and moisture content of the wood. The amount of
swelling at equilibrium generally decreases with temperature. Comparisons with other studies indicate a large difference in swelling
properties between wood species.
Beall et al[1966] investigated the direct and radio-frequency heat curing of basswood plastic composites impregnated with methyl
methacrylate containing a peroxide initiator. Three types and two concentrations of organic peroxides were used for the initiation of
the polymerization of methyl methacrylate. Heating the wood-monomer composite by radio-frequency energy to a temperature
between 90C and 98C caused a rate of free radical generation sufficient to overcome the inhibitor concentration and to promote
polymerization in about 20 percent of the time required for direct heating. The polymer retention for radio-frequency heating was
approximately 80 percent of that for direct heating. Most of the monomer loss could be accounted for in the handling techniques used
in radio-frequency. Young and Meyer [1968] compared the properties of heartwood and sapwood before and after impregnations with
MMA monomer and polymerization using the catalyst-heat method. Vazo as a free radical catalyst was used. Eight species of wood
including yellow birch and red maple were investigated. After treatment, the sapwood exhibited greater increases in compressive
strength perpendicular to the grain direction, tangential hardness, and density than did the heartwood. Sapwood also showed a
greater reduction in permeability and it had a higher polymer retention, which meant that more of the calculated void volume was
filled with polymer (Table 1.3). Beall et al[1973] compared the hardness and hardness modulus of wood-polymer composites that
were impregnated with MMA of red oak, aspen, and hard maple. It was found that the hardness of untreated wood was mainly related
to the sample density. For the treated wood, the hardness was significantly related to the hardness modulus, density and polymer
loading.

Table 1.3 Summary of tangential hardness and air permeability data [Young and Meyer 1968].

S=sapwood, H=heartwood, 1 standard deviation from 6 measurements.


Schneider [1995] proposed a method to prepare cell wall and cell lumen wood polymer composites by using two monomers, one of
which can swell the cell wall. Furfuryl alcohol-based (FA) cell wall and methyl methacrylate-based (MMA) combination formulations
were used. Swelling gradients developed during preparation. Properties (density, hardness, water extractables, antiswelling efficiency)

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

gradients were observed along samples in 5 of the 7 treating formulations. Two cell wall treatments based on furfuryl alcohol did not
show the gradients. The results suggest treating solution separations occurred during impregnation and the resulting
nonhomogeneous chemical yielded the properties gradients. WPC made using the combination formulations based on MMA had
properties that fell between cell wall and cell lumen formulations.
A combination of wood cell-wall modification with acetic anhydride and lumen filled with methyl methacrylate (MMA) treatments were
studied for their effectiveness in reducing the rate of moisture sorption and the degradative effects of accelerated weathering [Feist et
al 1991]. It was found that the combined treatment of acetylation followed by methacrylate impregnation was the most effective in
reducing the rate and extent of swelling and reducing erosion caused by accelerated weathering (85%).Both acetylation and
methacrylate treatments,or a combination of the two, reduced the loss of surface lignin with subsequent reduction in weathering. It is
obvious that treatment of wood with methacrylate polymers decreased the rate of moisture uptake and improved mechanical
properties, including modulus of elasticity and rupture, fiber stress at proportional limit, work to maximum load, maximum crushing
strength, and hardness index, compared to those properties of untreated wood [Langwig et al 1968, Rowell et al 1982]. However,
treatment with metharylate polymers causes high polymer weight gain and is more expensive.
Vinyl Chloride
The study of impregnation and polymerization of vinyl chloride in wood was carried out under a wide variety of conditions as
described in the article of Juneja and Hodgins [1970]. The polymerization reaction is highly exothermic, which causes distortion and
discoloration of the substrate. Degradation of the PVC releases hydrogen chloride, leaving behind conjugated double bonds along the
chain, which are responsible for the polymer discoloration. Thus, vinyl chloride was considered unsuitable as a monomer for
producing acceptable WPC material mainly because of the following reasons: 1) the polymer is a mixture of powder and glassy
polyvinyl chloride; 2) mechanical properties are only slightly improved; 3) heats of reaction are so high that discoloration and
deformation occur; and 4) polyvinyl chloride is insoluble in the monomer and precipitates as a fine powder during the reaction. This
powdery form offers no contribution to the enhancement of the mechanical properties. Also PVC generally has a bad environmental
image.
Hydroxymethylacrylate (HMA) and Ethyl- - Hydroxymethylacrylate (EHMA)
Mathias and Wright [1989] developed wood-polymer composites by ways of impregnation and in situ

polymerization

of

hydroxymethylacrylates. It was observed that ethyl alpha-hydroxymethylacrylate (EHMA) is a good monomer for use in polymer-wood
composites. The monomer as reported can penetrate the cell walls of the wood fibers, thus forming a completely filled polymer-wood
composite. Wright and Mathias [1993] also developed a lightweight materials based on balsa wood-polymer composed with the
combination of ethyl -(hydroxylmethyl)acrylate (EHMA) and styrene. The copolymer EHMA-styrene improved the dimensional
stability (water soak test) and mechanical properties of balsa wood. Improvements in specific modulus and specific toughness
(absolute properties divided by specific gravity) were also achieved using an EHMA-styrene monomer mixture with polybutadiene
diacrylate as cross-linker and toughening agent. The best results were obtained at low weight gain (10-40%). These improvements in
modulus and toughness were ascribed to efficient penetration of monomers into the cell walls. The combination of a light weight,
renewable wood precursor with synergistic reinforcement by this combination of monomers offers unique opportunities for increased
use of wood-polymer composites in a wide variety of structural and insulating applications. Table 1.4 shows their results obtained
from the EHMA-styrene-modified wood composites and some other polymer modified wood data from the literature [Wright and
Mathias 1993].
Water dispersion phenolic resin can diffuse into the cell walls and swell the wood. After that, the water can be removed by drying the
wood at low temperature, and then the resin is cured by heating the impregnated wood at curing temperature. Bryant [1966]
investigated the effect of phenolic resin impregnation on mechanical properties of wood and the effect of relative molecular size of
phenolic resin and concentration on the dimensional stabilization of wood. Water-soluble phenolic resins were used to impregnate
wood.

Table 1.4 Comparison of the measured properties and the specific properties (Measured divided by density) from the
literature data vs. the data for balsa impregnated with EHMA-ST-pBDDA [Wright and Matias 1993].

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

=(Valuetreated Valuecontrol)/Valuecontrol 100


Properties such as the compressive strength and moisture-related shrinking and swelling were improved (Tables 1.5, and 1.6).
Phenolic resins penetrate and bulk the cell wall structure, preventing shrinkage of the wood upon drying. Chui et al[1994]
investigated the effects of phenol resin impregnation and process parameters on some properties of poplar laminated veneer lumber
(LVL). Compared with untreated LVL, mechanical properties and dimensional stability are improved when all veneers or surface
veneers are impregnated with phenol formaldehyde (PF) resin. Increasing press pressure, reducing time to maximum pressure, and
increasing initial veneer moisture content lead to greater compression of LVL boards. Increased compression of boards directly leads
to an increase in modulus of rupture, modulus of elasticity and thickness swelling. It should be noted that all phenolic resin modified
wood products are dark brown, thus their applications are limited.

Table 1.5 Effect of relative molecular size of phenolic resin and concentration on the dimensional stabilization of wood
[Bryant 1966].

1 solid

content with weight percentage.

2 swelling

percentage is the volumen percentage

Polyurethane
Hartman [1969] used aqueous polyurethane emulsion to modify wood because of its availability of the polyurethane as low molecular
weight components in aqueous media. Four kinds of wood were impregnated with different molecular polyurethanes. The aqueous
polyurethane systems could be in a high solid concentration at low viscosity and are cured readily at elevated temperature, and could
be used over and over without effecting polymerization at room temperature. It was found that lower molecular weight resin has
better diffusion and penetration into the wood [Hartman 1969]. Khan and Idriss [1992] investigated the effect of urea on the
mechanical strength of wood-plastic composites. It was suggested that the use of urea in the wood plastic composite formation open
up a wide scope of preparative technique of WPC when high polymer loading along with strong tensile strength is considered.
Melamine-(Urea)-Formaldehyde
Polymers of melamine (1,3,5-triamino-2,4,6-triazine) and formaldehyde form an important class of amino resins, which have been
commercially used for over 60 years. Melamine-formaldehyde (MF) itself is one of the hardest and stiffest isotropic polymeric
materials used in decorative laminates, moulding compounds, adhesives, coatings and other products. Due to its advantageous
properties, i.e. high hardness and stiffness, and low flammability, MF resins have potential to improve properties of solid wood.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

Impregnation of solid wood with water-soluble MF resin has led to a significant improvement of surface hardness and modulus of
elasticity (MOE) [Inoue et al 1993c, Miroy et al 1995,Deka and Saikia 2000].

Table 1.6 The effect of phenolic resin impregnation on mechanical properties of wood [Bryant 1966].

All resins were approximately at pH 10 in 95 percent ethanal.

Based on original dimension of untreated yellow poplar specimens at 12 percent moisture content. Specimens were 1/10th-inch

thick, 3-inches in radial width, and 1-inch along the grain.


By impregnating wood with methoxymethyl melamine, which is not capable of reacting with itself, but can react with hydroxy groups
in wood, an increase of Brinell hardness was achieved [Miroy et al 1995]. Aqueous melamine-formaldehyde (MF) resins can penetrate
into the wood cell wall [Rapp et al 1999, Gindl et al 2002, Gindl and Gupta 2002] and the amorphous region of cellulose fibrils [Hua
et al 1987a,b]. The influence of melamine-formaldehyde impregnation on the strength and stiffness of spruce in transverse
compression was investigated [Gindl et al 2003a]. After polymerization, the tangential compression strength of treated samples was
83 % higher compared to the untreated reference, and radial compression strength had increased by 290 %. Less than 2 % of the
cell cavities (lumina) were found to be filled with resin. Thus the improvement of strength and stiffness obtained is attributed to
modification of the cell wall and not to filing of tracheid lumina. Yielding-behavior under excessive compression load changed from
plastic buckling in the reference samples to brittle fracture of cell walls in the treated samples. Gindl and Gupta [2002] investigated
the cell-wall hardness and Youngs modulus of melamine-modified spruce wood by nano-indentation. It was demonstrated that
modification of wood with melamine-formaldehyde resin changes cell wall mechanical properties. Gindl et al[2003] also examined the
selected factors influencing the uptake of MF resin into the cell wall of softwood. Using UV-microspectroscopy, it was found that water
soluble MF diffused well into the cell wall. Moisture content, high water content of the resin used for impregnation, and low extractive
content are factors, which are favorable for MF resin uptake into the cell wall. For dry cell walls, solvent exchange drying improved
resin uptake to a similar extent, as was the case when cell walls were soaked in water.
Poly(ethylene glycols) PEGs
Poly(ethylene glycols) are widely used in the field of the dimensional stabilization of wood [Stamm 1959, 1964, Hoffmann 1988,
1990]. One classical approach is to fill wet wood with polyethylene glycol (PEG), a water-soluble synthetic wax. The wood is put in hot
PEG solutions of increasing strength, and the PEG diffuses into the wood replacing the water. On cooling to room temperature the
PEG solidifies, and thus hinders the structure from shrinkage and collapse. Hoffmann [1988, 1990] investigated the low molecular
PEG impregnation waterlogged softwoods and oakwood. It was found that to reach optimal stabilization less PEG per wood substance
is needed for the softwoods than for oakwood. A 20 % solution of PEG 400, and a 50 % solution of PEG 4000 will bring the required
amounts of PEG into the wood compared to 50 % and 70 % solutions for oak wood. Strict relations exist between PEG uptake for
optimal stabilization and the density of the wood. Optimally treated wood does not become damp in RH lower than 86 %.
Impregnation with PEG aqueous solutions allows the bulking of wood cell walls, preventing shrinkage, but leads only to the deposition
of chemicals in the wood that are nonbonded and consequently easily leachable. Soulounganga et al [2003] developed the synthesis
of poly(glycerol methacrylate) (PGMA) and applied to dimensional stabilization of wood. PGMA was synthesized in good yields starting
from polyglycerol (PG) and glycerol methacrylate (GMA). Since soluble in water, PGMA allows waterborne treatment of wood, limiting
the use of volatile organic compounds. After in situ polymerization, the treated wood shows increased dimensional stability compared
to the untreated wood. The dimensional stabilization and the level of polymerization are in direct relation with the amount of PGMA
used for

the impregnation

solution. Microscopic observation

coupled with

picture analysis indicates that

poly(glycerol-

poly(methacrylate)) is able to bulk the cell walls, explaining the observed dimensional stabilization. Since polyglycerols are
polyethers presenting several similarities with polyethyleneglycols, they possess numerous hydroxyl groups, which permit to form
tridimensional network and decrease their leachability. Similarly, for enhancing wood durability, chemical modification of wood by
polyglycerol (PG)/maleic anhydride (MA) treatment was investigated by Roussel and coworkers [2001]. Conditions recommended for
producing PG/MA treated wood durable for a long time are vacuum/pressure impregnation of aqueous solutions of PG/MA adduct
(30%) in the presence of 2-butanone peroxide (2%) and cobalt naphthenate (2%).
Alkylene Oxide Treatments

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

Alkylene oxide treatments can change the hydrophilic nature of wood, which have been used on the dimensional stability of wood by
many researchers [Rowell 1975, Rowell et al 1976, Guevara and Moslemi 1984]. Rowell et al[1982] investigated the combination
treatment of cell wall grafting with alkylene oxides and lumen treatments with methyl methacrylate to prepare wood-polymer
composites. Using a two-stage treatment, they chemically modified the cell walls of southern pine and sugar maple with propylene
oxide and then impregnated the void volumes with methyl methacrylate, resulting in highly stabilized wood-polymer combinations.
While propylene oxide was effective in reducing the wood hygroscopicity and bulking the cell wall, it did not necessarily provide a
good adhering surface for the poly(methyl methacrylate). The propylene oxide treatment reduced the flexural strength of the wood.
The affected properties were largely restored, however, by filling the lumens with the polymer. A desired pretreatment is one that
reduces wood hygroscopicity, but not the wood strength, and also promotes bonding with vinyl polymers. Schneider and Brebner
[1985] investigated the combination of chemical modification of wood with alkoxysilane coupling agents. Many other cell-wallmodifying chemicals could be studied along with other lumen-polymer-fill treatment.
1.2.6.1.2 Other Treatments

Inorganic Wood Composites


It is noteworthy that wood-inorganic composites, in which the woods cell lumens and/or cell-walls are filled with inorganic
substances, have been developed recently [Furuno et al 1991, 1992a,b, 1993, Saka et al 1992, Ogiso and Saka 1993, Saka and
Yakake 1993, Yamaguchi 1994a,b, Zollfrank and Wegener 2002]. One obvious improvement in wood properties for inorganic-wood
composites is the fire and decay resistance of the wood. Silicified wood is a type of fossil produced by the replacement of some or all
of the wood substances with SiO2 mineral (silica) during burial over extremely long periods, and it is almost petrified.
A combination of acetylated and propionylated wood-silicate composites was investigated by Li et al[2001]. The acetylated
wood-silicate composites and propionylated wood-silicate were evaluated, impregnated with an aqueous sodium silicate solution. The
resulting composites showed good dimensional stability. The silicate gels fixing in composites endowed the composites with flameresistance, while the oxygen index of the composites increased with an increase in weight percent gain of silicate gels.
Yalinkilic et al[1998] developed a new method in which boron treatment was combined with vinyl polymerization to improve leaching
resistance of boron from wood, as well as dimensional stability, biological and thermal resistance of wood. Boric acid (BA) was
impregnated into wood specimens (sapwood of Cryptomeria japonica) as 1% aqueous solution prior to vinyl monomer treatment.
Styrene, methyl metacrylate and their mixture (50:50, v/v) were impregnated in the presence of catalyzer and crosslinker.
Polymerization was conducted by heat radiation method at 90C for 4h. Treated specimens were then subjected to decay and termite
tests, as well as oxygen index determination. Anti-swelling efficiency and water absorption levels were also measured. Vinyl
monomers succeeded in reducing water absorption of wood to minimum level and delaying boron leaching considerably. The treated
wood proved to be resistant against two decay fungi. It is well known that vinyl monomers contribute to dimensional stability and
strength properties of wood, and they are expected to provide a delayed leaching of boron from wood. Boron is considered to enhance
the biological and fire resistance of vinyl polymerized wood in a combination treatment system as vinyl monomers are not toxic to
fungi after polymerization.
Combination of two or three monomers
Schneider [1995] investigated cell wall and cell lumen wood polymer composites. Furfuryl alcohol-based (FA) cell wall and methyl
methacrylate-based (MMA) combination formulations were used to make wood polymer composites. A combination of wood cell-wall
modification with acetic anhydride and lumen fill with methyl methacrylate (MMA) treatments were studied for their effectiveness in
reducing the rate of moisture sorption and the degradative effects of accelerated weathering [Feist et al 1991]. It was found that the
combined treatment of acetylation followed by methacrylate impregnation was the most effective in reducing the rate and extent of
swelling and reducing erosion caused by accelerated weathering (85%). Both acetylation and methacrylate treatments, or a
combination of the two, reduced the loss of surface lignin with subsequent reduction in weathering.
Gaylord [1973]developed a method for impregnating wood in situ with styrene and maleic anhydride monomers that formed a 1:1
charge transfer complex. At low temperatures, heating produced uncatalyzed polymerization. Forest Products Laboratory researchers
have been investigating bonding of fluorophenyl isocyanates to wood cell wall polymers and reaction of 8-hydroxyquinoline,
pentachlorophenol, or -naphthol with ethylene-maleic anhydride copolymer to cell wall polymers [Chen and Rowell 1986]. Their goal
is to lower chemical load by using bonded biocides as controlled-release compounds. However, filling cell lumens with preformed
bioactive polymer leads to problems with solubility and penetration because of the reason that generally polymers have low solubility,
high viscosity, and thus to diffuse with difficulty into wood. Therefore, it is preferable to synthesize a monomer with a bioactive group,
fill the wood with monomer, and then in situ polymerize or copolymerize with a carrier monomer using a catalyst. This will result in
higher loading of the bioactive polymer into the wood.
A combination of styrene-acrylonitrile wood composite was developed by Juneja and Hodgins [1970]. It was found that the styreneacrylonitrile-wood composite had properties that compare favorably with the best materials produced by radiation polymerization.
Ellis and ODell [1999] developed wood-composites made with acrylic monomers, isocyanate, and maleic anhydride combination. Pine,
maple, and oak solid wood were combined with different combinations of hexanediol diacrylate, hydroxyethyl methacrylate,
hexamethylene diisocyanate, and maleid anhydri. It was found that treatment slowed the rates of water vapor and liquid water
absorption. Although the resultant dimensional stability was not permanent, the rate of swelling of WPC specimens was less than that
of unmodified wood specimens. In general, WPC prepared using hydroxyethyl methacrylate were harder than specimens made
without hydroxyethyl methacrylate and excluded water and moisture more effectively.
Using swelling solvents to dilute vinyl monomers and allow the monomer to enter cell walls is one approach, which has been used to
form vinyl WPC with moisture-resistant cell walls [Furono and Goto 1973]. Other approaches have been performed to use polar vinyl

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

monomer molecules or additives mixed with the monomer to cause wood swelling [Loos and Robinson 1968, Rowell et al 1982,
Schneider and Brebner 1985]. Such combination treatments (treating cell lumens and walls) show promise for producing WPC with
the advantages of vinyl types-wood composite but improved permanent dimensional stability.
Wax oil treatment
Wax oil can be employed for impregnation of wood for moisture proofing. Both natural and synthetic wax can be used. But generally
the resulting surface appears greasy and dirty. (WO0139917).
1.2.6.2 Compression
Wood can be largely compressed transversely under heat with little damage to its cell wall structure, and the deformation imposed is
fixed by drying under restraint. The compressed wood, because of its higher strength properties, has a potential for use in flooring
board, furniture, interior materials, engineered composite materials, surface-hardened softwoods, and so on [Dwianto et al 1999].
However, the fixation of deformation is apparent because it recovers its original thickness almost completely following application of
moisture and heat [Norimoto and Gril 1989]. To fix the deformation permanently, chemical treatments of compressed wood such as
the formation of cross linking between wood components [Inoue et al 1994], the substitution of hydroxyl groups by hydrophobic
groups, and resin impregnation into the wood cell wall [Inoue et al 1993b] have been attempted. Heating or steaming provides an
additional mode of flexibility in the permanent fixation [Inoue et al 1993a]. For example, the deformation is permanently fixed by
heating under restraint at 180C for 20h. This has been considered to be due mainly to the release of stresses stored in the cell wall
polymers following their decomposition [Dwianto et al 1997]. Heat treatment seems up to date from the standpoint of environmental
protection, since no chemical reagent is needed in the process, though the treatment is lengthy. In contrast to heating (dry heating),
high temperature heating in the presence of water (high temperature steaming, or steam heating) gives quite different results, even
at the same temperature. Compressed wood steamed at 180C is fixed completely within 8 minutes and the resulting wood showed
little lowering of modulus of elasticity (MOE) and modulus of rupture (MOR) [Inoue et al 1993a].
Researchers are searching for an effective method to stabilize the dimension of wood during changes in moisture content. Many
methods have been studied, ranging from bulking the cell wall with simple aqueous sugar solutions to chemical reactions with the
cell-wall polymers. Most methods use chemicals that bulk, react, crosslink, or coexist within the wood cell wall. Theoretically, wood is
stabilized by decreasing its hygroscopicity after thermally degrading the hemicelluloses (the most hygroscopic polymers in the cell
wall). It is observed that dimensional stability can be improved by either steaming or heating wood, while the wood is in a
compressed state. The effect of steam or heat on fixation of compression set and the effect of these treatments on hardness,
mechanical properties, and color of compressed and uncompressed wood specimens were investigated. Burmester [1973] found that
under optimum moisture content, for the heat and pressure treated wood, deformation caused by moisture swelling was reduced 75
% for oak heartwood, 60 % for beech sapwood, 55 % for pine sapwood and heartwood, and 52 % for spruce. Giebeler [1983] found
that swelling decreases of 50 % to 80 % when beech, birch, poplar, pine, and spruce were heated at 180C to 200C in an inert gas
atmosphere of 0.8 to 1.0 MPa. Inoue et al[1993a] found that the specific gravity increased from 0.36 to 0.50 with a compression set
of 30 %. At a compression set of 60 %, specific gravity increased to 0.9g/cm3, which was more than twice the initial specific gravity.
The fixation of the compressed wood structure by steaming is probably a result of the steam softening of the lignin-hemicellulose
matrix in which the crystalline cellulose microfibrils are embedded. Tanahashi et al[1989] showed that the steaming of wood results
in an increase in cellulose crystallinity, microfibril width, and micelle width. Inoue et al[1993a] found that almost complete fixation of
compression set in wood can be achieved by steaming wood while compressing. Compressed wood steamed at 180C for 8 min or at
200C for 1 min showed no recovery of set, large increases in hardness, minimum decreases in mechanical properties, and slight
darkening. Dry heating of compressed wood for lengthy periods can also give complete fixation of compressive set but with significant
decreases in mechanical properties and severe darkening.
Methods to enhance wood hardness and dimensional stability include various combinations of compression, heating, and chemical
treatment. Inoue et al[1993 b] investigated wood chemical treatment then compression, which was treated with increasing
concentrations of a low molecular weight, water-soluble melamine-formaldehyde resin solution (MW, 380) and compressed while
heated. This method achieved a maximum bulking efficiency of 5 % and an anti-shrink efficiency of 45 %, showing that the chemical
had not completely penetrated the cell wall. Once the wood was treated, its ability to retain the compressed state was tested by
immersing wood specimens in water at different temperatures. Specimens treated with an 8 % resin solution retained almost
complete fixation when soaked in room-temperature water, while those treated with a 25 % solution retained fixation in boiling
water. A solution of 25 % resin and a compression of 54 % increased hardness from 0.48 to 0.72 MPa. According to these reports,
size stability is achieved not by the dissolution of hemicellulose but by heat plasticity of matrix, i.e. cutting hemicellulose and part of
lignin molecules, which are rearranged under compressed condition. It is generally agreed that fixation of compressed wood structure
by steam treatment is a result of steam softening of lignin-hemicellulose matrix.
Wellons et al[1983]examined the effects of veneer moisture content (1% to 6%), press temperature (132C to 165C) and pressing
cycles on compression of Douglas fir, southern yellow pine and hem-fir plywood panels. A maximum compression of 11% was
obtained when a steady pressure of 1380 kPa was applied, and when both veneer moisture content and press temperature were at
their maximum value. Microscopic examination revealed that when compression was less than about 5%, the glue line was the major
zone of compression. At higher compression the thin walled fiber of the early wood near the glue lines collapsed. Zhang et al[1994]
investigated the processing for hot-pressing of poplar laminate veneer lumber (LVL) in order to control the compression effectively. It
is observed that compression has significant effects on modulus of elasticity, modulus of rupture, specific gravity and thickness
swelling of poplar LVL. Modulus of elasticity, modulus of rupture and specific gravity appear to be directly proportional to compression
within the compression range of 5 % to 20 %. Horizontal shear strength results indicate that, due to inadequate contact, proper glue
bond may not be achieved between veneers of LVL with low compression. Thickness swelling appears not sensitive to compression
between the compression range of 4 % to 10 %.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

Table 1.7 Compression set (C2) and set-recovery (R2) of compressively molded squares [Ito et al 1998a].

Ito et al[1998a] presented a new method of processing logs into square lumbers without cutting them but by treating them with high
pressure steaming in an autoclave, for the purpose of utilizing coniferous thinning. They attempted to compression-mold logs of sugi
(Cryptomeria japonica D. Don) to squares, and to permanently fix the transformed shape. The transformation and fixation were
successively conducted under high-temperature saturated steam atmosphere by a newly designed apparatus that is a combined
pressure vessel and a press machine. The results are summarized as follows: 1) the logs were successively softened by steaming at
200C in the apparatus; 2) the shape of square fixed by high-pressure steam-treatment did not recover to original log shape by
water-soaking and boiling. The result shows that the high-pressure steaming is effective for permanent fixation of the transformed
shape; 3) the condition for sufficient fixation of the compressive transformation was steaming at 200C for 3 min when a 15cm of
diameter and 10cm of length sugi-thinning was used as a specimen; 4) internal stress within a compressed wood was gradually
reduced during steaming and, finally, the stress was removed completely. The relationship between the steam temperature and
compression set is given in Table 1.7. It is obvious that the value of set-recovery (R2) decreases with an increase in fixation
temperature. It is also evident that steaming at 200C for 3 min is required for the complete fixation of 15 cm diameter and 10 cm
length logs.
Ito et al[1998b] investigated the mechanism of permanent fixation of the compressive molding of wood by high-pressure steam
treatment. The conditions under which the fixed specimen would not recover to their original shape were found to be steaming with
saturated steam at 200C for 4 min or longer or at 180C for 8 min or longer time. The mechanism of fixation of compressive
transformation is supposed to be that the internal stress is released because the paracrystalline region of cellulose, which is distorted
by compressive transformation, is partially hydrolyzed. On the other hand, steam-rearrangement of the hydrolyzed constituent into
crystalline region may occur, thus keeping the transformed shape intact. Dwianto et al[1999] clarified the mechanism of the
permanent fixation of compressive deformation of wood by high temperature steaming, stress relaxation and stress-strain
relationship in the radial compression for Sugi (Cryptomeria japonica D. Don) wood under steam at temperatures up to 200C. The
stress relaxation curves above 100C were quite different in shape from those blow 100C, showing a rapid decrease in stress with
increasing temperature. In the stress-strain relationships measured above 140C, the stress reduced as pre-steaming time increased
with steaming time and reached almost 0 in 10 min at 200C. The relationship between the residual stress and the stain recovery at
the end of relaxation measurements could be expressed by a single curve regardless of time and temperature. The permanent
fixation of deformation by steaming below 200C was considered to be due to chain scission of hemicelluloses with a slight cleavage
of lignin. In some cases, the increase in regularity of the crystalline lattice space of microfibrils or the formation of crosslinks between
the cell wall polymers seemed to play an important role in the permanent fixation of compressive deformation.
1.2.6.3 Chemical Impregnation and Compression of Wood
The hardening of wood has been widely investigated for value added applications since the introduction of phenol-formaldehyde (PF)
impregnated compressed (Compreg) high-density wood for manufacturing airplane propellers during the early 1940s. Today,
compregnated wood has established itself for a variety of industrial applications. Products such as bolts, rivets, gears, high and low
tension circuit breakers, turbo generators, sports goods, chair seats, etc. are being manufactured.
Vinyl monomers
Yalinkilic et al[1998] have explored compressed-wood polymer composites (CWPC) by in situ polymerization of the impregnated vinyl
monomers such as styrene, methylmethacrylate, and their combination (50:50, v/v) under hot-compression of the sample to 50-70
% of that original radial dimension. Boric acid (BA) was impregnated into the wood at 1 % aqueous solution concentration prior to
monomer treatment. CWPC with and without BA-pretreatment was tested in terms of biological resistance and mechanical and
thermal properties. Surface hardness of CWPC was superior to wood polymer composite (WPC) obtained at the same polymerization
temperature and time by a conventional heat process in an oven without compression. Modulus of elasticity and rupture were also
considerably improved with this newly introduced in situ polymerization process, suggesting the great potential of CWPC for exterior
use. Table 1.8 shows surface hardness (Brinell hardness) values for the CWPC, WPC, and untreated compressed wood. It is seen that
CWPC is superior to WPC even at almost half the level of monomer loading, concerning hardness.
Ammonia treatment
Anhydrous ammonia is a strong, hydrogen bonding, low molecular weight solvent, which penetrates not only into amorphous areas of
the fiber cell wall but also into the lignin binding material of wood. Some of the hydrogen bonds responsible for the rigidity of wood

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

are broken by the ammonia. This results in a softening or plasticizing of the fiber structure so that the wood exhibits sponge-like
characteristics and can be compressed at moderate pressure. As the ammonia is removed from the wood, hydrogen bonds are again
formed between the polymer chains, although not necessarily at the same locations on the polymer chains or between the same
microfibrils, resulting in the wood once again becoming rigid [Favot 1986].

Table 1.8 Brinell hardness of CWPC at 50% compression set and WPC obtained by vinyl polymerization with and without
boric acid (BA) pretreatment. [Yalinkilic et al 1999]

SD: Standard deviation; ST: styrene; SG: Specific gravity


Wood plasticized with anhydrous ammonia can be compressed by cold compression in a press to convert the wood to a much harder,
mar-resistant, glossy-surfaced product. Although such processes have been known for many years, practical and inexpensive
commercial systems have not been developed. Favot [1986] proposed a method in which a low quality, low density solid wood of high
moisture content was firstly impregnated with anhydrous ammonia. The wood was plasticized to a sponge-like form saturated with
water and ammonia. The plasticized wood is then placed between the perforated press plates of a cyclic press, while maintaining the
temperature of the plasticized wood member below 100C. The wood is subjected to a plurality of compression cycles with the entire
wood being substantially compressed to a predetermined thickness, held at that thickness for a short period of time and released
during each cycle. In this manner, water and ammonia are squeezed out of the wood to a moisture content of less than 30 % to
obtain a damp wood with predetermined reduced thickness. This damp compressed wood is then dried to obtain a dry, permanently
densified solid wood product having the characteristics of high quality natural hardwood.

1.2.7 Effect of Wood Permeability on Properties of WPC


Wood impregnation depends greatly on its permeability to the impregnating medium. Experimental techniques have been developed
and data on wood permeability coefficients for gases and liquids (phenolic alcohol, urea-formaldehyde, polyester resins, and styrene)
are accumulated [Galperin et al 1995]. The transport of gas and liquid through the wood generally follows Darcys law. The pores
connecting cells in the wood create much of the resistance to flow of gases and liquids. The coefficient of wood permeability along the
grain is generally 100-1000 times larger than that across the grain. The main flow paths in wood are the large conducting anatomical
elements such as vessels and tracheids. The pathway of the penetration of liquids into wood has been examined using aqueous
inorganic solutions [Rudman 1966a,b, Bailey and Preston 1969]. The enhancement of wood properties is dependent primarily on the
chemical used and their interaction with wood cell walls. In addition, the extent of impregnation of these substances into wood and
the pattern of their distribution within wood may also be important factors in achieving desired wood properties. Pits are considered
to be important pathways for liquid flow in wood [Kininmonth 1972, Murmanis and Chudnoff 1979].
Siau et al[1978] investigated the fractional volumetric retentions of monomer and polymer that were related to the longitudinal air
permeabilities and the anatomical structures of 22 species of wood. It was found that all five diffuse-porous species including red
maple, yellow-poplar and the gums are easily treatable, while on the other hand, the white oaks, hackberry, black oak, and blackjack
oak were difficult to treat and therefore are considered unsuitable for wood-polymer composites. The remaining species were
moderately treatable and their suitability for composites would depend on specimen size and treatment conditions. Singh et al[1999]
investigated the relationship between pit membrane ultrastructure and chemical impregnability of wood. It was found that the pattern
of wood cell impregnation for alder differed significantly from that of Eucalypt. In alder wood all cell types e.g. vessels, fibers and
rays, were impregnated in similar proportions. However, in eucalypt wood the impregnation material was largely confined to ray cells
and the lumina of vessels; other cell types were either not impregnated or impregnated in very small numbers.
For WPC preparation, low or medium viscosity of the treating fluid, such as monomers or prepolymers, is needed, which can be forced
into wood using vacuum or pressure. It has been observed that high molecular weight polymer, e.g. polyester resin, fills only
microcapillary cavities and cannot penetrate into the cell walls, while phenol resins penetrate well into vessels and tracheid cell
cavities as well as into cell walls of birch and pine. The polymer distribution in the wood can be determined by scanning electronic
microscopy and IR-spectroscopy studies.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

1.2.8 Summary of Wood Quality Improvement Methods and Technologies


In summary, it is observed that wood has a variety of properties as an industrial raw material, and it has also some unfavourable
drawbacks, such as high moisture uptake, biodegradation, and dimensional variations. These defects could be ascribed to the
presence of numerous hydroxyl groups in wood components and various cavities in the wood. It is also observed that wood could be
modified through different methods such as by impregnation with reactive chemicals, by compression, or by a combination method of
chemical impregnation and compression. Various modifications of wood, either by physical or chemical treatment, have been
attempted to improve its properties. A number of studies have been directed to chemical modification of wood to improve its
hardness, mechanical properties and dimension stability. Efforts to reinforce wood with polymers have been extensively made in the
past. Both thermoplastic and thermosetting systems have been used. The most common and inexpensive technique developed is in
situ polymerization of monomers such as styrene and methyl methacrylate, or prepolymers such as phenol-formaldehyde. This type
of treatment has been performed commercially since the early 1960s. Most of the commercial vinyl monomers do not bond strongly
to the components of wood. The polymers formed thus only fill the empty lumens of the wood. This leads to a mixture of two
materials rather than a composite. It has been observed that the polymer existing in the cell wall of a WPC has a greater influence on
the dimensional stability of the WPC, and is more important than the polymer existing in the voids like the cell lumen. Polar
monomers can swell cell wall to give better chemical resistance and dimensional stability of the resulting WPC products. Thus
modification of vinyl monomers by introduction of polar groups has been attempted such as ethyl alpha-hydroxymethylacrylate
(EHMA), which was found to be a good monomer for cell wall swelling, but the monomer modification is a time-consuming process.
Wood impregnated with thermosetting resins such as water-soluble phenolic resins, urea-formaldehyde and melamine-formaldehyde
prepolymers, improves its compressive strength properties and moisture-related shrinking and swelling behaviors. The thermosetting
resins can penetrate and bulk the cell wall, and may react with the hydroxyl groups of the wood components, preventing shrinkage of
the wood upon drying. But in most cases, the resulting wood-polymer composites are quite brittle, and the good toughness properties
of the original wood are deteriorated.
Compression of wood has been investigated. The mechanical properties of wood such as stiffness, strength, surface hardness, and
abrasion resistance can be improved by compression perpendicular to the grain. However, most compressed wood appears to expand
almost to its initial uncompressed state under the influence of moisture and heat. It is observed that dimensional stability of wood can
be improved by steaming or heating the wood when it is in a compressed state.

1.3 Literature Review on Polymer/Layered Silicate Nanocomposites


Polymer-clay nanocomposites were first reported in the literature as early as 1961, when Blumstein demonstrated polymerization of
vinyl monomers intercalated into montmorillonite clay [Blumstein 1961].Toyota researchers work in nanocomposites highlighted to
academic and industrial researchers the potential of nanocomposites [Kojimaet al 1993a, b; Usuki et al 1993a, b].In comparison with
the conventional polymer composites such as glass fiber and carbon fiber reinforced polymer composites, polymer-clay
nanocomposites are a new class of composites that have a dispersed phase with at least one ultrafine dimension in a few nanometers.
Clay layers dispersed at the nanoscale in a polymer matrix act as a reinforcing phase to form an important class of organic-inorganic
nanocomposites. These nanocomposites are also called polymer-clay nanocomposites. With the addition of a very small amount of
nanofiller into the polymer matrix, the resulting nanocomposites exhibit substantial improvement in many physical properties,
including mechanical properties (tensile modulus and strength, and flexural modulus and strength), decreased thermal expansion
coefficient, decreased gas permeability, increased swelling resistance, enhanced ionic conductivity, and flame resistance. [Gilman
1999, Kornmann et al 2001, Lan et al 1995, Byun et al 2001, Su and Wilkie 2003, Dietsche et al 2000]. The main reason that
nanocomposites possess special properties not shared by conventional composites is the large interfacial area per unit of weight of
the dispersed phase (e.g., 750m2/g).The clay particles are essentially bundles of nanoclay sheets that are roughly 1 nm in thickness
and 100-1000 nm in breadth. For instance, an 8-m clay particle has more than 3000 nanoclay sheets of 1 nm thickness and 20-200
nm in diameter. Since the hydrophilic clay sheets are not compatible with hydrophobic polymer matrices, and the spacing between the
clay sheets is extremely narrow, direct diffusion of polymer chains in the clay galleries is not likely. This often leads to aggregation of
clay particles, and the aggregated clay sheets act as stress-concentrated sites in the polymer matrix. Thus chemical treatment or
functionalization of the clay by adsorption of organic molecules is generally needed so as to weaken the interlayer cohesive energy,
leading to a better interaction between the nanoclay and the polymer matrix. A brief review regarding recent developments in the
synthesis and properties of modified clay and polymer-clay nanocomposites is given in this part.

1.3.1 Clay Structure and the Challenge of Processing


Clay consists of small crystalline particles made up of aluminosilicates of various compositions, with possible iron and magnesium
substitutions by alkalis and alkaline earth elements. The basic silicon-oxygen unit is a tetrahedron, with four oxygen atoms
surrounding the central silicon. The tetrahedral are linked to form hexagonal rings. This pattern repeats in two dimensions to form a
sheet. Aluminum, in combination with oxygen, forms an octahedron, with the aluminum at the center, and the octahedral link to form
a more closely packed two-dimensional sheet. There are two basic types of clay structures (1:1 and 2:1). Kaolinite is 1:1 type of
nonswelling dioctahedral clay. The kaolinite crystal is a sheet of alumina octahedral sitting on top of a sheet of silica tetrahedral. The
apical oxygen atoms from the silica are shared with the aluminum atoms of the upper layer. The other basic type of clay is of the 2:1
type (i.e., two sheets of silica to one of alumina or two sheets of silica at one of magnesium oxide). The two parent materials are
pyrophyllite and talc, with alumina and magnesia, respectively, in the central layer.
Clays used in preparing polymer-clay nanocomposites belong to the 2:1 layered structure type. Montmorillonite, hectorite, and
saponite are the most commonly used layered silicates clay. Table 1.9 gives the chemical formula and characteristic parameters of
commonly used 2:1 phyllosilicates clay. Montmorillonite is one of the most interesting and widely investigated clays for polymer
nanocomposites. The structure of montmorillonite consists of layers made up of one octahedral alumina sheet sandwiched between
two tetrahedral silica sheets, as shown in Figure 1.1. Stacking of the silicate layers leads to a regular van der Waals gap between the

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

layers. Approximately one in six of the aluminum ions in the octahedral layers of montmorillonite is isomorphously substituted by
magnesium or other divalent ions. The isomorphic substitution renders negative charges that are counterbalanced by cations residing
in the interlayer.

Table 1.9 Chemical formula and characteristic parameters of commonly used 2:1 phyllosilicates [Ray and Okamoto 2003].

M monovalent cation ; x degree of isomorphous substitution (between 0.5 and 1.3)


CEC cation exchange capacity.

Figure 1.1 Idealized structure of a montmorillonite layer showing two tetrahedral-site sheets fused to an octahedral-site
sheet (2:1 type) [Outubuddin and Fu 2002, Ray and Okamoto 2003].

Silicated nanolayers due to their high aspect ratio and surface area are particularly desirable for polymer reinforcement. The clay
particles are essentially bundles of nanoclay platelets. An 8-m clay particle has more than 3000 nanoclay platelets of 1 nm thickness
and 20-200 nm in diameter (Figure 1.2). In the synthesis of polymer-clay nanocomposites, colloid and surface chemistry play
important roles. Due to high interlayer cohesive energy, dispersion of clay layers in polymers is hindered by the inherent tendency to
form facetoface stacks in agglomerated tactoids of the clay. Exfoliation (e.g. sheets separation) is further prevented by the
incompatibility between hydrophilic clay and hydrophobic polymers. This often leads to aggregation of clay particles, and the
aggregated clay sheets act as stress-concentrated sites in the polymer matrix. Thus chemical treatment or functionalization of the
clay by adsorption of organic molecules is generally needed so as to weaken the interlayer cohesive energy and enhance the
compatibility between the nanoclay and the polymer matrix. The processing challenge is to disperse not only the clay particles but
also the platelets, so that the high aspect ratio (>50) and high surface area (>750 m2/g) of the clay nanolayers could ultimately be
utilized.

Figure 1.2 Processing challenge for exfoliated nanoclay [Dennis et al., Southern Clay Products].

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

1.3.2 Structure and Properties of Organically Modified Layered Silicate


The physical mixture of a polymer and layered silicate clay may not form a nanocomposite. This situation is analogous to polymer
blends, and in most cases separation into discrete phases takes place. In immiscible systems, which typically correspond to the more
conventionally filled polymers, the poor physical interaction between the organic and the inorganic components leads to poor
mechanical and thermal properties. In contrast, strong interactions between the polymer and the layered silicate in polymer
nanocomposites lead to the organic and inorganic phases being dispersed at the nanometer level. As a result, nanocomposites exhibit
unique properties not shared by their micro counterparts or conventionally filled polymers. Pristine clay usually contains hydrated
inorganic cations such as Na+, K+, and Ca2+. In this pristine state, layered silicate clays are not miscible with hydrophobic polymers.
Generally, to change the clay surface from hydrophilic to hydrophobic, one can use ion-exchange reactions with cationic surfactants
including primary, secondary, tertiary, and quaternary alkylammonium or alkylphosphonium cations. When the inorganic cations are
exchanged by organic cations, such as surfactants and polyelectrolytes, the clay surfaces change from hydrophilic to hydrophobic or
organophilic. The organic cations lower the surface energy and decrease the cohesive energy by expanding the interlayer distance,
thus facilitating the wetting and intercalation of monomer or polymer. In addition, the organic cations may contain various functional
groups that react with monomer or polymer resin to improve interfacial adhesion between clay nanolayers and polymer matrix.
Fourier transform spectroscopy (FTIR), x-ray diffraction (XRD), and differential scanning calorimetry (DSC) can be used to probe the
interlayer structure and pack of intercalated cationic surfactants. Figure 1.3 shows the possible configurations such as flat monolayer,
bilayer, pseudo-trilayer, and inclined paraffin structure.

Figure 1.3 Orientations of alkylammonium ions in the galleries of clay layers with different layer charge densities
[Qutubuddin and Fu 2002].

Complete dispersion or exfoliation of clay tactoids in a monomer or polymer matrix may involve three steps similar to the dispersion
of powders in liquids, as identified by Parfitt [1981]. The first step is wetting the surface of clay tactoids by monomer or polymer
molecules. The second step is intercalation or infiltration of the monomer or polymer into the clay galleries, and the third step is
exfoliation of clay layers. The first and second steps are determined by thermodynamics, while the third step is controlled by
mechanical and reaction driving forces. The dispersion of clay tactoids in a polymer matrix can result in the formation of three types
of composites, as shown in Figure 1.4. The first type is a conventional composite that contains clay tactoids dispersed simply as a
segregated phase, resulting in poor mechanical properties of the composite. The second type is intercalated polymer-clay
nanocomposite, which is formed by the infiltration of one or more molecular layers of polymer into the clay host galleries. The last
type is exfoliated polymer-clay nanocomposites, characterized by low clay content, a monolithic structure, and a separation between
clay layers that depends on the polymer content of the composite. Exfoliation is particularly desirable for improving specific properties
that are affected by the degree of dispersion and resulting in interfacial area between polymer and clay nanolayers. Exfoliated
nanocomposites usually provide the best properties enhancement due to the large aspect ratio and surface area of the clay.
Homogeneous dispersion of clay nanolayers in a polymer matrix provides maximum reinforcement via distribution of stress and
deflection of cracks resulting from an applied load. Interactions between exfoliated nanolayers with large interfacial area and
surrounding polymer matrix lead to higher tensile strength, modulus, and thermal stability. Conventional polymer-filler composites
containing micro-size aggregated tactoids also improve stiffness, but with less improvement in strength, elongation, and toughness of
the resulting composites. However, exfoliated clay nanocomposites of Nylon-6 and epoxy have shown improvements in all aspects of
thermomechanical behaviour. Exfoliation of silicate nanolayers with high aspect ratio also provides other performance enhancements
that are not achievable with conventional particulate composites such as barrier and reduced flammability properties. The
impermeable nanoclay layers could provide a tortuous pathway for a permeant to diffuse through the nanocomposite. The hindered
diffusion in nanocomposites leads to enhanced barrier property, reduced swelling by solvent, and improvements in chemical stability
and flame retardance.

Figure 1.4 Schematic illustrations of three types of polymer-clay composites.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

1.3.3 Techniques Used for the Characterization of Nanocomposites


Generally the structure of nanocomposites has typically been estabilished using X-ray diffraction (XRD) analysis and transmission
electron micrography (TEM) observation. XRD is most commonly used to probe the nanocomposite structure. By monitoring the
position, shape, and intensity of the basal reflections from the distributed silicate layers, the nanocomposite structure (intercalated or
exfoliated) may be identified. In an exfoliated nanocomposite, the extensive layer separation associated with the delamination of the
original silicate layers in the polymer matrix results in the eventual disappearance of any coherent X-ray diffraction from the
distributed silicate layers. On the other hand, for intercalated nanocomposites, the finite layer expansion associated with the polymer
intercalation results in the appearance of a new basal reflection corresponding to the larger gallery height. TEM allows a qualitative
understanding of the internal structure, spatial distribution of the various phases, and views of the defect structure through direct
visualization. Figure 1.5 shows four samples of X-ray diffraction, montmorillonite (MMT) with its ordered platy morphology causes
x-rays to diffract indicating the distance from the top of one platelet to the top of the next platelet (Basal spacing or D001 spacing).
The organoclay Cloisite 15A has a d-spacing of about 32.When Cloisite15A is compounded with polyamide 6 (PA6) in a single screw
extruder, the peak intensity decreases, the peak shape broadens but the peak position remains at about 32.Co Rotating Low Shear
and Medium Shear samples further decrease in size and show the platelets move further apart, 34 and 38, respectively. The
sample made in the Tangential Medium Shear extruder showed no XRD peak. The lack of XRD peak suggests either an exfoliated or
an intercalated disordered nanocomposite. TEM is required to see what the platelet distribution looks like. Figure 1.6 shows the TEM
image of the sample from the Single Screw extruder shows big ribbons that are of stacks of intercalated platelets. The samples from
the Co Rotating extruder show increasing dispersion as the ribbons get smaller and some single platelets can be seen, particularly in
the sample from the Medium Shear extruder.

Figure 1.5 Four samples of X-ray diffraction [Dennis et al, Southern Clay Products].

VanderHart et al [2001] first used solid-state nuclear magnetic resonance (NMR) (1H and 13C) as a tool for gaining greater insight
about the morphology, surface chemistry, and the dynamics of exfoliated polymer clay nanocomposites. They were especially
interested in developing NMR methods to quantify the level of clay exfoliation, a very important facet of nanocomposite
characterization. Some researchers [Wu 2001, Loo and Gleason 2003] also used FTIR spectroscopy to examine the structure of the
nanocomposites. Very recently, Nascimento et al [2002] presented the resonance Raman characterization of a polymer/clay
nanocomposite formed by aniline polymerization in the presence of montmorillonite.

1.3.4 Synthesis and Properties of Polymer-clay Nanocomposites


There are three general approaches to the synthesis of polymer-clay nanocomposites. Figure 1.7 shows the dispersion mechanism of
these different synthesis polymer-clay nanocomposites. In the first approach, a monomer or prepolymer is mixed with organophilic
clay and followed by polymerization. This in situ polymerization technique was first developed by Toyota group to make Nylon-6

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

nanocomposites from caprolactam monomer. For clays modified with a long-chain surfactant, the galleries swollen by the monomer or
prepolymer show a d-spacing indicative of a paraffin monolayer arrangement, as illustrated in Figure 1.8. Upon polymerization, the
clay nanolayers are forced apart and no longer interact through the surfactant chains. Thus, highly exfoliated nanocomposites are
formed. In terms of both experimental results and thermodynamic considerations, this method is most promising for the synthesis of
highly exfoliated nanocomposites.

Figure 1.6 TEM imagine of PA6/nanoclay by extruder process [Dennis et al, Southern Clay Products].

Figure 1.7 Dispersion mechanism of nanoclay to a polymer.

Figure 1.8 Model of alkylammonium exchanged clay swollen by monomer or polymer precursors such as styrene,
-caprolactam, and epoxide [Qutubuddin and Fu 2003].

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

The second method of nanocomposite synthesis involves dissolving a polymer in a solvent, mixing with organophilic clay, and then
removing the solvent. This is based on a solvent system in which the polymer or prepolymer is soluble and the silicate layers are
swellable. The layered silicate is first swollen in a solvent, such as water, chloroform, or toluene. When the polymer and layered
silicate solutions are mixed, the polymer chains intercalate and displace the solvent within the interlayer of the silicate. Upon solvent
removal, the intercalated structure remains, resulting in polymer layered silicate nanocomposites. The aggregate formation of
nanoclay sheets can be somewhat alleviated if nanoparticles are mixed in polymer solutions, but the solvent recovery will add extra
costs and pose environmental concerns.
The third approach is melt intercalation and involves heating a mixture of polymer and organophilic clay above the glass transition or
melting temperature. The properties of nanocomposites depend on the compatibility or interaction between the polymer and
organophilic clay. A major difficulty in nanocomposite synthesis is that nonpolar polymers such as polypropylene do not easily
intercalate into clay galleries. Melt intercalation is very practical from the point of view of direct manufacturing of composite articles.
This method, however, suffers from such problems as degradation and aggregate formation. The matrix polymers may potentially
degrade due to shear heating, especially if the mixing time is long and the mixing temperature is high, and obviously, one cannot
heat thermosetting polymers. The nanofiller particles, on the other hand, may form aggregates if not chemically treated before use.
1.3.4.1 Polymer clay Nanocomposites Synthesized from Monomers or Prepolymers in situ Polymerization
Although inter-lamellar polymerization techniques using appropriately modified layered silicate or synthetic layered silicates have
long been known, the field of polymer layered silicate nanocomposites gained momentum recently due to the report of a PA6/MMT
nanocomposite from the Toyota research group [Kurauchi et al 1991], where very small amounts (4.2%) of layered silicate loadings
resulted in significant improvements in thermal and mechanical properties (Table 1.10). Exfoliated Nylon-6 nanocomposites also
demonstrated significant improvements in dimensional stability, barrier properties, and flame retardant properties [Kojima et al
1993]. Inorganic cations were replaced by alkylammonium fatty acids to make the clay surface compatible with the monomer,
-caprolactam. The monomer was polymerized in the interlayer galleries of modified montmorillonite (MMT) to form Nylon-6-clay
hybrids or nanocomposites. The protonated alkylammonium acidic cations catalyze the intragallery polymerization of caprolactam,
thereby providing a driving force for nanolayer exfoliation in the resulting composite. Individual silicate layers (less than 1-nm
thickness) of montmorillonite were completely exfoliated and homogeneously dispersed in Nylon-6 matrix, as revealed by X-ray
diffraction and transmission electron microscopy (TEM). PMMA, PS, PE and PP are also reported to prepare nanocomposites through in
situ polymerization [Garces et al 2000, Su and Wilkie 2003, Zerda et al 2003, Lee and Jang 1996].

Table1.10 Mechanical and thermal properties of Nylon-6 and Nylon-6-Clay nanocomposites [Kurauchi 1991].

Chen et al [1999] used polymer-clay based nanocomposites synthesis technique for the preparation of novel segmented
polyurethane/clay nanocomposites on diphenylmethane diisocyanate, butanediol and preformed polycaprolactone diol. Wang and
Pinnavaia [1994, 1998] reported the preparation of polyurethane-MMT nanocomposites using a direct in situ

intercalative

polymerization technique. WAXD analyses of these nanocomposites established the formation of intercalated structure.
Epoxy-clay nanocomposites from epoxide prepolymer have been investigated by some researchers [Wang and Pinnavaia 1994, 1998,
Shiet al 1996, Feng et al 2002]. In general, the synthesis is similar to that of Nylon-6 and PS described earlier, but with the
additional need for a curing agent. First, the clay is rendered hydrophobic by cationic exchange with appropriate surfactant molecules.
Next, the organoclay is dispersed in a mixture of epoxy resin and curing agent. Finally, the temperature is increased to cure the
resin. Acidic onium ions catalyze intragallery polymerization at a rate that is comparable to extragallery polymerization. The relative
rates of reagent intercalation, chain formation, and network cross-linking have an important effect on the initial gelation and final
curing of epoxy-clay exfoliated nanocomposite. Aliphatic amine, aromatic amine, anhydride, and catalytic curing agents have been
investigated to form an epoxy composite with broad glass transition temperature, Tg. The thermomechanical properties of epoxy
nanocomposites show dramatic improvements, particularly in the rubbery state.
1.3.4.2 Polymer-clay Nanocomposites Synthesized from Polymer Solution
It is well known that water soluble polymers easily intercalated into clay galleries in aqueous suspension. Recent interest has focused
on the blending of organoclay with polymers dissolved in organic solvents, Ogata et al [1997] investigated polyethylene oxide (PEO)
mixed with disearyldimethyl ammonium exchanged organoclay in chloroform. After the chloroform was evaporated, micrometer thick
films were obtained. The organoclay accelerated the crystallization of PEO and induced a preferred orientation of the polymer.
Dramatic improvements in barrier properties, thermal stability, and modulus were observed for these nanocomposites. The polymer
solution approach was also used to prepare polyurethane-clay nanocomposites. 12-Aminolauric acid and benxidine ion-exchanged
organoclays and polyurethane were mixed in dimethylformamide solvent. After degassing of the mixture and removal of solvent at
80C, an elastic nanocomposite film was obtained. Exfoliation of polyurethane-clay nanocomposites was indicated by x-ray diffraction

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

and TEM. The tensile strength and elongation were significantly enhanced, while water absorption decreased.
1.3.4.3 Polymer-Clay Nanocomposites Synthesized via Melt Intercalation
Vaia and coworkers [1995, 1996] prepared intercalated PS-clay nanocomposites via polymer melt intercalation using long-chain
primary and quaternary alkylammonium-exchanged clays. The organoclay was mixed with commercial available PS at a temperature
above the Tg via melt processing. The diffusion of PS into the clay galleries is slow and depends on many factors, including polymer
molecular weight (MW), processing temperature, surfactant properties, and interactions between the polymer and the organoclay.
Dennis et al [Southern Nanoclay Products]investigated the synthesis PA6/clay nanocomposites by extruder processing. The
delamination dispersion is monitored by X-ray diffraction and by transmission electron microscopy (TEM). The sample from the
Tangential Medium Shear extruder shows excellent delamination and dispersion. Figure 1.9 gives the relationship between the
dispersion and the tensile modulus. A few mechanical properties were also measured. This figure shows the tensile modulus increases
as dispersion increases with the TEM micrograph indication. A big increase in tensile modulus occurs by reinforcing the PA6 with
ribbons of clay tactoids or intercalants. Byun et al [2001] prepared the resol type phenolic resin/layered silicate nanocompoistes by
melt intercalation using various layered silicates such as sodium MMT and -amino acid modified MMT. The mechanical properties of
the MMT reinforced phenolic resin are improved more significantly than pristine phenolic resin.
Silicated clay-polymer nanocomposites generally showed excellent barrier properties. Several-fold reduction in the permeability of
small gases, e.g. O2, H2O and CO2 has been observed for some of the polymer/layered silicate nanocomposites, because of the
formation of a tortuous path in the presence of clay in the nanocomposites. For example, it is observed that at 2% clay loading, the
permeability of water vapour in the polyimide/layered silicate nanocomposites was reduced ten-fold compared to that of pristine
polyimide [Ray and Okamoto 2003]. Figures 1.10 and 1.11 illustrate the gas permeation behaviour of O2 (Fig. 1.10) and water
vapour (Fig.1.11) in the nanoclay-nylon6 packaging films with different clay loadings. As one can see that both O2 gas and water
vapour permeabilities are significantly reduced with addition of a small amount of nanoclay into the nanocomposite films.

Figure 1.9 Relationship between the dispersion and the mechanical properties [Dennis et al Southern Clay Products, Inc.].

Figure 1.10 Effect of relative humidity on O2 transmission at various clay loadings [Dennis et al, Southern Clay Products,
Inc.].

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch01.html

Figure 1.11 Water vapour transmission vs. clay content [Dennis et al, Southern Clay Products, Inc.].

1.3.5 Summary of Polymer/Layered Silicate Nanocomposites


Clays functionalized via ion exchange with cationic surfactants have been successfully used to prepare polymer-clay nanocomposites.
Organoclay nanolayers exfoliated in polymer matrix can dramatically improve mechanical reinforcement, thermal and chemical
stability, barrier properties, etc. These improved properties are generally achieved at lower silicate content ( 5%) compared with
that of conventionally filled systems. For these reasons, the silicated clay nanocomposites are far lighter in weight than conventional
composites, and make them competitive with other materials for specific applications. A major difficulty in the synthesis of new
polymer nanocomposites is the compatibility of the organoclay surface with the polymer, particularly if it is hydrophobic. The
interfacial adhesion between polymer and clay is very important for the reinforcement, due to the tremendous surface area of clay
nanolayers. Of the three main approaches for the synthesis of polymer-clay nanocomposites, in situ polymerization of the monomer
or prepolymer seems to be most effective in achieving exfoliation of the clay nanolayers. The polymer-clay nanocomposites showed
excellent barrier properties. Several-fold reduction in the permeability of small gases, e.g. O2, H2O and CO2 has been observed for
some of the polymer/layered silicate nanocomposites, because of the formation of a tortuous path in the presence of clay in the
nanocomposites. Barrier properties, fire resistance and good mechanical properties are of great importance for the application of a
wood product. Most of the published literature on polymer-clay nanocomposites focuses on synthesis and characterization. It is
necessary to conduct research issues on novel applications, scale-up, and industrial processing of nanocompoistes, including control of
rheology, morphology, and specific performance. Fundamental research is needed to improve the understanding of the mechanism of
exfoliation, interfacial adhesion, and overall reinforcement. New synthesis methods and models for the prediction of nanocomposite
formation and properties are obviously desirable.
It would be attractive to prepare a WPC product via nano technique using the layered clay nanofillers and above properties may be
improved simultaneously. This led us in the present work to investigate the feasibility to prepare a solid wood polymer nanocomposite
from solid wood using nano technique approach.

1.4 Motivation for the Study


Following this literature survey, it was decided upon making this study of wood/nanoclay/MUF composites. Reasons for this were:
- There are very few research projects on thermoset/nanoclay systems, most of what is found in the literature in on nanoclay with
thermoplastics.
- Many studies on thermoplastic nanocomposites cite nanoclays as reinforcing agents.
- Nanoclays of several varieties can be obtained on the market. That is not so for other nanoparticles such as aluminum oxide or
titanium oxides, which are, in addition, very difficult to modify chemically.
- Modification with MUF will increase several properties of wood, but without nanoclays, or nanoparticles we do not expect resistance
to wear or scratch to be optimized.
-We hope nanoclays, due to their small size, may modify the cell wall and improve dimensional stability.
- Nanoparticles, due to their small size, are hard to see with naked eyes and do not modify appearance of composites which in field of
parquets/wood floors is an obvious prequisite.
For all these reasons this particular kind of composite was selected almost promising.
Xiaolin Cai, 2007

Chapter 2 Experimental and Methodology

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Chapter 2 Experimental and Methodology

Chapter 2 Experimental and Methodology


Table des matires
2.1 Survey of Experimental Design
2.2 Wood Sample Preparation
2.2.1 Wood Cutting Design
2.2.2 Lumber Drying and Sample Preparation
2.3 Synthesis of Low Viscosity Melamine-Urea-Formaldehyde (MUF) Resin and Mix of Nanoclay and MUF Resin
2.3.1 Synthesis of Low Viscosity Melamine-Urea-Formaldehyde (MUF) Resin
2.3.2 Mixture of MUF Resin and Nanoclay
2.3.3 Characterization Ball Mill Ground Layered Aluminium Silicate Nanoclays
2.3.4 Characterizing Nanoparticles of MUF/Nanoclay System
2.4 Color Measurement
2.5 Characterizing the Interphase and Morphology of Wood MUF Nanoclay with Electron Probe Micro-Analysis (EPMA)
Element Distribution

2.1 Survey of Experimental Design


A general description of experiment and methodology survey, experimental techniques and methods involved in this work is provided
in this section. Figure 2.1 indicates the survey of the experimental design, which contains the preparation of wood samples, the
synthesis of melamine-urea-formaldehyde resin and the blending of MUF/nanoclay. The effect of layered aluminium silicate nanoclays
on the curing behaviour and the viscoelasticity of melamineureaformaldehyde resin were investigated through DSC/TGA analysis,
DMTA and SEM observation. The dispersion of nanoclay in the MUF was investigated through X-ray diffraction analysis, and Zetasizer
observation. First, defect free wood samples used in this study were prepared from Trembling aspen green logs, and low viscosity MUF
resin designed for impregnation purposes was synthesized in the laboratory of PFInovation Forintek (Eastern Region). Second, the
impregnation solutions of nanoclay/MUF mixtures were prepared by mixing the nanoclay fillers with the MUF resin at high mixing
speed with a high-speed mixer. After that, the wood samples were impregnated at selected vacuum then followed a pressure
procedure, and wood polymer composites were prepared through in situ polymerization of the impregnated wood at selected
temperature, time and pressure. The obtained products were stored in a 20 C / 65% RH conditioning chamber for at least 3 weeks
before being measured the physical/mechanical properties such as static bending, surface hardness, dimensional stability and water
absorption. In order to investigate the interphase, morphology and bonding between the woodMUFnanoclay and the effect of these
characters on the mechanical/physical properties on the wood polymer nanocomposites, X-ray microprobe, scanning electron
microscope (SEM), transmission electron microscope (TEM) and atomic force microscope (AFM) analysis were used.
The details of experiment and methodology description are divided in six parts which include wood samples preparation, low viscosity
MUF resin synthesis, the mixing of layered aluminium silicate nanoclays and MUF resin, impregnation and in situ polymerization, the
measurement of mechanical/physical properties and the investigation of interphase and morphology of wood MUF nanoclay. The
following paragraphs describe the experimental procedure we used and which is not included in the publications.

Figure 2.1 Survey of the experimental procedure.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

2.2 Wood Sample Preparation


2.2.1 Wood Cutting Design
Trembling aspen (Populus tremuloides) used in this work was provided by a local tree farmer in Ste-Foy (Qubec, Canada). To
compare the effect of different chemical formulations of impregnation on the mechanical/physical properties of wood polymer
nanocomposites, defect-free samples were prepared carefully through strict end-match selection to diminish the effect of the variation
of wood properties. Figure 2.2 shows the wood samples cutting design. Figure 2.2(a) is the trembling aspen green log received from
the supplier. Figure 2.2 (b) shows as sketch that the reveals method of cutting up the bolt and marking the sticks, AS01E1 (where
A-aspen, S-sapwood, 01-log lumber, E-eastern direction, 1-lumber number) and AH01-1 (where A-aspen, H-heartwood, 01-log
number, 1-heartwood lumber number), respectively. Figure 2.2 (c) shows the arrangement of samples end match, 1, 2,... with
different treatments, where Treat1, Treat2, etc. are different treatments, respectively. The green logs (Figure 2.3 (a)) were cut to
lumber by Wood-Mizer (Canada, Figure 2.3).

Figure 2.2 (a)Aspen green log before cutting.

Figure 2.2 (b) Sketch showing method of cutting the bolt and marking the sticks. For example in lumber AS01E1 (stands for:
A aspen, S sapwood, 01 log lumber, E eastern direction, 1 - lumber number) and AH01-1 (Stands for: A aspen, H
heartwood, 01 log number, 1 heartwood lumber number), respectively.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

Figure 2.2 Wood sample cutting design (a) Aspen green log before cutting, (b) Sketch showing method of cutting up the bolt
and marking the sticks, (c) Samples end match.

Figure 2.3 Forintek staff is preparing specimens with Wood-Mizer.

2.2.2 Lumber Drying and Sample Preparation


The trembling aspen green logs were cut into lumber and then dried with (Figure 2.4) the procedure description in Drying aspen and
birch Northern hardwoods need a special approach at the FPInnovations Forintek (Eastern Region), Qubec. After 6-8 days kiln
drying, the lumber with 8-10% moisture content was removed from the kiln. Wood samples were chosen from defect-free boards.
End-matched samples for both the modified and the control samples were prepared with dimensions of 3 2 0.5 (L T R) for
Brinell hardness according to Europee standard for measuring Wood and parquet flooring Determination of resistance to indentation

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

(EN1534)and water absorption/dimensional stability measurements according to ASTM D1037, 4 4 0.5 (L T R) for
abrasion resistance evaluations according to ASTM D4060-95, and 8 0.5 0.5 (L T R) for static bending tests according to
ASTM D-143.

Figure 2.4 Preparation of kiln drying green aspen lumbers.

2.3 Synthesis of Low Viscosity Melamine-Urea-Formaldehyde (MUF) Resin and Mix of Nanoclay
and MUF Resin
2.3.1 Synthesis of Low Viscosity Melamine-Urea-Formaldehyde (MUF) Resin
Melamine-urea-formaldehyde (MUF) is a kind of resin widely used in wood industry. It was synthesized based on urea-formaldehyde
resin with a small amount of melamine added to improve the water repellency and dimensional stability. Compared to the phenolic
resin, MUF resion is inexpensive in price and low viscosity melamine-urea-formaldehyde resin is almost transparent that is a very
good characteristic for impregnation. The color of the wood products will not be influenced by the MUF impregnation process. The low
viscosity MUF resin was synthesized specially for impregnation purpose at FPInnovations Forintek (Eastern Region). All materials
including melamine, urea and para-formaldehyde are industrial grade, which were provided kindly by TEMBEC (Canada). To select the
recipe of melamineureaformaldehyde (MUF) resin for impregnation purpose, different receipts with different melamine content
were tested at the beginning of the project.
Recipe 1
Industrial grades of p-formaldehyde (93.5%), urea (98%), melamine (99%) were used as reactants without any purification. The
synthesis process contained two main stages. First, the calculated amounts of reactants and water were charged in the reactor,
equipped with a condenser (Figure 2.5). 848.5 g water (A1), 912.4 g p-formaldehyde (F1) were charged to a reactor, the pH value of
the mixture was adjusted to 7.0 8.0 and then was heated to 75 80C, 807 mg urea (U1) was charged to the system slowly which
was heated to 90C and maintained 30 minutes. The pH was adjusted to 4.6 with HCl and maintained 30 minutes, check the viscosity
until obtained the viscosity C, decreased the temperature to 60C, adjusted the pH value to 7.5-8.0, 136 g water (A2) and 146.2 g
formaldehyde (F2) were charged to the system, followed 227 g melamine were charged to the system, adjust pH to 6.7 heated the
system to 85-90C, maintained the temperature 30 minutes and check the viscosity. Adjust the pH value to 8.0-9.0, decrease
temperature to 60C, 789 g urea (U2) were charged, checked the viscosity until the target viscosity achieved, cooled down the
system and remove the samples to a plastic container.
Recipe 2
The same types of materials provided by TEMBEC were used as reactants without any purification as in recipe 1. The synthetic
process was made up of two main stages. First, the calculated amounts of reactants and water were charged in the reactor, equipped
with a condenser. First, 1590 g of water, 1450 g of p-Formaldehyde and 816 g of urea were charged into a reactor and the pH was
adjusted with NaOH (50%) to 7.0-8.0 with constant stirring at room temperature. The temperature was raised to 75C-80C slowly
within about 30 minutes, maintained 1 hour at this temperature (verify the pH value regularly) then the solution cooled to 60C. The
first batch of melamine (75 g) was then slowly added. After the melamine had completely dissolved, the pH was adjusted with HCl
(10%) to 4.8 and the temperature was raised to 85C in 20 minutes and maintained for 60 minutes. The second batch of melamine
(75 g) was added and maintained until the target A-B viscosity of V/W was attained. The system was cooled to room temperature and
adjusted pH value to neutral (7.0-8.0). The MUF resin synthesis equipment is shown in Figure 2.5.

Figure 2.5 Equipment of melamine urea formaldehyde resin synthesis.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

The different synthesized MUF resins were used to impregnate aspen wood. The resulted impregnated wood samples from recipe 1
had a high ratio of splitting once the samples were in situ polymerized at certain temperature and pressure. The impregnated wood
samples with recipe 1 MUF resin were stored for 3 weeks in a conditioning chamber. An even higher proportion of samples split after
3 weeks condition at 21C and 65 % moisture content. So the MUF resin of recipe 2 was adopted for the impregnation purpose in this
project. All the methods and procedures in the following parts were developed based on the MUF resin of recipe 2. The solid content
of the MUF resin with recipe 2 was measured in an oven at 103C for 12 hours, whose solid content is around 45.9%. The
characterization of melamine urea formaldehyde resin including the gel time test, curing behaviour with differential scanning
calorimeter (DSC) and dynamical mechanical thermal analysis (DMTA) were described later in the section on the mixing of nanoclay
and MUF resin.

2.3.2 Mixture of MUF Resin and Nanoclay


Montmorillonite is classified as Magnesium Aluminium Silicate and it tends to have a sheet-like morphology. Montmorillonite crystals
have a flat, thin sheet, morphology. They have an irregular overall shape and can be up to 1,000 nm in the largest dimension.
However, a side view of the crystal reveals a uniform size of 0.92 nm thickness. The result of this morphology is an extremely large
surface area of about 750 m2/g. Surfaces tend to structure water that is in close proximity to them; therefore, the greater the
surface area, the greater the effect. Being able to generate such large surface area allows nanomontmorillonite to have large impact
on the rheology of the water with a small amount of material.Three layered aluminum silicate nanoclays with different surface
polarity were selected in this project. Nanoclays Cloisite 30B (NF1), Claytone APA (NF2) and Cloisite Na+ (NF3) were provided
by Southern Clay Products, Inc. (USA). The parameters of the physical properties of the nanoclays are given in Chapter 4. The
average particle size of the clay powder received is 8 m, and each particle contains at least 3,000 platelets. Even dispersion of
nanoparticles into a polymer matrix is crucial in preparing a nanocomposite. The process challenge of using the nanoclay is not only
to disperse the nanoparticles into the MUF resin, but also to exfoliate the platelets to achieve the maximum surface area (e.g.,
750m2/g) and high aspect ratio. With this aim, the clay powder was further ground with a ball-mill before being mixed with MUF
resin, which was performed with a ball mill (U.S. Stoneware) equipped with a 2 lceramic jar (Figure 2.6). In the jar 20 porcelain balls
(Fisher Scientific) with a diameter of 2.5 cm were utilized. 100 g of the clays were put into the ceramic jar and milled at a rolling
speed of 162 rpm for the rolls, and 50 rpm for the jar for 48 hours. The impregnation solutions were prepared by adding 1% (wt/wt)
layered aluminosilicate nanofiller for both the ball-mill treated (coded as NFx-m, where x =1,2,3) and unmilled nanoclays (NF1, NF2
and NF3) into the low viscosity MUF resin at a mixing speed of 3050 rpm for 20 min in a Desintegrator from Mitchel (Canada, Figure
2.7).Later, for comparison, both the ball-mill treated and unmilled nanoclays were mixed with the MUF prepolymer to form
impregnation solutions that were subsequently used to impregnate the solid aspen wood samples under specific conditions.

Figure 2.6 Ball mill equipment.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

2.3.3 Characterization Ball Mill Ground Layered Aluminium Silicate Nanoclays


The thermogravimetric analysis (TGA) of the specimens was conducted on a TGA/SDTA851e (TA Instruments) instrument under air
(atmosphere) at a heating rate of 20C min-1 from 30C to 1000C for the received nanoclay. X-ray diffraction (XRD) experiments
were performed directly on the nanoclay received and ground with the ball mill samples using a Siemens D5000 apparatus with CuK
radiation (=0.154 nm) at the Department of Chemistry (CERSIM) of Universit Laval. Both received and ball milled layered silicate
powder were put into a glass capillary with 0.5 mm diameter. The morphologies of the received and ball mill ground nanoclay were
observed and analyzed using a JOEL JSM-840A scanning electron microscope (SEM) at Dpartement degnie des mines, de la
mtallurgie et des matriaux at Universit Laval. Prior to SEM observation, the nanoclay was cast on a support and a gold/palladium
alloy was sputtered on the surfaces. The X-ray composition analysis of montmorillonites received and ball mill ground were performed
with a JOEL JSM-840A to observe the morphology of the clay.

Figure 2.7 Desintegrator for high speed mixing MUF resin with nanoclay.

2.3.4 Characterizing Nanoparticles of MUF/Nanoclay System


The physical and optical properties of nano-particles are related strongly to their size. To find a reliable, accurate and rapid mean of
particle size, measurement and materials characterization in the nanometer size range is crucial. The diversity of nanoparticles and
their applications mean that they are likely to be present in a variety of formats and in concentrations ranging from the very dilute
through to highly concentrated suspensions or emulsions. Being able to measure these particles in their native formats may be
important. The technique of dynamic light scattering (DLS) is ideally suited to the determination of the size of particles in the
nanometer range. The Malvern Zetasizer Nano series uses patented optics that provided exceptional levels of sensitivity and enable
the determination of particle size in samples that contain very small particles and/or particles that are present at very low
concentrations. The particle size of measurement is from 0.6 nm to 6 m. The dispersion and the particle size of nanoclay in the MUF
resin system were measured via Zetasizer Nano ZS (Malvern Instruments, USA). Figure 2.8 shows the image of the equipment.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

Figure 2.8 Zetarsizer Nano ZS.

2.4 Color Measurement


The BYK-Gardner Color-Guide is widely used to measure the color change of wood products due to drying or preservative. The
principle and calculation of color measurement is showed in Figure 2.9 and Figure 2.10, respectively. The colors of the control, MUFand nanoclay/MUF-treated wood samples were measured using a Color-guide BYK-Gardner GmbH equipment (Germany), according to
the procedure of the Color-guide 45o/0o, Cat. No.6805 (Figure 2.11). The color change of each sample was calculated and compared
through the method of Figure 2.9 and 2.10.

Figure 2.9 Principle of color meaurement.

Figure 2.10 Calculation of color differences.

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

Figure 2.11 Color-guide BYK-Gardner GmbH equipment

2.5 Characterizing the Interphase and Morphology of Wood MUF Nanoclay with Electron
Probe Micro-Analysis (EPMA) Element Distribution
Electron microprobe technique is one of the most precise and accurate methods for surface elemental analysis. How EPMA works is
illustrated in the Figure 2.12. The distribution of nanofillers in the wood for the nanofiller/MUF treated wood samples was further
investigated using an Electron Microprobe Analyzer (CAMECA S100, France). The samples used for the EPMA analysis were prepared
using a way similar to the SEM samples before coating. The cross-section of the wood samples was coated with carbon (25 nm) prior
to the measurements in order to make it conductive and to protect the surface from beam damage. Electron microprobe analysis was
performed in a mapping mode with an accelerating voltage of 15 kV and 10 nA. The images of elements Al and Si, which are mainly
from aluminosilcate nanofillers for the nanofiller/MUF impregnated wood, were captured digitally to allow for enhanced analysis of the
samples.
EPMA basically works by bombarding a micro-volume of a sample with a focused electron beam (typical energy = 5-30 keV) and
collecting the X-ray photons thereby induced and emitted by the various elemental species. Because the wavelengths of these X-rays
are characteristic of the emitting species, the sample composition can be easily identified by recording WDS spectra (Wavelength
Dispersive Spectroscopy). WDS spectrometers are based on Bragg's law and use various moveable, shaped monocrystals as
monochromators (see: http://www.cameca.fr/html/epma_technique.htm).
The area of aluminium elemental distribution image through microprobe analysis was measured by tracing the pixels with Win/Mac
Cell V. 5.6 D software (Regent Instrument Inc., Quebec, Canada).

Figure 2.12 Electron Microprobe Analyzer (CAMECA S100, France).

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch02.html

Xiaolin Cai, 2007

Chapter 1 Wood Quality Improvement Methods and

Chapter 3 Effects of Aluminium Silicate Clays on

Technologies, and Polymer/Layered Silicate

the Morphology and Curing Behaviour of Melamine-

Nanocomposites: General Introduction and

Urea-Formaldehyde Resin

Literature Review

08-Aug-15 00:14

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Chapter 4 Formation and Properties of Nanocomposites Made up from Solid Aspen Wood, MelamineUrea-Formaldehyde, and Clay

Chapter 4 Formation and Properties of Nanocomposites Made up from Solid Aspen Wood,
Melamine-Urea-Formaldehyde, and Clay
Table des matires
Rsum
Abstract
4.1 Introduction
4.2 Experimental
4.2.1 Materials and Preparation of Impregnation Solutions
4.2.2 Impregnation Process
4.2.3 Physical and Mechanical Properties Measurements
4.2.4 Thermogravimetric Analysis (TGA)
4.2.5 Statistical Analysis
4.3 Results and discussion
4.3.1 Polymer Retention
4.3.2 Surface Hardness
4.3.3 Static Bending (MOE/MOR) Analysis
4.3.4 Wear Resistance
4.4 Conclusions
Acknowledgments

Rsum
Des nanocomposites en bois / polymre ont t prpars partir de bois massif de tremble, de rsine hydrosoluble de mlamineure-formaldhyde (MUF), et de nanoargiles de silicate. Les nanorenforts ont t moulus avec un moulin billes avant d'tre
mlangs de la rsine MUF et imprgns dans le bois. Le prpolymre hydrosoluble a t mlang aux nanoargiles une vitesse de
mlange de 3050 t/min pendant 20 minutes pour former des solutions d'imprgnation. Le bois a t imprgn de rsine qui a
polymris sous certaines conditions in-situ. Les proprits physiques et mcaniques du composite et de l'effet de la rduction de
dimension au moulin billes sur le traitement des nanorenforts ont t tudies. Des amliorations significatives des proprits
physiques et mcaniques, telles que la densit, la duret de surface, et du module d'lasticit ont t obtenues pour les spcimens
imprgns des mlanges de rsine MUF et de rsine de nanoargiles-MUF. Le traitement au moulin billes favorise la dispersion des
nanoargiles dans le bois, alors que pour certaines proprits ce traitement interfre avec l'adhrence particule-rsine.

Abstract
Wood polymer nanocomposites were prepared from solid aspen wood, water-soluble melamine-urea-formaldehyde (MUF) resin, and
silicate nanoclays. The nanofillers were ground with a ball-mill before being mixed with the MUF resin and impregnated into the
wood. The MUF resin was mixed with the nanoclays at a mixing speed of 3050 rpm for 20 min to form impregnation solutions. Wood
was impregnated with the resin, which polymerized in-situ under certain conditions. The physical and mechanical properties of the
composite and the effect of ball-milling treatment of nanofillers on the physical and mechanical properties were investigated.
Significant improvements in physical and mechanical properties, such as density, surface hardness, and modulus of elasticity were
obtained for the specimens impregnated with MUF resin and nanoclay-MUF resin mixtures. Ball-mill treatment favors the dispersion
of the nanofillers into the wood, while for certain properties interfering with particleresin adhesion.

4.1 Introduction
A shortage of high quality hardwoods has driven researchers and manufacturers to search for alternative resources, such as
softwoods and some low-density hardwoods, for value-added applications [Ayer et al 2003, Chao and Lee 2003]. Low-quality wood
can be modified through proper physical or chemical techniques to improve properties such as dimensional stability, durability,
mechanical strength and hardness, so as to meet specific end-use requirements. In addition, these treatments could improve
low-density wood mechanical properties, making them even better than hardwoods [Schneider 1994].
The formation of wood polymer composites (WPC) is a promising way to achieve this goal. Efforts to reinforce solid wood with
polymers have been extensive in the past decades. Both thermoplastic and thermosetting systems used for this purpose showed
limitations [Ayer et al 2003, Deka and Saikia 2000, Kumar 1994, Schneider 2000]. Thermoplastic-related monomers such as
acrylates or methacrylates, for instance, do not efficiently improve the dimensional stability because these monomers do not
penetrate the cell wall. Many polymers, even though formed in situ, only filled the empty lumens, which leads to a mixture of two
materials rather than to a true composite. Accordingly, the resulting products were still subject to dimensional changes with water

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

uptake. Wood impregnated with thermosetting resins such as water-soluble phenolic resins, urea-formaldehyde and melamineformaldehyde prepolymers show improvements in its compressive strength properties and moisture-related shrinking and swelling
behaviors [Schneider 2000]. These thermosetting resins can penetrate and bulk the cell wall, and may react with the hydroxyl groups
of the wood [Gindl and Gupta 2002]. However, the resulting wood-polymer composites may be more brittle, and show marginal
improvement in wear/scratch resistance.
Nanocomposite technology with organophilic layered silicates as in situ nanoreinforcement offers new opportunities for the
modification of thermoset properties. Essential improvements of physical and mechanical properties including tensile modulus and
strength, flexural modulus and strength, thermal stability, flame resistance, and barrier resistance have been observed for various
thermoplastic and thermoset nanocomposites at low silicate content [Qutubuddin and Fu 2002, Ray and Okamoto 2003]. Barrier
properties, fire resistance and mechanical properties are of great importance for the successful application of selected wood products.
Nanotechnological preparation of WPCs could be a promising new approach to obtain better products. Few attempts, however, have
been made in this regard.
The main objective of this work is to evaluate the feasibility of preparing wood polymer nanocomposites from solid wood. Trembling
aspen wood was selected as raw material. This wood is abundant, has low density and may be modified for value added applications.
The melamine-urea-formaldehyde (MUF) prepolymer was chosen for impregnation. It is a typical thermoset resin, the potential of
which to improve solid wood properties such as water absorption, hardness, and modulus of elasticity (MOE) has already been
demonstrated. Moreover, MUF resin-modified wood products retain their natural color compared with those obtained by phenolformaldehyde (PF) and other thermosets. Three kinds of aluminosilicate nanofillers (NF), with different hydrophobicity properties,
named Cloisite 30B (NF1, hydrophobic), Claytone APA (NF2, hydrophobic) and Cloisite Na+ (NF3, a pristine nanoclay,
hydrophilic) were investigated. Table 4.1 summarizes the physical properties of these montmorillonite nanofillers. Dispersion of
nanoparticles into a polymer matrix is crucial to prepare a nanocomposite. Aiming at this, the fillers were further ground with a
ball-mill before being mixed with the MUF resin andimpregnatedinto the wood. The resulting physical and mechanical properties of
the WPCsand the effect of ball-milling treatment of nanofillers on these properties were investigated.

4.2 Experimental
4.2.1 Materials and Preparation of Impregnation Solutions
Trembling aspen (Populus tremuloides) logs were provided by a local forest producer in Ste-Foy (Qubec, Canada). The green log was
cut into lumber and was dried in a kiln at temperatures from 54C to 71C for 6-8 days. Wood samples were chosen from defect-free
boards. End-matched samples for both the modified and the control samples were prepared with dimensions of 7.5(L)5(T)1.25(R)
cm3 for Brinell hardness measurements, 10101.25 cm3 for wear resistance evaluations, and 201.251.25 cm3 for static bending
tests. The urea, formaldehyde and melamine used for MUF resin cooking were industrial samples. The low viscosity MUF resins, with a
solid MUF content around 50%, were prepared and applied as described in the literature (Kim 2001).
Nanoclays Cloisite 30B (NF1), Claytone APA (NF2) and Cloisite Na+ (NF3) were provided by Southern Clay Products, Inc. (USA).
The average particle size of the clay powder received is 8 m, and each nanoparticle is made up of at least 3,000 platelets.
Ball-milling treatments of above clays were performed with a ball mill (U.S. Stoneware) equipped with a 2-literceramic jar, in which
20 porcelain balls (Fisher Scientific) with a diameter of 2.5 cm were added. 100 g of the clays were put into the ceramic jar without
water and solvent and milled for 48 h at a speed of 162 rpm for the rolls, and 50 rpm for the jar. The impregnation solutions were
prepared by adding 1% (wt/wt) layered aluminosilicate nanofiller for both the ball-mill treated nanoclays (coded as NFx-m, where x
=1, 2, 3) and untreated nanoclays (NF1, NF2 and NF3) into the low viscosity MUF resin at a mixing speed of 3050 rpm for 20 min in
a Desintegrator from Mitchel (Canada). For comparison, both the ball-mill treated and untreated nanoclays were mixed with the MUF
prepolymer to form impregnation solutions that were subsequently used to impregnate the solid aspen wood samples under specific
conditions.

4.2.2 Impregnation Process


The solid aspen wood samples, at approx. 8% MC, were oven dried to constant weight at 103C for 24 h. They were then placed into
an impregnation container, in which there was no contact between samples, and they were covered completely by MUF resin or
nanoclay/MUF mixtures. The flotation was impeded by weights on the samples. The system was evacuated to 27 mmHg for 15 min.
After that, compressed air was applied to the system and maintained at a pressure of 0.52 MPa for 15 min, then released. The excess
chemicals were wiped off the samples. Specimens were weighed and dried in fume hood for 24 h by air circulation, followed by oven
drying at 70C for 24 h. The samples were polymerized in a compression mold press at 140C for 20 min. The resulting WPC samples
were then sanded to remove excess polymer from the surface. A minimum of 20 specimens were used for the impregnation of each
treatment combination.

4.2.3 Physical and Mechanical Properties Measurements


For comparison, the density profiles of both untreated and chemically impregnated wood samples were measured and automatically
recorded using X-Ray density profile equipment. The density profiles for the impregnated wood were obtained from samples that were
cured under procedures described in the above impregnation process section. The retention of the polymer system (MUF resin or the
nanoclay/MUF) was determined according to the weight change before and after impregnation.
The hardness was measured according to the European Standard EN 1534, using Alliance RT50 systems from MTS Systems
Corporation (Eden Prairie, MN 55344-2290 USA). The hardness modulus was calculated as the slope of load vs. indentation within the
20% to 60% indentation range. Three points were tested for each sample and at least 9 samples were tested for each combination.
The modulus of elasticity (MOE) and modulus of rupture (MOR) for both treated and untreated wood samples were evaluated

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

according to ASTM D143 using a three-point bending test. The wear resistance of the treated and untreated samples was estimated
according to ASTM D-4060-95, where an abrasive wheel was mounted on a test wood specimen and rotated for a specified number of
cycles, e.g. at 100, 300 and 500 cycles. The weight losses of the specimens after rotation were determined and compared. The wear
resistance index was calculated as: Wear resistance index = (W1-W2)/number of cycles 100, where W1 and W2 correspond to the
mass of a sample before and after abrasion, respectively.

4.2.4 Thermogravimetric Analysis (TGA)


The thermogravimetric analysis (TGA) was conducted with a TGA/SDTA851 (TA Instruments) under air atmosphere at a heating rate
of 20C min-1 from 30C to 1000C.

4.2.5 Statistical Analysis


The effects of different treatments on the polymer loading, static bending strength, wood hardness and wear resistance were
determined by Tukeys Studentized and t-test Range Test (SAS, 1999).

4.3 Results and discussion


4.3.1 Polymer Retention
The treated wood was put in the fume hood for 24 h at room temperature to evaporate the water contained in the MUF resin after
impregnation. Then the samples were put into an oven at 70C for about 24 h to evaporate the remaining water and polymerized in
situ at 140C for 20 min to obtain WPC composites via a hot press procedure. Figure 4.1 shows the density profiles of the treated
wood samples with different treatments after polymerization and that of the control samples, and their average densities are listed in
Table 4.2. As readily visible, wood treatment with MUF and nanoclay significantly increased the densities. The resins are both in the
cell wall and in the vessels, as observed through scanning electron microscopy (SEM) [Cai et al 2007b]. Table 4.2 shows more details.
It is obvious that wood treated with pure MUF resin has a mean loading or weight gain of 72.7%. For the samples impregnated with
the hydrophobic nanoclay (either ball-mill treated or untreated) and MUF, similar weight gains (around 70%) can be obtained. For the
wood impregnated with hydrophilic nanofiller NF3/MUF, lower weight gains are typical, for example, 55.3% for the unmilled and
64.9% for the ball-milled wood (in both cases nanoclay/MUF impregnated). It is evident that ball-mill treatment favors the dispersion
of the nanofillers into the wood. The weight gains are higher for wood samples treated with ball-milled nanoclay/MUF as compared
with the unmilled reference. The wood polymer composite prepared from MUF and milled nanoclay such as NF1-m, NF2-m, and
NF3-m show much higher average densities than the corresponding unmilled samples (Figure 4.1). Moreover, the density profile
(Fig.4.1) of the samples impregnated with ball-milled nanofiller/MUF indicates that the nanofillers were evenly introduced into the
wood. Density profile peaks on both sides near the sample surfaces are perceptible only for the wood impregnated with unmilled
(either hydrophobic or hydrophilic) nanoclay/MUF. Especially for hydrophilic NF3/MUF formulation, the peaks of density profile on the
surface sides are much sharper than the others (Figure 4.1), which indicated high aggregation of nanoclay on the surface of treated
wood. The aggregation of nanoclay on the surface blocked the flow path of NF3/MUF resin that results the lowest average density for
NF3/MUF resin treated wood compare to all other formulations. Accordingly, the nanofiller distribution is not uniform for unmilled
nanofillers.
Thermogravimetric analysis was also performed at a heating rate of 20C/min up to 1000C. About 3% of nanofiller/wood residue
was left, if samples from the surface of treated wood (unmilled NF3/MUF) were analysed. Only 1.5% of residues were left in samples
taken from the center. The milled NF3/MUF impregnated wood, however, gives rise to 2.4% residues from the surface and about
1.8% from the center. Similar results were observed for wood samples impregnated with the hydrophobic nanoclay (NF1 or
NF2)/MUF.

4.3.2 Surface Hardness


Table 4.3 summarizes the hardness results. All formulations of wood treatment, either MUF resin alone or the combination of
nanoclay (both hydrophilic and hydrophobic) and MUF resin, displayed significant improvements in hardness compared with untreated
blanks. The best hardness corresponds to the formulation of nanoclay NF2/MUF resin mixture, where the hydrophobic nanoclay was
used directly without ball-mill treatment. The hardness of NF2/MUF impregnated wood was improved from 1.09 MPa to 3.25 MPa, i. e.
it is an almost 3-fold improvement vs. untreated wood. The improvement of wood hardness through impregnation can be attributed to
the rigid polymeric network of the cured MUF resin and to the nanofiller reinforcement by nanoclay. The increased hardness is a
result of the synergistic effect of the interaction in the system of nanoclay/MUF resin/wood.
Nanoclays have a strong influence on the hardness of the resulting WPC products. Both ball-mill treated and untreated NF3/MUF
resins caused less hardness improvement than pure MUF resins. The hydrophobic NF1/MUF yielded, on the other hand, significant
hardness increase. WPC hardness was elevated to 64% (NF2/MUF resin formulation) and to 30% (NF1/MUF formulation) compared
with pure MUF resin impregnated wood. The hardness of the wood impregnated with hydrophobic NF/MUF resins was higher in the
presence of 1 % nanoclay than that impregnated with pure MUF resin. This effect was not influenced by milling. In general, a polymer
does not intercalate into layered silicate because the layered silicate is hydrophilic and the polymer is generally hydrophobic. Thus
layered silicate was modified with alkylammonium fatty acids (NF1 and NF2), which confers hydrophobicity and increases the
d-spacing as Table 4.1 shows. However, some polymers with a hydrophilic structure such as poly(ethylene oxide) can be easily
intercalated in pristine layered silicate without modification with alkylammonium chain [Choi et al 2000].
Table 4.1 shows the layered silicate NF1 and NF2 have higher d-spacing values compare with the pristine NF3, which can help the
MUF resin better intercalate into layered silicate NF1 and NF2. The different alkylammonium chain (Table 4.1) used to modify the
layered silicate may have different compatibility with the MUF resin, which give better interaction between the layered silicate and

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

MUF resin [Cai et al 2007b]. The NF1 and NF2 combinations with MUF resin have different influence on the surface hardness and the
other properties. The hydrophobic nanofillers have better interphase interactions with the polymer matrix than NF3 filler. The latter is
hydrophilic, which helps in mixing with the MUF prepolymer, but is incompatible with the cured MUF matrix because of the polarity
change of the MUF from hydrophilic to hydrophobic after curing. This leads to poor interphase interactions of the hydrophilic
nanofiller and the cured hydrophobic polymer matrix. As a result, the WPC composites have poor mechanical properties.
The effect of ball-mill treatment of the nanofillers on hardness is less straightforward. For the hydrophobic nanofillers, as one can see
from Figure 4.2 a), ball-mill treatment of the NF2 nanofiller resulted in lower hardness than in the case of unmilled NF2/MUF resin
composites. It is supposed that the ball-mill treatment destroys some functional groups grafted on the nanofiller surfaces [Porter and
Casale 2004, Ruan et al 2004, Tanaka et al 2004], and this leads to poorer interphase interactions between the nanofiller and the
polymer matrix. However, this interpretation is not valid for the nanofiller NF1. In the case of wood impregnated with hydrophilic
nanofiller/MUF similar observation have been made, even though the weight gain was higher (Table 4.2).

4.3.3 Static Bending (MOE/MOR) Analysis


Rowell et al [1982] studied the effects of cell wall and cell lumen treatment. In the first case, propylene oxide triethylamine (POT)
was used. POT reacts with the hydroxyl groups of the cell wall. In the second case, methyl methacrylate (MMA) filled only the voids of
wood. The POT treated wood composite exhibited both lower MOE and MOR, while the MMA impregnated wood displayed both
increased MOE and MOR. The combination of POT and MMA treatment also resulted in relatively lower MOE and MOR compared to
untreated wood.
Treatment with the resins caused drastic improvements in the MOEs. (Table 4.4). The average MOE for treated samples is almost
double of that of the untreated references. A comparison of the effects of different nanoclay/MUF formulations on MOE with those of
pure MUF resin reveals no significant differences, although slightly higher MOEs are perceptible for the formulations containing NF2
(both ball-mill treated and untreated) than for pure MUF impregnated wood (Table 4.4). For the NF3/MUF formulations, lower MOEs
are observed, regardless of the milling. This could be attributed to the incompatibility between the cured MUF resin and the NF3.
The highest improvement in MOE was achieved with unmilled NF2/MUF. However, at the 95% probability level, no significant
differences were found between the MOEs concerning the effects of ball-mill treatment of nanofilllers (Table 4.4), though a tendency
of a higher MOE was observed for samples impregnated with milled hydrophilic nanofiller/MUF.
Concerning the modulus of rupture (MOR), no significant difference was found between the treated and untreated wood samples
(Table 4.4). Relatively higher MORs were obtained when pure MUF resin and unmilled organophilic nanofiller/MUF were used for
impregnation. For the hydrophilic nanofiller NF3, lower MORs were observed for samples both impregnated with ball-mill treated and
untreated NF3/MUF. Ball-mill treatment of nanofillers resulted in slightly lower MORs of nanoclay/MUF/wood composites as compared
with control samples (Table 4.4).
There is no linear or quadratic relationship visible from Figure 4.3 which depicts the relationship between MOR and MOE. However,
three different domains are marked as A, B, and C. Almost all control samples are distributed in the area A, for which relatively low
MOE and MOR data are typical. The hydrophilic nanofiller/MUF/wood composites and all wood composites impregnated with
ball-milled nanofillers (both hydrophilic and organophilic) and MUF are in region B. Here, the MOEs of treated wood composites are
higher and the MORs are lower or equal to that of untreated wood. In area C, samples are placed with improved MOE and MOR. The
positive quality effect is due to impregnation with MUF, NF1/MUF, and NF2/MUF. Usually, MOR/MOE improvements are correlated with
the increase in density. The improvement in MOR and MOE for the samples in area C can be interpreted by density improvement and
by a better interphase interaction.

4.3.4 Wear Resistance


Table 4.5 presents the mean wear resistances and Figure 4.2 b) illustrates the same values measured at 100, 300, and 500 cycles.
During the first 100 cycles, the wear resistance is the highest except for the NF3-m/MUF resin formulation. After that, wear
resistance becomes relatively low for samples impregnated with MUF resin and hydrophobic nanofiller/MUF mixtures. The wear
resistance index decreases with increased number of abrasion cycles except for the NF3-m/MUF resin formulation (Fig. 4.2 b). The
NF2/MUF resin formulation has the highest wear resistance index (592% of the control samples). The resistance of wood impregnated
with pure MUF resin was almost triple of that of the control sample.
Nanoclay/MUF formulations compared with pure MUF impregnated wood show significant increases in wear resistance with the
exception of the NF1/MUF formulations. The wear resistance index or wear loss is affected by several factors, typically, by the friction
factor, hardness, and impact strength of the material [Huang et al.1997]. Generally, wear loss increases with an increase in friction
factor and decreases with increase in hardness and impact strength. The impact strength of the treated wood was obviously reduced,
though the hardness was improved. The treated wood has a more rigid surface, higher density, and better mechanical properties.
Woods impregnated with MUF and NF2/MUF have a better wear resistance than those impregnated with NF1/MUF and NF3/MUF.

4.4 Conclusions
Wood polymer nanocomposites were prepared by impregnation of solid aspen wood with water-soluble melamine-urea-formaldehyde
(MUF) prepolymer in combination with hydrophilic and hydrophobic nanoclays. The clays were ground with a ball-mill before being
mixed with the MUF prepolymer. Significant improvements in physical and mechanical properties (density, surface hardness, MOE)
can be obtained in the case of impregnation with MUF and nanoclay/MUF. Ball-mill treatment was found to favor the dispersion of the
nanoclays into the wood. The hardness of wood impregnated with hydrophobic nanoclay/MUF was improved from 1.09 MPa to 3.25
MPa. Impregnation with either MUF or nanoclay/MUF caused essential improvement of the MOEs. The average MOE for the treated

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

wood was almost twice as much as for untreated wood. The MOR, does not differ significantly between treated and untreated woods.

Acknowledgments
The financial support of the Natural Sciences and Engineering Research Council of Canada (NSERC), the Fonds Qubcois pour la
Recherche sur la nature et les technologies (FQRT) and FPInovation Forintek (Eastern Region) for this project is gratefully
acknowledged. Also, thanks are given to Mme. Francine Cot for the measurement of the physical properties and to Dr. Chuangmin
Liu for the statistical analysis using SAS program from FPInovation Forintek (Eastern Region).

Table 4.1 Properties of nanoclays.

1 MT2EtOH:

methyl, tallow, bis-2-hydroxyethyl, quaternary ammonium; R1 through R4 = combination of aliphatic chains, methyl and

(or benzyl groups), N=nitrogen.


Table 4.2 T-test analysis of density and weight gain of MUF/nanoclay treated and untreated aspen1.

1 1Each
2 NF1,
3 The

value is the average of 10 measurements.

NF2 and NF3 are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled nanofillers.

same letters are not significantly different at =5%.

Table 4.3 t-test analysis of Brinell hardness of MUF/nanoclay treated and untreated aspen1.

1 Each

value is the average of 10 samples with minimum 3 point measurements for each sample.

2 NF1,

NF2 and NF3 are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled nanofillers.

= ((treatment-control)/control))100, is just ratio.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...


4 The

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

same letters are not significantly different at =5%.

Table 4.4 t-test analysis of modulus of elasticity (MOE) and modulus of rupture (MOR) of MUF/nanoclay treated and
untreated aspen1.

1 Each

value is the average of 10 measurements.

2 NF1,

NF2 and NF3 are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled nanofillers; _c is the control.

3 The

same letters are not significantly different at =5%.

Table 4.5 Tukey analysis of wear resistance index of MUF/nanoclay treated and untreated aspen1.

1 Each

value is the average of 10 measurements.

2 NF1,

NF2 and NF3 are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled nanofillers.

3 The

same letters are not significantly different at =5%.

Figure 4.1 Density profiles of MUF and nanoclay/MUF treated and untreated aspen wood after hot-pressing: NF1, NF2 and
NF3 are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled nanofillers.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

Figure 4.2 Mechanical properties of aspen and impregnated aspen with MUF and 1 wt % montmorillonite/MUF. a) Brinell
hardness, b) Wear resistance of MUF and nanoclay/MUF (100cycles/300cycles/500cycle) treated samples NF1, NF2 and NF3
are montmorillo

Figure 4.3 Relationship between strength at rupture and elasticity modulus (MOR/MOE) of aspen and impregnated aspen
with MUF and montmorillonite/MUF. NF1, NF2 and NF3 are montmorillonite nanofillers; NF1-m, NF2-m and NF3-m are milled
nanofillers.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch04.html

Xiaolin Cai, 2007

Chapter 3 Effects of Aluminium Silicate Clays on the

Chapter 5 Effects of Nanofillers on Water Resistance

Morphology and Curing Behaviour of Melamine-

and Dimensional Stability of Solid Wood Modified by

Urea-Formaldehyde Resin

Melamine-Urea-Formaldehyde Resin

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Chapter 5 Effects of Nanofillers on Water Resistance and Dimensional Stability of Solid Wood Modified by MelamineUrea-Formaldehyde Resin

Chapter 5 Effects of Nanofillers on Water Resistance and Dimensional Stability of Solid Wood
Modified by Melamine-Urea-Formaldehyde Resin
Table des matires
Rsum
Abstract
5.1 Introduction
5.2 Experiment
5.2.1 Materials and Sample Preparation
5.2.2 Moisture Absorption
5.2.3 Water Absorption and Dimensional Stability Measurements
5.2.4 Anti-Swelling-Efficiency Measurements
5.2.5 Scanning Electron Microscopy (SEM) measurement
5.2.6 Electron Probe Micro-Analysis (EPMA)
5.2.7 Color Measurement
5.3 Results and Discussion
5.3.1 Microscopy Investigations
5.3.2 Moisture Absorption
5.3.3 Water Absorption
5.3.4 Dimensional Stability
5.3.5 Color Variations
5.4 Conclusions
Acknowledgments

Rsum
L'absorption de l'eau et la stabilit dimensionnelle du bois imprgn de mlamine-ure-formaldhyde (MUF) et du bois imprgn avec
diffrentes formulations de nanorenfort/MUF ont t tudies. Trois genres de nanoparticules, Cloisite 30B, Claytone APA et
Cloisite Na+, ont t choisis et mlangs de la rsine MUF, et plus tard imprgns dans le bois massif de tremble par un processus
de vide et de pression. Les nanocomposites en bois / polymre ont t prpars par polymrisation in situ par condensation dans le
bois imprgn dans des conditions spcifiques. Des amliorations significatives de la rsistance l'eau et une meilleure stabilit
dimensionnelle ont t obtenues pour le bois trait par nanorenfort/MUF. Le bois non trait a absorb autour de 63% d'humidit
aprs 24 h de trempage dans l'eau tandis que labsorption d'eau tait denviron 125% aprs 1 semaine dimmersion dans l'eau. Le
bois imprgn de rsine MUF a absorb environ 8.3% et 38.5% d'humidit aprs 24 h et une semaine dimmersions dans l'eau,
respectivement. Pour le bois imprgn de rsine hydrophobe de nanoargiles/MUF, on a observ une absorption deau bien infrieure,
soit environ 5% sur 24 h et 22% aprs une semaine. L'efficacit antigonflement (ASE) a t galement amliore de 63.3%
125.6% pour le bois trait par nanorenfort/MUF. L'amlioration significative de la rsistance l'eau et de la stabilit dimensionnelle
des nanocomposites en bois avec de tels polymres peut tre attribue l'introduction de MUF et de-nanorenforts dans le bois. La
fluorescence au rayon-X montre que beaucoup de nanoparticules ont migr dans la paroi cellulaire. Les traitements du bois avec MUF
et nanorenfort/MUF n'ont montr aucune influence significative sur la couleur du bois, ce qui est important pour les applications
pratiques du bois trait dans certains secteurs spcifiques tels que les planchers et les parquets.

Abstract
The water absorption and dimensional stability of wood impregnated with melamine-urea-formaldehyde (MUF) and wood impregnated
with different nanofiller/MUF formulations were investigated. Three kinds of nanoparticles, Cloisite 30B, Claytone APA and
Cloisite Na+, were selected and mixed with MUF resin, and subsequently impregnated into solid aspen wood through a vacuum and
pressure process. The wood polymer nanocomposites were prepared by in situ condensation polymerization of the impregnated wood
under specific conditions. Significant improvements in water repellency and better dimensional stabilities were obtained for the
nanofiller/MUF treated wood. The untreated wood absorbed around 63% of moisture after 24 h soaking in water while water uptake
was about 125% after one week immersion in water. The MUF resin impregnated wood absorbed about 8.3% and 38.5% of moisture
after 24 h and one week immersion in water, respectively. For the organophilic nanoclay/MUF resin-impregnated wood, much lower
water absorption in the amounts of around 5% water uptake in 24 h and 22% after one week were observed. The antiswelling
efficiency (ASE) was also improved from 63.3% to 125.6% for the nanofiller/MUF treated wood. The significant improvement in water
resistance and dimensional stability of the resulting wood polymer nanocomposites can be attributed to the introduction of MUF and
nanofillers into the wood. X-ray fluorescence shows that some nanoparticles have migrated into the wood cell wall. Wood treatments
with MUF and nanofiller/MUF showed no significant influence on the color of the wood, which is important for practical application of
the treated wood in some specific areas such as flooring.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

5.1 Introduction
Wood is a three-dimensional polymeric material, made up mainly of cellulose, hemicellulose and lignin. These constituents are
responsible for most of the physical and chemical properties of wood. Because of its strong physical strength, aesthetically pleasant
characteristics and low processing cost, wood is an important building and industrial raw material. But wood has some drawbacks such
as high moisture uptake, biodegradation and dimensional change with environmental variations [Kumar 1994, Galperin et al 1995],
which limit its use. These defects have primarily been ascribed to the presence of numerous hydroxyl groups in the three major wood
components (cellulose, hemicellulose, and lignin) and its various cavities. The hydroxyl groups attract water molecules through
hydrogen bonding from the surrounding environment, causing swelling. The cell cavities and lumens are major paths for moisture
movement. This process is reversible; wood shrinks as it loses moisture upon drying.
The dimensional changes of wood due to atmospheric moisture can be minimized by appropriate chemical treatments such as the
formation of wood polymer composites (WPC), which is a promising way to improve wood properties [Schneider et al 1991, Hartley
and Schneider 1993, Rowell 1983]. Attempts to reinforce solid wood with polymers have been extensive in the past decades. Both
thermoplastic and thermosetting systems have been used and have achieved certain improvements in wood properties, but both
showed limitations [Kumar 1994, Deka and Saikia 2000, Schneider 2000, Ayer et al 2003]. Thermoplastic-related monomers such as
acrylates or methacrylates, for instance, do not improve the dimensional stability because these monomers do not penetrate the cell
wall or do not react with the wood. Many polymers, even though formed in situ, only filled the empty lumens in the wood, which leads
to a mixture of two materials rather than a real composite. Accordingly, the resulting products were still subject to dimensional
changes with water uptake.
Nanocomposite technology with organophilic layered silicates as in situ nano-reinforcement offers new opportunities for the
modification of polymeric material properties and has been intensively investigated in recent years [Schadler 2003]. Polymer/layered
silicate nanocomposites were first developed based on thermoplastic polyamide 6 [Kojima et al 1993]. Since then, a great deal of
research has been carried out in the field of polymer nanocomposites using various thermoplastic and thermoset polymers over the
past decade [Schadler 2003; Ray and Okamoto 2003]. Essential improvements of physical and mechanical properties including tensile
modulus and strength, flexural modulus and strength, thermal stability, flame resistance, and barrier resistance have been observed
for various polymer nanocomposites of low silicate content [Byun et al 2001, Qutubuddin and Fu 2002, Schadler 2003]. Barrier
properties, fire resistance and mechanical properties are of great importance for the successful application of selected wood products.
Nanotechnological preparation of WPCs could be a promising new approach to obtain better products. Little work, however, has been
devoted to the preparation of a wood polymer composite using nanotechnology. This led us to investigate the feasibility of preparing a
wood polymer nanocomposite through impregnation of nanoparticles into the wood. In a previous work [Cai et al 2007a], we reported
on a procedure to prepare wood polymer nanocomposites from solid aspen wood with impregnation of a MUF resin and different
layered aluminosilicate nanofillers. Significant improvements in various physical and mechanical properties of the resulting
nanocomposites such as surface hardness and modulus of elasticity were found. In this paper, we have investigated the location of
MUF resin and nanofillers in the wood and the effects of nanofillers on water absorption, dimensional stability and color of the
resulting wood polymer nanocomposites.

5.2 Experiment
5.2.1 Materials and Sample Preparation
Trembling aspen (Populus tremuloides), an abundant, low density and fast growing wood logs, used in this work were provided by a
local forest farmer from Qubec City area (Qubec, Canada). The green log was cut into lumber and low temperature kiln-dried for 2
weeks. Wood samples were chosen from defect-free boards. End-matched samples with dimensions in longitudinal, tangential and
radial directions of 7.551.25 cmwere prepared. The urea, formaldehyde and melamine used for MUF resin cooking were industrial
samples. Low viscosity MUF resins, with a solid MUF content of around 50%, were used as prepolymers to impregnate into the wood
samples and were prepared as described else where [Cai et al 2007d; Kim 2001]. The layered aluminosilicate nanofillers (NF),
specifically Cloisite 30B (NF1, organophilic, density, 1.98 g/cm3), Claytone APA (NF2, organophilic, density, 1.70 g/cm3) and
Cloisite Na+ (NF3, a pristine nanoclay, hydrophilic, density, 2.86 g/cm3) were obtained from Southern Clay Products, Inc. (USA),
and were used as received. The average particle size of the clay powder received is 8 m, and each nanoparticle contains at least
3,000 platelets. The impregnation solutions were prepared by adding 1% (wt/wt) nanoclay into the low viscosity MUF resin at a
mixing speed of 3050 rpm for 20 min to form suspensions that were subsequently used to impregnate the solid aspen wood samples
under specific conditions. A minimum of 20 specimens was used for the impregnation of each combination. The wood polymer
composites and nanocomposites were prepared by the procedure described in our previous work [Cai et al 2007a, b]. In short, a low
viscosity MUF solution containing 1% nanoparticules, of mean 100 nm in length and 10 nm in thickness, was impregnated into the
wood and cured in situ at 140C for 20 min in a compression mold press.

5.2.2 Moisture Absorption


The controls and treated wood samples were oven dried at 103C for 24 hours. They were then placed in a conditioning chamber at a
temperature of 21C and a relative humidity of 65% for approximately 6 weeks. After 6 weeks, the weight of each sample was
measured. The moisture content (MC) was calculated as follows:

(5.1)
where M

is the mass of the aspen block after 6 weeks condition, and M

is the oven-dried mass of the control or treated aspen

block sample.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

5.2.3 Water Absorption and Dimensional Stability Measurements


Water absorption and dimensional stability for both the treated wood specimens and the control samples were measured according to
ASTM-D037. All of the samples were conditioned as described above. The mass gain and dimensional changes of each sample were
measured and recorded (denoted as the dry data). After that, the conditioned specimens were immersed in distilled water at a
temperature of 201C for either 24 h or 1 week, when the specified samples were taken out. The excess water on the soaked
samples was wiped off, and the weight and dimensional changes of the wood were recorded (the wet data). The dimensional swelling
and water absorption of the samples were calculated using the dry and the wet data obtained. The swelling coefficient and
anti-swelling efficiencies in the longitudinal, radial and tangential directions were also calculated using data obtained for the wood
polymer nanocomposites and corresponding data for the control samples. For each combination, there were at least 20 values for the
swelling coefficient and anti-swelling efficiency at each direction (longitudinal, radial or tangential).
The water repellency efficiency was calculated by Eq. (5.2):
(5.2)
where Wc is the water absorption of a wood control sample and Wt is the water absorption of a treated sample, which was calculated
by Eq. (5.3):
(5.3)
where w0 is the initial weight of an oven-dried sample, and wi is the weight after water immersion for 1 day and 7 days at 21C,
respectively.

5.2.4 Anti-Swelling-Efficiency Measurements


The anti-swelling-efficiency (ASE) was determined after test samples were soaked in water at 21C at a water flow rate of 20 ml/s for
7 days. The volumetric swelling coefficients were calculated according to the formula:
(5.4)
where V

is the volume of the water saturated wood and V

is the sample volume of the dry untreated and treated wood,

respectively.
The percentage of ASE was calculated for the wet and oven dried volumes of the treated and untreated blocks according to:
(5.5)
where S

is the volumetric swelling coefficient of the control samples and S

is the volumetric swelling coefficient of the treated

samples.

5.2.5 Scanning Electron Microscopy (SEM) measurement


The morphologies of the treated and untreated wood samples in transverse surface were observed and analyzed using a Jeol
JSM-840A scanning electron microscope (SEM). The pure wood, MUF- and nanofiller/MUF-treated wood blocks were prepared with a
razor blade mounted onto a microtome by carefully cutting one of the end-grain surfaces to a depth of about 3 mm. All blocks were
desiccated with phosphorus pentoxide for 2 weeks. A gold/palladium alloy was sputtered on the cutting surfaces prior to the
investigation.

5.2.6 Electron Probe Micro-Analysis (EPMA)


Electron microprobe technique is one of the most precise and accurate methods for surface elemental analysis. The distribution of
nanofillers in the wood for the nanofiller/MUF treated wood samples was investigated using an Electron Microprobe Analyzer
(CAMECA S100, France). The samples used for the EPMA analysis were prepared using a way similar to the SEM samples before
coating. The cross-section of the wood samples was coated with carbon (25 nm) prior to the measurements in order to make it
conductive and to protect the surface from beam damage to some extent. Electron microprobe analysis was performed in a mapping
mode with an accelerating voltage of 15 kV and 10 nA. The images of elements Al and Si, which are mainly from aluminosilicate
nanofillers for the nanofiller/MUF impregnated wood, were captured digitally to allow for enhanced analysis of the samples.

5.2.7 Color Measurement


The colors of the control, MUF- and nanoclay/MUF-treated wood samples were measured using a Color-guide BYK-Gardner GmbH
equipment (Germany), according to the procedure of the Color-guide 45o/0o, Kat. Nr./Cat. No.6805. The color change of each sample
was calculated by Eq. (5.6):

(5.6)
Where L

is the lightness, a

= + red/ - green, b

= + yellow/ - blue, = sample standard. E

defines the total color difference

between sample and standard samples.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

5.3 Results and Discussion


5.3.1 Microscopy Investigations
Figure 5.1 shows micrographs of the untreated wood (Fig. 5.1(a)) and treated wood samples (Fig. 5.1(b), (c), (d) and (e)) in a typical
transversal section. It is clear that good penetration and adhesion of the MUF resin to the cell wall was observed for the treated wood.
For the NF3/MUF treated wood, relatively lower polymer loading (nanofiller/polymer content, 35.6% by weight) was obtained, and
the impregnated materials seem to be located mainly in the cell wall also a lot in the cell lumen (Fig.5.1 (e)). Higher weight gains
were achieved for other treatments with MUF (polymer content, 42.1%) or the organophilic nanoclay/MUF formulations (NF1/MUF,
40.9% and NF2/MUF, 40.6%), and the impregnated materials are found in both the cell wall and vessels of the wood. Figure 5.2
illustrates the electron microprobe analysis results of the nanofiller/MUF treated wood samples. Figure 5.2(a) shows, for example, the
cross-section morphology for the NF2/MUF treated wood, and Figure 5.2 (b) shows the captured Al and its distribution in the
corresponding section of the sample. Similar results were also observed for other aluminosilicate nanofiller/MUF treated wood
samples. From Figure 5.2(b), the middle lamella and cell wall have the highest concentration of Al, implying that the aluminosilicate
nanofillers diffuse in the lamella and the cell wall. Work has also been done through transmission electron microscopy (TEM) analysis
of the cell wall of both the control and nanofiller/MUF treated samples, confirming that nanoparticles have been impregnated into the
cell wall [Cai et al 2007c]. Thus for the nanofiller/MUF treated wood, it could be said that both MUF resin and the nanofillers were
successfully impregnated into the wood, and into the cell wall. This not only result in enhanced mechanical properties of the wood,
but contribute to decrease water absorption and increase dimensional stability of the treated wood as well. In a previous paper [Cai et
al 2007a], it has been shown that mechanical properties of such composites were much better than untreated wood, but diffusion of
nanoparticles in the cell wall had not been discussed.

5.3.2 Moisture Absorption


The moisture absorption (MA) values for the control and treated samples after condition (21C, 65% R.H.) were measured and are
summarized in Table 5.1. From Table 5.1, it is MUF resin and different nanoclay/MUF mixture treated wood samples exhibited much
lower MAs than did control samples. For untreated wood, around 8.5% water adsorption was observed, while less than half as much
water adsorption occurred in the wood containing MUF or nanoclay/MUF, i.e. 2.8% for the MUF-treated wood, and about 3.1%, 3.9%
and 2.5%, for the NF1/MUF, NF2/MUF and NF3/MUF treated wood samples, respectively. The reduction in moisture absorption for the
treated wood could be attributed to the impregnated polymer MUF and/or the nanoclay/MUF, which may block sorption sites on the
interior of wood cell lumens and in the cell walls [Gindl and Gupta 2002].
From Table 5.1, it is also apparent that different combinations of nanofiller/MUF treatment provide different levels of protection from
moisture absorption. It is interesting to see that the hydrophilic nanoclay NF3/MUF-treated wood showed relatively lower moisture
absorption. The nanofiller NF3 is hydrophilic, which makes it compatible with the MUF prepolymer during the preparation process of
the impregnation solution. However, this makes it incompatible with the cured MUF matrix, because the polarity of the MUF resin is
changed from hydrophilic to hydrophobic during the curing process. This could result in poor interphase interactions of the hydrophilic
nanofiller and the cured hydrophobic polymer matrix, causing poor mechanical properties of the resulting WPC composites. It has
been emphasized from our previous work that lower surface hardness and MOE were observed for the hydrophilic nanofiller/MUFtreated wood samples than those of the organophilic nanoclay/MUF modified wood [Cai et al 2007a]. Poor dispersion of NF3 into the
wood was also observed as shown by large amount of NF3 nanoclay, as aggregates, left on the surface of the NF3/MUF treated wood
samples. Even though poorly-dispersed nanoclays could result in poor mechanical properties of the resulting WPC composites, the
higher concentration of the accumulated nanoclays on the top skin of the sample may inhibit moisture absorption on the surface and
delay the moisture diffusion in and through the wood. Thus better moisture repellency would be expected. For the organophilic
nanofillers (NF1 and NF2), better dispersion behavior was observed, and better interphase interactions of the organophilic nanofillers
with the polymer matrix were expected and achieved as observed in our previous work [Cai et al 2006a], in Figure 5.1 and Figure
5.2. The difference in the effect of different nanofillers on moisture absorption of treated samples, however, was not obvious as
compared to the MUF treated samples.

5.3.3 Water Absorption


The water absorption of the treated and untreated wood samples in water at 21C after 1 day and 7 days was measured and results
are given in Table 5.2. As one can see, the untreated wood absorbed around 63% of its weight water after a 24 h soaking and about
125% after 1 week immersion. The MUF-impregnated wood absorbed about 8.3% and 38.5% after 24 h and 1 week immersion,
respectively. Thus, MUF treatment significantly reduces water absorption of wood. For the nanofiller/MUF-impregnated wood, much
lower water absorption was observed, e.g. around 5% in 24 h and 22% in 1 week for the NF2/MUF-impregnated wood. Similar water
absorption behaviors could also be observed for wood composites made with other nanofillers (both organophilic and hydrophilic:
Table 5.2).
Figure 5.3 illustrates the water repellency efficiencies (WREs) of MUF and nanofillers/MUF-treated wood samples. The WREs of
MUF-treated wood samples were 63.8% and 59.6% for 24 h and one week immersions, respectively. The WREs of nanofillers/MUFtreated wood samples were between 90%-92% for 24 h and between 78%-83% for 1 week water immersion, respectively. The WREs
of nanofillers/MUF-treated wood samples were much higher than that of pure MUF-treated wood samples for both 24 h and 1 week
water immersions. A significant improvement in water repellency was achieved for the nanofiller/MUF impregnated wood. The WRE,
however, was not significantly different between the hydrophilic (NF3) and the organophilic (NF1 and NF2) nanofiller/MUF
treatments. Several reasons could be ascribed to the decrease in water absorption for the treated wood. First, the cured hydrophobic
polymer MUF shields the wood surfaces and remains in the cell wall and lumen, resulting in less water penetration into the wood.
Secondly, the MUF resin may react with the numerous hydroxyl groups contained in the wood components [Gindl and Gupta 2002],
thus fewer water absorption sites would remain, which would also contribute to the reduction in water uptake by the treated wood.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

For the nanoclay/MUF-impregnated wood, since the nanolayers themselves are impermeable, the impregnated nanolayers could
generate a tortuous pathway for a permeant, e.g. the water molecules, to diffuse through the nanocomposite. The hindered diffusion
of a permeant in the wood could lead to enhanced barrier properties, reduced swelling behavior, and probably improvement in
chemical stability and flame retardance as well [Gilman 1999].

5.3.4 Dimensional Stability


The results of thickness swelling (in radial direction) of MUF- and nanofillers/MUF-treated wood are given in Table 5.3. From Table
5.3, all treatments significantly decreased the thickness swelling of the wood. The untreated wood showed a thickness swelling of
3.8% and 4.1% after 24 h and 1 week soaking in water, respectively. For the MUF-treated wood, a thickness swelling of 2.1% for 24
h and 3.0% for 1 week was obtained. Compared to the pure MUF-treated wood samples, much lower thickness swelling for
nanoclay/MUF-treated wood was observed, e.g. 0.75% in 24 h and 1.81% in 7 days for the NF3/MUF-treated wood. Thus the addition
of nanofillers into MUF significantly decreased thickness swelling. The lowest thickness swelling of 0.54% for 24 h and 1.50% for 1
week immersion, respectively, was achieved with the combination of NF2/MUF.
Table 5.4 gives the results for tangential swelling for the control, MUF- and three formulations of nanofiller/MUF-treated wood
samples. Similar swelling behavior was seen in the radial direction. For the untreated wood, the highest values for swelling in the
tangential direction of 5.29% for 24 h and 6.37% for 1 week immersion in water were observed. Much lower swelling was achieved
for both MUF- and nanofillers/MUF-treated wood samples, e.g. 1.05% for NF3/MUF treated wood samples after 24 h soaking (Table
5.4). Comparing the pure MUF- and nanofiller/MUF-treated wood samples, better results were achieved for the formulations
containing nanofillers. For instance, NF2/MUF gave the lowest swelling in the tangential direction of 2.84% after 1 week water
immersion, which was significantly lower than the 6.37% for the control samples and 5.06% for pure MUF-treated wood.
Table 5.5 gives the results of swelling in the longitudinal direction. Although minimal, the swelling of the control samples was still
significantly higher than those of MUF- and nanofiller/MUF-treated wood samples. The difference in the longitudinal swelling between
the pure MUF resin-treated and the nanofiller/MUF-treated wood samples is not significant for 24 hours water immersion, but
significant for 1 week water immersion, which indicates that nanofillers decreased the swelling in the longitudinal direction as well.
Figure 5.4 (a) gives ASEs in tangential and radial directions for both pure MUF and nanofillers/MUF-treated wood samples. Compared
to the pure MUF-treated wood, all the nanofiller/MUF-treated wood samples have increased ASE, both in tangential and radial
directions. The improvement in ASE, in both tangential and radial directions of nanofillers/MUF treated wood samples, is significant
compared to pure MUF resin-treated wood samples. The effect of different nanofillers/MUF combinations on ASE, both in tangential
and radial directions, were not significantly different from each other, even though the ASE for NF2/MUF-treated wood was slightly
higher than that of NF1/MUF- and NF3/MUF-treated wood samples. The results of ASE in volume are given in Figure 5.4(b). The ASE
of the MUF-treated wood samples was 63.3%, while the ASEs of NF1/MUF- NF2/MUF- and NF3/MUF-treated wood samples were
102.9%, 125.6% and 94.2%, respectively. All the nanofillers/MUF treatments improved ASE significantly as compared to the pure
MUF resin, which means that better dimensional stability could be obtained using nanofiller/MUF impregnated wood.

5.3.5 Color Variations


The color change is an important factor for the final using of the wood polymer comoposites, for example, WPCs as wood flooring
materials. The characterization of color change of each sample after wood treatment with MUF and/or nanofiller/MUF was
investigated. Table 5.6 gives differences in color change of the samples as compared to the controls. As one can see all the color
variation values are less than 5. This means that wood treatments with MUF and nanofiller/MUF showed no significant influence on
the color of the wood, which is important for practical application of the treated wood in some specific areas such as flooring.

5.4 Conclusions
Wood polymer composites and nanocomposites were prepared through impregnation of melamine-urea-formaldehyde (MUF)
prepolymer and different nanoclay/MUF formulations into aspen. Significant improvements in water repellency and better dimensional
stabilities were obtained for the nanoclay/MUF-treated wood samples. The untreated wood absorbed around 63% of its weight in
water after 24 h of soaking and about 125% after 1 week immersion. The MUF resin impregnated wood absorbed about 8.3% and
38.5% after 24h and 1 week immersion, respectively. For the nanoclay/MUF resin-impregnated wood samples, much lower water
absorption was observed, e.g. around 5% water uptake in 24 h and 22% in 1 week for the organophilic nanoclay/MUF impregnated
samples. Antiswelling efficiency was also improved from 63.3% to 125.6% for the nanoclay/MUF treated wood. The significant
improvement in water resistance and dimensional stability of the resulting wood polymer nanocomposites could be attributed to the
introduction of MUF and nanofillers into the wood. Wood treatments with MUF and nanoclay/MUF showed no significant influence on
the color of the wood, which is important for practical application of the treated wood in some specific area such as flooring. Finally,
with high resolution elemental mapping and our further TEM observation, some nanoparticules used in MUF formulations have been
shown to diffuse into the wood cell wall.

Acknowledgments
The financial support of the Natural Sciences and Engineering Research Council of Canada (NSERC), le Fonds qubcois de la
recherche sur la nature et les technologies (FQRNT) and FPInovations Forintek (Eastern Region) for this project is gratefully
acknowledged.

Table 5.1 Moisture absorption and moisture content at equilibrium (at 21C, 65% MC for 6 weeks) of MUF and nanoclay/MUF
treated and untreated aspen1.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

1 Each
2 The

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

value is the average of 10 specimens.

same letters are not significantly different at =5%.

Table 5.2 Water absorption in one day and one week of MUF and nanoclay/MUF treated and untreated aspen1.

1 Each
2 The

value is the average of 10 specimens.

same letters are not significantly different at =5%.

Table 5.3 Thickness swelling (radial) after 1 day and 1 week in water of MUF and nanoclay/MUF treated and untreated
aspen1 samples.

1 Each
2 The

value is the average of 10 specimens.

same letters are not significantly different at =5%.

Table 5.4 Swelling in width (tangential) (1 day and 1 week) for MUF and nanoclay/MUF treated and untreated aspen1
samples.

1 Each
2 The

value is the average of 10 specimens.

same letters are not significantly different at =5%.

Table 5.5 Swelling in length (1 day and 1 week) for MUF and nanoclay/MUF treated and untreated aspen1 samples.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

1 Each
2 The

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

value is the average of 10 specimens.

same letters are not significantly different at =5%.

Table 5.6 Color variation values of the MUF- and nanofiller/MUF-treated wood samples as compared to the controls.

Figure 5.1 Scanning electron microscopy photographs of (a) control wood samples, (b) MUF-treated wood sample,
containing 42.1% polymer, (c) NF1/MUF-treated wood sample, NF1/MUF content 40.9%, (d) NF2/MUF-treated wood sample,
containing 40.6% NF2/MUF, and (e) NF3/MUF-treated wood sample, containing 35.6% NF3/MUF.

Figure 5.2 Electron Microprobe Analysis of NF2/MUF treated wood, (a) transversal face, and (b) Al elemental distribution.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

Figure 5.3 Water repellency efficiency.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch05.html

Figure 5.4 Antiswelling efficiency, (a) tangential direction and radial direction, (b) ASE in volume.

Xiaolin Cai, 2007

Chapter 4 Formation and Properties of


Nanocomposites Made up from Solid Aspen Wood,

Chapter 6 Impact of the nature of nanofillers on the


Performances of Wood Polymer Nanocomposites

Melamine-Urea-Formaldehyde, and Clay

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Chapter 6 Impact of the nature of nanofillers on the Performances of Wood Polymer Nanocomposites

Chapter 6 Impact of the nature of nanofillers on the Performances of Wood Polymer


Nanocomposites
Table des matires
Rsum
Abstract
6.1 Introduction
6.2 Experiment
6.2.1 Materials
6.2.2 Water Absorption
6.2.3 Microscopic Analysis of the WPCs
6.2.4 Thermogravimetric and Dynamical Mechanical Properties Measurements
6.3 Results and Discussion
6.3.1 Ball-milling Treatment of Nanofillers
6.3.2 Enhanced Properties of the WPC Composites
6.3.3 Improvement in Water Absorption
6.3.4 Microscopy investigations
6.3.5 Dynamic Mechanical Property Investigations
6.4 Conclusions
Acknowledgements

Rsum
Des nanocomposites bois / polymre ont t prpars partir dchantillons en bois massifs de tremble en utilisant la mlamineure-formaldhyde hydrosoluble (MUF) et de deux genres de nanorenforts d'aluminosilicate, un tant hydrophile et l'autre
hydrophobe. Les silicates ont t moulus avec un moulin billes avant d'tre mlangs de la rsine MUF et tre imprgns dans le
bois. Le prpolymre hydrosoluble a t mlang aux nanoargiles une vitesse de mlange de 3050 t/min pendant 20 minutes pour
former les solutions d'imprgnation qui ont t plus tard imprgnes dans le bois et polymrises in situ. L'influence de la nature des
nanorenforts et des interactions linterface entre les nanorenforts, la MUF et le bois sur les proprits physiques et mcaniques des
nanocomposites en bois rsultants a t tudie en utilisant la spectroscopie lectronique de rayons-X, la microscopie des rayons-X
balayage, limagerie en transmission des rayons X et les mthodes spectroscopiques de micro-sonde (fluorescence) de rayon-X. Des
amliorations significatives des proprits du bois, telles que la duret de surface, le module d'lasticit et la rsistance l'eau, ont
t obtenues avec l'addition des nanorenforts hydrophobes dans le bois. Les proprits en bois amliores ont pu tre attribues aux
proprits inhrentes aussi bien qu de meilleures interactions interfaciales du bois, de la MUF et de nanorenforts. Le traitement au
moulin billes a favoris la dispersion des nanorenforts dans le bois, mais a endommag les groupes fonctionnels sur la surface des
nanorenforts hydrophobes, ce qui tait nuisible pour ladhsion entre le nanorenfort et la matrice de MUF.

Abstract
Wood polymer nanocomposites have been prepared from aspen solid wood samples using water-soluble melamine-urea-formaldehyde
(MUF) prepolymer and three kinds of aluminosilicate nanofillers, one being hydrophilic and the other two hydrophobic. The layered
silicates were ground with a ball mill before being mixed with the MUF resin and impregnated into the wood. The water-soluble
prepolymer was mixed with the nanoclays at a mixing speed of 3050 rpm for 20 minutes to form impregnation solutions that were
subsequently impregnated into the wood and in situ polymerized. The influence of the nature of nanofillers and interphase
interactions between the nanofillers, melamine-urea-formaldehyde and wood on the physical and mechanical properties of the
resulting wood polymer nanocomposites was investigated using X-Ray electron spectroscopy, SEM, TEM and X-ray microprobe
methods. Significant improvements in wood properties, such as surface hardness, modulus of elasticity and water repellence, were
obtained with the addition of hydrophobic nanofillers into the wood. The improved wood properties could be ascribed to inherent
properties as well as better interphase interactions of wood, MUF and nanofillers. Ball-mill treatment favored the dispersion of the
nanofillers into the wood, but broke down functional groups on the hydrophobic nanoclay surface, which was detrimental for the
bonding between the nanofiller and the MUF matrix.

6.1 Introduction
Nanocomposite technology with layered silicate nanoclays as in situ reinforcement has been intensively investigated in recent years
[Qutubuddin and Fu 2002, Ray and Okamoto 2003, Schmidt et al 2002, Lan and Pinnavaia 1994, Wang and Pinnavaia 1998a,b,
Messersmith and Giannelis 1994, Byun et al 2001]. Essential improvements of physical and mechanical properties including tensile
modulus and strength, flexural modulus and strength, thermal stability, flame resistance, and barrier resistance have been observed
for various thermoplastic and thermoset nanocomposites at low silicate content [Qutubuddin and Fu 2002, Ray and Okamoto 2003].

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

Barrier properties, fire resistance and mechanical properties are of great importance for the successful application of selected wood
products. Many efforts have been made in the formation of wood polymer composite (WPC), to improve such properties so as to meet
specific end-use requirements. Both thermoplastic and thermosetting systems have been used and have achieved certain
improvements in wood properties, but both showed limitations [Kumar 1994, Deka and Saikia 2000, Schneider et al 1991, Ayer et al
2003]. Thermoplastic-related monomers such as acrylates or methacrylates, for instance, do not improve the dimensional stability
because these monomers do not penetrate the cell wall or do not react with the wood. Many polymers, even though formed in situ,
only filled the empty lumens in the wood, which leads to a mixture of two materials rather than a true composite. Accordingly, the
resulting products were still subject to dimensional changes with water uptake. Nano-modified WPCs could be a promising new
approach to obtain better products. Few attempts, however, have been made with this regard. The main objective of this project is to
evaluate the feasibility of preparing wood polymer nanocomposites from solid wood.
In our previous work [Cai et al 2007a], we reported on a procedure to prepare wood polymer nanocomposites by impregnation of a
MUF resin and different layered aluminosilicate nanofillers, both hydrophobic and hydrophilic, into solid aspen wood. Dispersion of
nanoparticles into a polymer matrix is crucial to prepare a nanocomposite. Aiming at this, the fillers were further ground with a
ball-mill before being mixed with the MUF resin andimpregnatedinto the wood. The water-soluble MUF prepolymer is hydrophilic
before curing and becomes hydrophobic when cured. Thus, the hydrophilic nanoclays may have good dispersion behavior in the MUF
prepolymer, while the hydrophobic nanofillers may have better interactions with the cured MUF matrix.Experimental results indicated
that significant improvements in wood and MUF/wood composites properties such as surface hardness, modulus of elasticity, water
repellence and dimensional stability were reached, even though only 1% nanofillers were added. It was also observed that use of
different nanofillers resulted in different reinforcements. Ball-milling treatment of the pristine nanofiller (hydrophilic) showed positive
effects on wood properties enhancement, while for the hydrophobic nanoclays, opposite effects were observed. It was suspected that
ball-milling treatment of the hydrophobic nanoclays may destroy some functional groups on the clay surface. This explanation is,
however, not experimentally verified. Then it is assumed that the dispersion of the nanofiller and the prepolymer into the wood and
interphase interactions between wood, MUF and nanofillers, play important roles on the performance of the resulting WPC product. In
this paper, we have investigated the location of MUF resin and nanofillers in the wood. The influence of the nature of nanofillers and
interphase interactions between the nanofillers, MUF and wood on the physical and mechanical properties, and water resistance of
the WPCs was investigated.

6.2 Experiment
6.2.1 Materials
Trembling aspen (Populus tremuloides) logs were obtained locally in in the Qubec City area (Qubec, Canada).The urea,
formaldehyde and melamine used for MUF resin cooking were industrial samples. Low viscosity MUF resin [Cai et al 2007d] was
synthesized in the FPInovations Forintek (Eastern Region). Montmorillonite (MMT) nanoclays, specifically Cloisite 30B (NF1,
organophilic), Claytone APA (NF2, organophilic) and Cloisite Na+ (NF3, a pristine nanoclay, hydrophilic) were obtained from
Southern Clay Products, Inc. (USA). Table 6.1 summarizes some physical properties of the nanoclays used in this study. The average
particle size of the clay powder received is 8 m, and each nanoparticle contains at least 3,000 platelets. Thus the process challenge
of using these nanoclays is not only to disperse the nanoparticles into the MUF resin, but also to exfoliate the platelets to achieve the
maximum surface area and high aspect ratio. Ball-milling treatments of above clays were performed with a ball mill (U.S. Stoneware)
equipped with a 2-literceramic jar, in which 20 porcelain balls (Fisher Scientific) with a diameter of 2.5 cm were utilized. The clays
were put into the ceramic jar and milled at a rolling speed of 162 rpm for the rolls, and 50 rpm for the jar for 48 h. The impregnation
solutions were prepared by adding 1% (wt/wt) nanoclays (both ball-mill treated and untreated) into the low viscosity MUF resin at a
mixing speed of 3050 rpm for 20 min to form suspensions that were subsequently used to impregnate the solid aspen wood samples
under specific conditions. A minimum of 20 specimens was used for different impregnation combinations. The detailed procedures for
the WPCs preparation and mechanical properties measurements were described in the previous work [Cai et al 2007b].

6.2.2 Water Absorption


The untreated and treated samples were submerged in water at 22C for 24 h and 7 days according to the ASTM D-1037 method. The
excess water on the soaked samples was wiped off, and the weight changes of the specimens were measured.

6.2.3 Microscopic Analysis of the WPCs


The morphologies of the treated and untreated wood samples in both radial section and transverse surface were observed using a Jeol
JSM-840A scanning electron microscope (SEM). In the cross section the morphologies of the samples were further investigated using
a Jeol 200 kV transmission electron microscope (TEM). For the SEM observation of the transverse surface, the pure wood, MUF- and
nanofiller/MUF-treated wood blocks were prepared with a razor blade mounted onto a microtome by carefully cutting one of the
end-grain surfaces to a depth of about 3 mm. A gold/palladium alloy was sputtered onto the cutting surfaces prior to investigation.
For radial section morphology observations, samples were immersed in liquid nitrogen and then fractured before being coated with
the gold/palladium alloy and the fracture surfaces were observed. The location of the impregnated MUF resin and nanoparticles in the
wood was identified through X-ray composition analysis of the samples at the same time when the scanning was performed. For TEM
observation, ultra-thin sections of toth treated and untreated samples with a thickness of about 50 nm were microtomed at -80C
using a Leica Ultracut E cryo-ultramicrotome. The staining was carried out using OsO4 in a water/formaldehyde mixture. The
distribution of nanofillers in the wood for the nanofiller/MUF treated wood samples was further investigated using an Electron
Microprobe Analyzer (CAMECA S100, France). The cross-section of the wood was coated with carbon (25 nm) prior to measurement
in order to make it conductive and to protect the surface from beam damage to some extent. Electron microprobe analysis was
performed in a mapping mode with an accelerating voltage of 15 kV and 10 nA. The images of elements Al and Si, which are mainly

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

from aluminosilicate nanofillers for the nanofiller/MUF impregnated wood, were captured digitally to allow for enhanced analysis of
the samples.

6.2.4 Thermogravimetric and Dynamical Mechanical Properties Measurements


The thermogravimetric analysis (TGA) of the specimens was conducted on a TGA/SDTA851e (TA Instrument) instrument under air
atmosphere at a heating rate of 20C min-1 from 30C to 1000C for both the treated and untreated samples. Dynamical mechanical
properties were measured on a DMA V instrument (Rheometric Scientific). A sample was made with a single piece of wood veneer
impregnated with MUF resin or nanofiller/MUF resin with dimension of 3690.6 mm. The solid contents of all resins used were
about 45.9%, and the amount of resin or resin/layered silicate loaded in each sample was about 60-70% of the untreated oven dry
wood weight. The grain direction of wood veneer parallels its longer side and was perpendicular to the loading force. The sample was
impregnated with the MUF resin and nanofiller/MUF resin before the test was performed, respectively. All samples were tested with
dual-cantilever mode at a heating rate of 2C/min and at a strain of 2.010-4 for a frequency of 1 Hz from 30C to 180C for the first
scan. The testing sample was then cooled down to room temperature and scanned again from room temperature to 250C with the
same heating rate and frequency as the first scan. All data reported here were obtained from the second scan of the samples.

6.3 Results and Discussion


6.3.1 Ball-milling Treatment of Nanofillers
Table 6.1 summarizes the physical properties of three montmorillonite nanofillers used in this study. NF1 and NF2 are hydrophobic
nanoclays with surface grafting of quaternary ammonium groups on the clay surfaces, while NF3 is pristine nanoclay (hydrophilic)
without any chemical modification. The pristine nanoclay exhibits the highest specific gravity of 2.86 g/cm3 and the lowest d-spacing
value of the fillers investigated. The inorganic solids content of pristine clay is about 92% after thermogravimetric analysis of the
filler up to a temperature of 1000C. The 8% of weight loss could be due to water adsorption and other contaminants in the filler. For
the hydrophobic nanoclays used, relatively lower inorganic solid contents, around 67% and 61% residue for NF1 and NF2 clays,
respectively, were observed. The higher loss parts of these nanofillers should be the grafted organic components by the nanoclays. As
the average particle size of the received nanoclay in powder form is around 8 m, and each nanoparticle contains at least 3,000
platelets, how to exfoliate the platelets and to disperse these nanofillers into the polymer matrix and the wood is crucial to prepare a
nano-WPC composite. Ball-milling of nanofillers may help in the dispersion of the nanofillers into the wood by reducing their sizes,
thus ball-milling treatment was performed as described in the experimental section. Fig.6.1 shows, as an example, the X-ray
diffraction results of the pristine nanofillers before and after ball-milling treatment. Ball-milling did have some effects on the
nanofillers and increased the d-spacing, that is, the mean distance between platelets, as seen by the much smaller peak, near 8
degrees. Similar results were also observed for other nanofillers. However, when the effect of ball-milling treatments on the nanofiller
size and size distribution was further investigated using a Mastersizer 2000 equipment (Malvem Instruments Ltd.), no significant
difference was found before and after ball-milling treatment of the nanofillers. Nevertheless, the ball milling treatment did have some
effect on the surface of the hydrophobic layered silicate. This will be further discussed later on in this paper.

6.3.2 Enhanced Properties of the WPC Composites


The physical and mechanical properties such as water absorption, hardness, modulus of elasticity (MOE), and modulus of rupture
(MOR) of the prepared WPCs after impregnation of MUF or nanofiller/MUF along with those of the control samples were measured and
summarized in Table 6.2. The weight gains of the oven dried wood polymer composites were also measured and listed in Table 6.2.
From Table 6.2, wood treated with pure MUF resin showed almost the highest weight gain of 72.7%, except for the milled NF2
nanofiller/MUF impregnated wood. For the hydrophobic nanoclay (either ball-mill treated or untreated)/MUF impregnated wood
samples, similar or slightly lower weight gains (around 70%) were observed. For the hydrophilic nanofiller NF3/MUF impregnated
wood samples, much lower weight gains were obtained e.g., 55.3% for the unmilled and 64.9% for the milled nanofiller/MUF
impregnated wood, respectively. Comparing all the nanofiller/MUF treated samples, it seems that ball-milling treatment favors the
dispersion of the nanofillers and MUF into the wood. As one can see from Table 6.2, significantly higher weight gains were seen for all
the ball-mill treated nanofiller impregnated samples as compared with those of corresponding unmilled nanoclay/MUF impregnated
wood.
Wood treatment with either MUF resin alone or the combination of nanofiller and MUF resin displayed significant improvement in
hardness of the wood compared with untreated samples. The highest hardness corresponds to the formulation of nanoclay NF2/MUF
resin mixture, where the hydrophobic nanofillers were used directly without ball-mill treatment. The hardness of NF2/MUF
impregnated wood was improved from 1.09 MPa to 3.25 MPa, an almost 3-fold improvement versus untreated wood. This can be
attributed: first, to the rigid polymeric network structure formed in the wood when the impregnated MUF resin was cured, which
almost doubles the hardness; second, to the nanofiller itself when a nanoclay was combined with the system; and third to the
interaction (adhesion) between the nanoparticles and the cured MUF resin [Cai et al 2007a].
The weight gain of the NF2/MUF impregnated wood is not the highest one as compared to the milled NF2/MUF treated wood. Thus,
the increased hardness is not only due to the increased weight gain of the polymer MUF or the nanoclay/MUF incorporated into the
wood, but also to the nanoclay/MUF resin/wood interfacial interactions. As one can see from Table 6.2, for both ball-mill treated and
untreated NF3/MUF resin impregnated samples, the hardness was lower than that of pure MUF resin impregnated wood. For the
hydrophobic NF1/MUF impregnated samples, significant increases in hardness were observed. Improvements in WPC hardness of 64%
for the NF2/MUF resin formulation and around 30% for the NF1/MUF combination were obtained compared with pure MUF resin
impregnated wood. The hardness of the hydrophobic layered silicate/MUF resin impregnated wood was higher than pure MUF resin
impregnated wood, although only 1% nanoclay, either ball-mill treated or untreated, was used. In general, a polymer does not
intercalate into a layered silicate because the layered silicate is hydrophilic and the polymer is generally hydrophobic. Thus the

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

layered silicate was modified with alkylammonium fatty acids (NF1 and NF2), which confers its hydrophobicity and increases the
d-spacing as Table 6.1 shows, which is a sign of intercalation of MUF between silicate layers. However, some polymers with a
hydrophilic structure such as poly(ethylene oxide) can be easily intercalated in pristine layered silicate without modification with
alkylammonium chain [Choi et al 2000]. The prepolymer MUF is hydrophilic and water soluble before curing, which is helpful in
dispersion of the pristine nanoclays into the prepolymer during the impregnation process. The polarity of the MUF resin is, however,
changed from hydrophilic to hydrophobic during the curing process, which is actually characterized in industry as water tolerance.
This may result in poor interphase interactions of the hydrophilic nanofiller and the cured hydrophobic polymer matrix, causing poor
mechanical properties of the resulting WPC composites, as seen in NF3 based composites.
The difference due to ball-milling treatment in hardness was observed for the hydrophobic nanoclays. For instance, ball-mill
treatment of the NF2 nanofiller resulted in lower hardness than measured for unmilled NF2/MUF resin composites. The reduction in
hardness for the ball-mill treated hydrophobic nanoclay impregnated wood is probably because of ball-mill treatment, which may
destroy some functional groups grafted on the nanofiller surfaces [Cai et al 2007 a,b,c], thus causing poorer interphase interactions
of the nanofiller and the polymer matrix. In order to verify this hypothesis, further work was done through X-ray fluorescence
composition analysis of the nanoclays before and after ball-milling treatment as Figure 6.2 illustrates. Figure 2a,c shows the
elemental compositions detected for the hydrophobic nanoclay NF1 and NF2 without milling, respectively, where the carbon and
oxygen peaks that are mainly from the surface grafted alkylammonium components could clearly be observed. While for the milled
nanoclays (Fig.6.2 b, d), even though a visible oxygen peak could be seen, the carbon in the clay disappears for NF2 and the peak of
carbon weaked for NF1, implying that the alkylammonium functional groups grafted on the nanoclays were removed after ball-milling
treatment for NF2 and the functional groups of NF1 was damaged somehow. This makes sense, since the ball mill treatment works by
fracturing the aluminosilicate particles, which will expose hydrophobic surfaces, as in pristine nanoclays. However, for the nanofiller
NF1, such effect is not obvious. This is probably because of the functional group of NF2 is more compatible with MUF resin than NF1.
Therefore with the ball milling, the damage of the functional group of NF2 showed more significant effect on the mechanical
properties than NF1, the functional groups are damaged at a different level (Fig. 6.2 c,d).
The effect of wood treatment with MUF and different NF/MUF resin formulations on the modulus of elasticity (MOE) and the modulus
of rupture (MOR) of the wood were investigated. It was observed that wood treatment with MUF and NF/MUF resin impregnation
caused drastic improvements in the MOEs of the resulting WPCs as compared with those of untreated samples (Table 6.2). The
average MOE for treated samples is almost double that of correspondingly untreated wood samples. Comparing the effects of different
nanoclay/MUF resin formulations on MOE with that of pure MUF resin showed that there was no significant difference, although
slightly higher MOEs were observed for the formulations containing NF2 (both ball-mill treated and untreated) than for pure MUF
impregnated wood. For the NF3/MUF formulations, lower MOEs were observed for both ball-mill treated and untreated nanoclay/MUF
impregnated wood compared to pure MUF resin impregnation. This could also be attributed to the incompatibility of the cured MUF
resin polymer matrix and the NF3. The highest improvement in MOE was seen with unmilled NF2/MUF. When comparing the effect of
ball-mill treatment of the nanofillers on the MOE of the resulting WPC, no significant difference was found. However, a higher MOE
was observed for the hydrophilic nanofiller/MUF impregnated wood samples when the nanofiller was milled.
Comparing the modulus of rupture (MOR), no significant difference was found between the treated and untreated wood samples
(Table 6.2). Correspondingly, relatively higher MORs were obtained when pure MUF resin and un-milled organophilic nanofiller/MUF
were used for impregnation. For the hydrophilic nanofiller NF3, lower MORs were observed for both the ball-mill treated and
untreated NF3/MUF impregnated wood. It can be seen that ball-mill treatment of nanofillers resulted in slightly lower MORs of
nanoclay/MUF wood composites as compared with control samples (Table 6.2).

6.3.3 Improvement in Water Absorption


The water absorption of the treated and untreated wood samples in water at 21C after 1 day and 7 days was measured and results
are also summarized in Table 6.2. The untreated wood absorbed around 63% of its weight water after a 24 h soak and about 125%
after 1 week immersion. The MUF-impregnated wood absorbed about 8.3% and 38.5% after 24 h and 1 week immersion,
respectively. Thus, MUF treatment significantly reduces water absorption by the wood. For the nanofiller/MUF-impregnated wood,
much lower water absorption was observed, e.g. around 5% in 24 h and 22% in 1 week for the NF2/MUF-impregnated wood. Similar
water absorption behaviour could also be observed for wood composites made with other nanofillers (both organophilic and
hydrophilic: Table 6.2). Several reasons could be ascribed to the decrease in water absorption for the treated wood. First, the cured
hydrophobic polymer MUF shields the wood surfaces and remains in the cell wall and lumen, resulting in less water penetration into
the wood. Secondly, the MUF resin may react with the numerous hydroxyl groups contained in wood components [Gindl and Gupta
2002], thus fewer water absorption sites would remain, which would also contribute to the reduction in water uptake by the treated
wood. For the nanoclay/MUF-impregnated wood, since the nanolayers themselves are impermeable, the impregnated nanolayers
could generate a tortuous pathway for a permeant, e.g. the water molecules, to diffuse through the nanocomposite.

6.3.4 Microscopy investigations


Figure 6.3 shows morphologies of the untreated (Fig. 6.3 (a)) and treated wood samples (Fig. 6.3 (b), (c) and (d)) fractured in liquid
N2 in a typical radial section. For the untreated wood, empty cell wall, the pit and parenchyma can easily be seen, while for the
treated wood, these empty places have been occupied by the MUF or nanofiller/MUF materials, even though some empty columns
(mainly cell lumens and vessels) could still be observed. From Fig. 6.3 (b) (c) and (d), as one can see that the impregnated materials
are mainly located in the cell walls, and good adhesion of the MUF resin to the cell wall could be observed for the treated wood [Cai.
et al2007b]. The trace of nanoclay can be detected with Scanning Electron Microscope as shown in Figs. 6.3 (e) and (f). The size of
the particles ranges from 100 nm to 1 m, which is smaller than the average size of the nanoparticles received 8 m. Further work
has been done through X-ray elemental analysis of the nanoparticles observed in the fracture surface for the nanoclay/MUF treated
wood samples. Elements such as Al and Si, which are mainly from the silicate nanofillers, could be detected, indicating that both MUF

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

and nanofillers had been successfully impregnated into the wood. Figure 6.4 shows the cross section as an example for the
nanofiller/MUF treated wood as well as the X-ray analysis results of the sample. It is clear that the impregnated MUF is mainly located
in the cell wall, and there are many white particles (mainly nanoclay aggregates) like small eggs sitting in the nets of the lumen. This
further proves that the nanoclays did go into the cell lumen with the MUF resin in the prepared wood polymer composite. From the
X-ray micro-probe analysis element distribution images (Fig. 6.4 b), it seems that some nanofillers had penetrated into the cell wall
as well. This could not only result in enhanced mechanical properties of the wood, but contribute to the observed decrease in water
absorption and increase in dimensional stability of the treated wood as well. Fig.6.5 shows the TEM photographs taken from the cell
wall of both the control (pure MUF treated sample) and nanofiller/MUF treated samples. Fig. 6.5a is the image of cell wall of pure MUF
treated wood and Fig. 6.5b is the cell wall of layered silicate NF2/MUF treated wood. The dark lines in Fig. 6.5b have cross sections of
around 10 , corresponding to the thickness of layered silicate. It is clearly visible that the silicate layers are almost homogeneously
dispersed in cell wall with the MUF matrix. This again shows that nanofillers have been impregnated to the cell wall. It should be
noted, however, that the distribution of the nanofillers into the wood is not uniform. As observed in the SEM photographs (Fig. 6.4),
there are some nano aggregates located mainly in the cell lumen or vessels, and some part of the smaller ones in the cell walls.
Further work will be done to improve phase interactions and distribution of the nanofillers into the wood.

6.3.5 Dynamic Mechanical Property Investigations


Dynamic mechanical analysis measures the modulus and damping properties of materials as the materials are deformed under
periodic stress. The dynamic mechanical characteristics of nanocomposite materials are related to the properties of the components,
the morphology of the system, and the nature of the interface between the layered silicate and polymer matrix. When layered
silicates are added to a matrix to form a polymer nanocomposite, an interphase is created. This interphase is very sensitive to the
surface modification of layered silicate, the compatibility between the layered silicate and the matrix. Thus dynamic measurements
can effectively be used to investigate these compatibilities and their impact on nanocomposite properties. Figures 6.6 and 6.7 show,
as examples, the viscoelastic properties of the resulting nanocomposites prepared from the pristine layered silicate (NF3)/MUF, and
an organophilic nanoclay (NF2)/MUF modified wood, along with the MUF resin impregnated wood and the pure wood, respectively.
One can see that the storage modulus (E) of the MUF and nanofiller/MUF modified wood (Fig.6.6) has been improved considerably
over that of the pure wood. This improvement is, however, more obvious for the surface modified layered silicate /MUF resin
impregnated wood than the pristine layered silicate (NF2)/MUF resin modified wood. For the NF3/MUF resin impregnated wood, even
though an enhanced initial storage modulus E was observed as compared to the pure wood, the increase in E values in the whole
temperature rage is much less than the MUF impregnated wood. Given that the poor interphase interactions of the hydrophilic
nanofiller NF3 and the cured hydrophobic polymer matrix, lead to poor mechanical properties of the resulting WPC composites.
The poor interphase interactions of NF3 with the cured MUF matrix was also reflected from the loss tangent, tan, of the material
(Fig.6.7). Broad shoulders, with virtually almost the same peak maximum of about 185 C, were clearly visible for the MUF and
NF3/MUF treated wood. No obvious tan peaks, however, could be detected for the pure wood in the temperature range measured.
This indicates that this transition (Tg, max) arises from the MUF impregnated into the wood, and the hydrophilic nanoclay NF3 had
almost no effect on the Tg, implying poor interactions of the NF3 with the impregnated MUF matrix. From Figure 6.7, it is also
observed that relatively lower tan peak height was seen for the NF3/MUF treated sample than that of pure MUF impregnated wood.
This could be due to relatively lower amount of MUF impregnated into the wood as observed for other NF3/MUF treated samples, on
the other hand, due to the effect of the impregnated nanoparticles. As has been observed in the fiber reinforced polymer composites,
the stiff fiber reduces the tan peak height by restricting the movement of the polymeric chains of the polymer matrix [SalehiMobarakeh et al 1998, Cai et al 2003]. It is interesting to note that for the NF2/MUF impregnated wood, even though almost the
same weight gain as that of the MUF impregnated wood was observed, much lower tan curve height was seen and no tan peaks
could be detected. This implies that strong interphase interactions of the nanofiller and the polymer matrix had been formed. The
highest E values are also obtained for the nanocomposite prepared from the organophilic nanoclay NF2/MUF modified wood. This
could be attributed to the functional groups of the NF2, that are more compatible with the MUF resin, and stronger layered
silicate/MUF resin matrix/wood adhesion could be, thus better performance could be obtained for the resulting WPC composite.

6.4 Conclusions
Wood polymer nanocomposites were prepared from aspen solid wood using water-soluble melamine-urea-formaldehyde (MUF) and
two kinds of nanofillers both hydrophilic and hydrophobic aluminosilicate. Significant improvements in wood properties, such as
surface hardness, modulus of elasticity and water repellency, were obtained with the addition of hydrophobic nanofillers into the
wood. The improved wood properties could be ascribed to better interphase interactions of wood, MUF and nanofillers. Ball-mill
treatment favors the dispersion of the nanofillers into the wood, but somehow breaks down some functional groups on the
hydrophobic nanoclay surface, which is a drawback for the bonding between the nanofiller and the MUF resin. There is strong
evidence that nanoparticles have diffused into the wood cell wall.

Acknowledgements
The authors thank NSERC, the Natural Sciences and Engineering Research Council of Canada, le Fonds qubcois de la recherche sur
la nature et les technologies (FQRNT) and FPInovations Forintek Division (eastern region) for funding of this project.

Table 6.1 Properties of nanoclays.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

1 MT2EtOH:
2R
1
3

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

methyl, tallow, bis-2-hydroxyethyl, quaternary ammonium;

through R4 = combination of aliphatic chains, methyl and (or benzyl groups), N=nitrogen.

X Ray diffraction results d spacing.

4 Thermogravimetric

analysis (TGA) from 30C-1000C, 20C/min, air as perge gas.

Table 6.2 Weight gains for the treated samples, and physical and mechanical properties of the control and the prepared
wood polymer composites

Figure 6.1 X-ray diffraction of the pristine nanofiller before and after the ball-mill treatment.

Figure 6.2 X-ray composition analysis of the hydrophobic nanoclay NF1, NF2 before (a,c) and after ball-milling treatment
(b,d), respectively.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

Figure 6.3 SEM photographs of the fracture surface of (a) untreated, (b) MUF treated, (c) NF3/MUF treated, and (d) NF2/MUF
treated wood. (e) and (f) are trace of nanoclay in the nanoclay/MUF treated wood.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

Figure 6.4 Electron microprobe analysis of aluminum distribution in the NF2/MUF treated wood: (a) transverse image, and
(b) Al distribution.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

Figure 6.5 TEM photographs of (a) MUF and (b) NF2/MUF treated wood cell wall.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

Figure 6.6 Storage modulus E profiles of pure control wood, and MUF, NF2/MUF and NF3/MUF impregnated wood.

Figure 6.7 tan profiles of pure control wood, MUF, NF2/MUF and NF3/MUF impregnated wood

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch06.html

Xiaolin Cai, 2007

Chapter 5 Effects of Nanofillers on Water Resistance

Chapter 7 Montmorillonite Nanoparticles

and Dimensional Stability of Solid Wood Modified by

Distribution and Morphology in Melamine-

Melamine-Urea-Formaldehyde Resin

Urea-Formaldehyde (MUF) Resin Impregnated Wood


Nanocomposites

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch08.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Conclusions and Recommendations

Conclusions and Recommendations


Table des matires
Recommendations for Future Work
Fast growth aspen wood was modified by prepolymer MUF resin in combination of nanoparticles through impregnation process. The
performance of modified aspen wood has been improved, such as surface hardness, modulus of elasticity (MOE), dimensional stability,
water repellence and abrasion resistance. As it has been known, recent developments in nanotechnology, material science and
engineering provide novel and unique methods to the production of advanced materials from natural products. The utilization of
nanotechnology to develop high value added wood products is a very attractive topic for the sustainability of the forest resource. This
project has successfully developed an easy approach to introduce nanotechnology into wood products, which enhance the
performances of the products to benefit the Canadian forest natural resources and wood products industry.
Low viscosity and low cost melamine-urea-formaldehyde was synthesized in the laboratory. One percentage (wt) nanofiller both
hydrophilic and hydrophobic was introduced in the low viscosity MUF resin via high speed mixer. The effect of nanofillers addition on
the curing behaviour and morphology of MUF resin was investigated by DSC and DMTA. Due to the nature of water soluble MUF resin
prepolymer, the hydrophilic nanofiller was compatible during the mixing process, while when the mixed solution was impregnated into
the aspen wood then in situ polymerized at a given temperature, the hydrophilic MUF resin turned to rather hydrophobic after totally
curing.
For better dispersion of the nanoclay into the MUF resin, the received nanoclays were ball-milled before being mixed with MUF resin.
The received nanoclay, and ball-milled nanoclays were analysed by X-ray diffraction. The d-spacing, i.e. spacing between nanosheets,
has been increased after ball milled treatment, and the X-ray density profile of characterized treated wood samples showed that
ball-milled nanoclay have been better dispersed into the wood polymer nanocomposite system. The mechanical/physical properties of
ball milled nanoclay wood polymer nanocomposites didnt improved mostly with the best dispersion as expected. Further investigation
that the ball mill treatment might damage the functional groups on the surface of nanoclay has been made by X-ray composition
analysis. The results of composition analysis showed that some functional peaks disappeared or weakened after the ball mill
treatment for the hydrophobic nanoclay, which was prior to ball milling modified by low molecular organic groups to help the
exfoliation of the nanofiller into different polymer system.
The interphase and morphology of nanoparticles, MUF resin and wood ultra-structure were investigated via different approaches such
as SEM, TEM, AFM and EPMA. TEM images provided evidence that nanoparticles have diffused into the wood cell wall, which explained
the improvement of dimensional stability of wood polymer nanocomposites. EPMA has been used to map the distribution of nanoclay
into the wood polymer nanocomposite via the detection of aluminium composition of layered silicate. It was found that the
distribution of montmorillonite nanoclay looks like a network along the layer of ML (middle lamella), M (compound middle lamella), P
(primary wall), S1 (out layer of secondary wall 1). For the nanoclay, the wood acts like a filter and some part of wood has more
capacity to capture those nanometric particles. An atomic force microscope (AFM) has been used to study the nanometric details of
the nanoclay/MUF wood surface. The adhesion between montmorillonite and MUF resin was observed by the AFM observation. It was
confirmed that the functional groups of the organophilic montmorillonites play an important role on the compatibility between
montmorillonite nanoclay and MUF resin, further to having strong influence on the physical/mechanical properties of the
nanoclay/MUF wood nanocomposites.

Recommendations for Future Work


Further investigation on the fire resistance of the wood polymer nanocomposites products could be investigated.
Different types of nanoclays or nanofillers could be introduced to the system on specific utilization purposes such as surface coating or
anti-UV. The main challenge for the clay series nanofiller is the problem of dispersion, the properties of solid powder has some
limitation and it is easy to form aggregation, especially for the porous materials such as wood products.
Different wood species could be selected to do more investigation on the performance of wood polymer nanocomposites. Both
hardwood and softwood could be utilized for this exploration. It is recommended to use these fast growing low quality and plentyful
resource species for the development of products with high value added and sustainability of nature resource.
Different processes might be used to adapt the nanotechnology to wood products industry, which might be an essential innovation to
benefit the wood product industry in a predictable future. The combination of traditional processes and nanotechnology might be a
quick and efficiency approach to introduce nanoscience into wood products industry to develop high value added products.
Some examples of further use of such nanotechnology would be to introduce iron or nanocarbon nanoparticles in fiber which may
make MDF fiber or paper sensitive to magnetic fields: the fiber could then be oriented. As mentioned before inclusion of inorganic
nanoparticles in wood may make wood much more resistant to insects and fire. Titanium oxide nanoparticles inclusion in exterior
wood might enhance considerably resistance to UV light. The list could be much longer. For the first time, our results show that

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/ch08.html

nanoparticles have the ability to diffuse into the cell wall, which opens endless possibilities for our industry.
Xiaolin Cai, 2007

Chapter 7 Montmorillonite Nanoparticles Distribution

Bibliographie

and Morphology in Melamine-Urea-Formaldehyde


(MUF) Resin Impregnated Wood Nanocomposites

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/apa.html

Collection Mmoires et thses lectroniques


Accueil

propos

Nous joindre

Bibliographie

Bibliographie
Aspen drying procedure Drying aspen and birch Northern hardwoods need a special approach (http://www.valeuraubois.ca
/imports/pdf/en/tech_profiles/TP-03-04W_Dry-Aspen-Birch_English.pdf)
ASTM D 143, AMERICAN SOCIETY FOR TESTING AND MATERIALS. 1994. Standard Methods of Testing Small Clear Specimens of
Timber.
ASTM D 1037 99, AMERICAN SOCIETY FOR TESTING AND MATERIALS. 1999. Standard Test Methods for Evaluating Properties of
Wood-Base Fiber and Particle Panel Materials.
ASTM D 4060-95, AMERICAN SOCIETY FOR TESTING AND MATERIALS. 2001. Abrasion resistance of organic coatings by the taber
abraser.
Ayer, S. W.; Fell, D. and Wan, H. 2003. Hardening of solid wood: Market opportunities and review of existing technologies, Forintek
Canada Corp. Project No.: 3678.
Baki, H.; Yalin rs, M. and Hakki, A. 1993. Improvement of wood properties by impregnation with macromonomeric initiators. J
Appl Polym Sci 47, 1097-1103.
Bailey, P. J. and Preston, R. D. 1969. Some aspects of softwood permeability I. Structural studies with douglas fir sapwood and
heartwood, Holzforschung, 23(4), 113-120.
Beall, F. C.; Meyer, J. A.; Skaar, C. 1966. Direct and RF heat curing of wood-plastic composites, Forest Prod. J., 16(9), 99-106.
Beall, F. C.; Witt, A. E. and Bosco, L. R. 1973. Hardness and hardness modulus of wood-polymer composites, Forest Prod. J., 23(1),
55-60.
Betty W.R. 1976. Dimensional stabilization of lignocellulosic materials, US Patent: 3968318.
Blumstein, A., 1961. tude des polymrisations en couche adsorbe I. Bull. Chim. Soc., 899-905.
Brady, R.L.; Porter, R. S. and Donovan, J.A. 1989. J. Mater. Sci., 24, 4138-4143.
Brebner, K. I.; Schneider, M. H. and St-Pierre L. E. 1985. Flexural strength of polymer-impregnated eastern white pine, Forest Prod.
J., 35(2), 22-27.
Brebner, K. I.; Schneider, M. H. and Jones, R. T. 1988. The influence of moisture content on the flexural strength of styrenepolymerized wood, Forest Prod. J., 38(4), 55-58.
Bryant, B. S. 1966. The chemical modification of wood -from the point of view of wood science and economics, Forest Prod. J., 16(2),
20-27.
Burmester V. A. 1973. Effect of heat-pressure-treatments of semi-dry wood on its dimensional stability. Holz Roh- Werkst. 31,
237-243.
Byun, H. Y.; Choi, M. H. and Chung, I. J. 2001. Synthesis and characterization of resol type phenolic resin/layered silicate
nanocomposite. Chem. Mater. 13, 4221-4226.
Cai, X.; Riedl, B. and Ait-Kadi, A. 2003. Cellulose fiber/poly(ethylene-co-methacrylic acid) composites with ionic interphase.
Composites Part A: applied science and manufacturing, 34, 1075-1084.
Cai, X.; Riedl, B.; Zhang, S.Y. and Wan, H. 2006. Effect of nanofiller on the performance of wood/melamine-urea-formaldehyde
nanocomposites. BioPlastics Montreal 2006, September 27-29.
Cai, X.; Riedl, B.; Zhang, S.Y. and Wan, H. 2007a.Formation and properties of nanocomposites made up from solid aspen wood,
melamine-urea-formaldehyde, and clay. Holzforschung. 61, 148-154.
Cai, X.; Riedl, B.; Zhang, S.Y. and Wan, H. 2007b. Effects of nanofillers on water resistance and dimensional stability of solid wood
modified by melamine-urea-formaldehyde resin. Wood Fiber Sci. 39(2), 307-318.
Cai, X.; Riedl, B.; Zhang, S.Y. and Wan, H. 2007c. The impact of interphase between wood, melamine-urea-formaldehyde and layered
silicate on the performance of wood polymer nanocomposites. Submitted to Composites Part A: applied science and manufacturing.
Cai, X.; Riedl, B.; Zhang, S.Y., Wan, H. and Wang, X.M. 2007d. The effect of layered aluminium silicates on the morphology and
curing behaviors of melamine-urea-formaldehyde resin, unpublished manuscript. Submitted to Polymer.
Chao, W. Y. and Lee, S. W. C. 2003. Properties of southern pine wood impregnated with styrene, Holzforschung, 57(3), 333-336.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/apa.html

Chen, G. C. and Rowell, R. 1986. Approaches to the improvement of biological resistance of wood through controlled-release
technology. In: proceedings of 13th International Symposium on Controlled-Release of Bioactive Materials. Norfolk, VA. Pp. 75-76.
Chen, T. K.; Tien, Y.I. and Wei, K.H. 1999. Synthesis and characterization of novel segmented polyurethane/clay nanocomposite via
poly(-caprolactone)/clay. J Polym Sci. Part A: Polym Chem, 37, 2225-2233.
Choi, M. H.; Chung, I. J. and Lee, J. D. 2000. Morphology and curing behaviours of phenolic resin-layered silicate nanocomposites
prepared by melt intercalation, Chem. Mater. 12, 2977-2983.
Chui, Y. H. Schneider, M. H. and Zhang, H. J. 1994, Effects of resin impregnation and process parameters on some properties of poplar
LVL, Forest Prod. J., 44(6), 74-78.
Deka, M. and Saikia, C. N. 2000. Chemical modification of wood with thermosetting resin: Effect on dimensional stability and strength
property, Bioresource Technology, 73, 179-181.
Dennis, H. R.; Hunter, D. L.; Cho, J. W.; Paul, D. R.; Chang, D.; Kim, S. and White J. L. Nanocomposites by Extruder Processing ,
Southern Clay Products Inc.
Devi, R. R.; Ali, I. and Maji, T. K. 2003. Chemical modification of rubber wood with styrene in combination with a crosslinker: Effect
on dimensional stability and strength property, Bioresource Technology, 88, 185-188.
Dietsche, F.; Thomann, Y.; Thomann, R. and Mulhaupt, R. 2000. Translucent acrylic nanocomposites containing anisotropic laminated
nanoparticles derived from interclated layered silicates, J Appl Polym Sci, 75, 396-405.
Duran, J. A. and Meyer, J. A. 1972. Exothermic heat released during catalytic polymerization of basswood-methyl methacrylate
composites, Wood Sci. and Tech., 6, 59-66.
Dwianto, W.; Inoue M. and M. Norimoto, 1997. Fixation of compressive deformation of wood by heat treatment. Mokuzai Gakkaishi 43
(4), 303-309.
Dwianto, W.; Morooka, T.; Norimoto, M. and Kitajima, T. 1999. Stress relaxation of Sugi (Cryptomeria japonica D. Don) wood in radial
compression under high temperature steam, Holzforschung, 53, 541-546.
Ellis, W. D. and O'Dell, J. L. 1999. Wood-polymer composites made with acrylic monomers, isocyanate, and maleic anhydride, J Appl
Polym Sci, 73, 2493-2505.
Ellwood, E. L.; Gilmore, R. C. and Stamm, A. J. 1972, Dimensional stabilization of wood with vinyl monomers, Wood Science, 4(3),
137-141.
EN1534, 2000. Wood and parquet flooring - Determination of resistance to indentation (Brinell).
Favot, P. 1986. Process for densifying low density woods, US Patent: 4606388.
Feist, W. C.; Rowell, R. M. and Ellis, D. W. 1991. Moisture sorption and accelerated weathering of acetylated and methacrylated
aspen, Wood and Fiber Science, 23(1), 128-136.
Feng, W.; Ait-Kadi, A. and Riedl,B. 2002. Polymerization Compounding: Epoxy-Montmorillonite Nanocomposites, Polymer Engineering
& Science, 42(9): 1827-1835.
Fengel, D. and Wegener, G. 1984. Wood - Chemistry, Ultrastructure, Reactions. De Gruyter, Berlin.
Fry H. J. 1976. Compression impregnation of wood veneers, US Patent: 3950577.
Furono, T. and Goto, T. 1973. The penetration of MMA monomer into the woody cell wall. Mokuzai Gakkaishi, 19 (6): 271-274.
Furuno, T.; Goto, T. 1978. Structure of the interface between wood and synthetic polymer. XI. The role of polymer in the cell wall on
the dimensional stability of wood-polymer composite (WPC). Mokuzai Gakkaishi, 24(5), 287-293.
Furuno, T.; Uehara, T. and Jodai, S. 1991. Combinations of wood and silicate I. Impregnation by water glass and applications of
aluminum sulfate and calcium chloride as reactants, Mokuzai Gakkaishi, 37(5), 462-472.
Furuno, T.; Shimada, K.; Uehara, T. and Jodai, S. 1992a. Combinations of wood and silicate II. Wood-mineral composites using water
glass and reactants of barium chloride, boric acid, and borax, and their properties, Mokuzai Gakkaishi 38(5), 448-457.
Furuno, T.; Uehara, T.; Jodai, S. 1992b. The role of wall polymer in the decay durabilities of wood-polymer composites, Makuzai
Gakkaishi, 38(3), 285-293.
Furuno, T. Uehara, T. and Jodai, S. 1993. Combinations of wood and silicate III. Some properties of wood-mineral composites using
the water glass-boron compound system, Makuzai Gakkaishi, 39(5), 561-570.
Garcs, J. M.; Moll, D. J.; Bicerano, J. Fibiger, R. and McLeod, D. G. 2000. Polymeric nanocomposites for automotive applications,
Adv. Mater. 12(23), 1835-1839.
Galperin, A. S.; Kuleshov, G. G.; Tarashkevich, V. I. and Shutov, G. M. 1995. Manufacturing and properties of modified wood: A
review of 25 years work. Holzforschung, 49(1), 45-50.
Gaylord, N. G. 1973, Process for impregnating porous, cellulosic material by in situ polymerization of styrene-maleic anhydride

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/apa.html

complex. U.S. Patent No. 3765934.


Giebeler, E. 1983 Dimensional stabilization of wood by moisture-heat-pressure-treatment. Holz Roh- Werkst. 41, 87-94.
Gilman, J. W. 1999. Flammability and thermal stability studies of polymer layered-silicate (clay) nanocomposites, Applied Clay
Science. 15, 31-49.
Gindl, W. and Gupta, H. S. 2002. Cell-wall hardness and Young's modulus of melamine-modified spruce wood by nano-indentation,
Composites Part A: applied science and manufacture, 33, 1141-1145.
Gindl, W.; Dessipri, E. and Wimmer, R. 2002. Using UV-microscopy to study diffusion of melamine-urea-formaldehyde resin in cell
walls of spruce wood, Holzforschung, 56, 103-107.
Gindl, W.; Muller, U. and Teischinger, A. 2003a. Transverse compression strength and fracture of spruce wood modified by melamineformaldehyde impregnation of cell walls, Wood and Fiber Science. 35(2), 239-246.
Gindl, W.; Zargar-Yaghubi, F. and Wimmer, R. 2003b. Impregnation of softwood cell walls with melamine-formaldehyde resin,
Bioresource Technology, 87, 325-330.
Guevara, R. and Moslemi, A. A. 1984. The effect of alkylene oxides, furan resin and vinylpyrrolidinone on wood dimensional stability,
Wood Sci. Technol. 18, 225-240.
Hagstrand, P. O.; Klason, C.; Svensson, L. and Lundmark, S. 1999. Rheokinetical behavior of melamine-formaldehyde resins, Polym.
Eng. Sci. 31(10), 2019-2029.
Hartley, I. D. and Schneider, M. H. 1993. Water vapour diffusion and adsorption characteristics of sugar maple (Acer saccharum,
Marsh.) wood polymer composites. Wood Sci. Technol. 27, 421-427.
Hartman, S. 1969. Modified wood with aqueous polyurethane systems, Forest Prod. J., 19(5), 39-42.
He, G. B. and Riedl, B. 2003 Phenol-urea-formaldehyde cocondensed resol resins: Their synthesis, curing kinetics, and network
properties. J. Polym. Sci. Part B: Polym. Phys. 41, 1929-1938.
Hoffmann, P. 1988. On the stabilization of waterlogged oakwood with polyethylene glycol (PEG), Holzforschung, 42(5), 289-294.
Hoffmann, P. 1990. On the stabiliztion of waterlogged softwoods with polyethylene glycol(PEG). four species from China and Korea,
Holzforschung, 44(2), 87-93.
Hua, L.; Zadorecki, P. and Flodin, P. 1987a. Cellulose fiber-polyester composites with reduced water sensitivity (1) Chemical
treatment and mechanical properties. Polym. Comp. 8, 199-202.
Hua L, Flodin P and Ronnhult T, 1987b. Cellulose fiber-polyester composites with reduced water sensitivity (2) surface analysis.
Polym. Comp. 8, 203-207.
Huang, J.; Wang, Q.; Kong, X. and Zhang, Y. 1997. Study on the mechanical and tribological properties of glassfiber reinforced
polyethylene composites. Acta Materiae Compositae Sinica. 14(4), 19-25.
Inoue, M.; Norimoto, M. Otsuka, Y. and Yamada, T. 1991a. Surface compression of coniferous wood lumber. I. Permanent set of
surface compressed layer by a resin. Mokuzai Gakkaishi, 35(3): 234-240.
Inoue, M.; Norimoto, M.; Otsuka, Y and Yamada, T. 1991b. Surface compression of coniferous wood lumber. II. Permanent set of
compression wood by low molecular weight phenolic resin and some physical properties of the products. Mokuzai Gakkaishi, 35(3):
227-233.
Inoue, M.; Norimoto, M.; Tanahashi, M. and Rowell, R. M. 1993a. Steam or heat fixation of compressed wood, Wood and Fiber Sci.,
25(3), 224-235.
Inoue, M.; Ogata, S.; Kawai, S.; Rowell, R. M. and Norimoto, M. 1993b. Fixation of compressed wood using melamine-formaldehyde
resin, Wood and Fiber Sci. 25(4), 404-410.
Inoue, M.; Ogata, S.; Nishikawa, M.; Otsuka, Y.; Kawai, S. and Norimoto, M. 1993c. Dimensional stability, mechanical properties, and
color changes of a low molecular weight melamine-formaldehyde resin impregnated wood. Mokuzai Gakkaishi, 39, 181-189.
Inoue, M.; Minato, K. and Norimoto, M. 1994. Permanent fixation of compressive deformation of wood by crosslinking. Mokuzai
Gakkaishi 40(9), 931-936.
Ito, Y.; Tanahashi, M.; Shigematsu, M.; Shinoda, Y. and Ohta, C. 1998a. Compressive-molding of wood by high-pressure steamtreatment: part 1. Development of compressively molded squares from thinnings, Holzforschung, 52, 211-216.
Ito Y.; Tanahashi, M.; Shigematsu, M. and Shinoda, Y. 1998b. Compressive-molding of wood by high-pressure steam-treatment: part
2. Mechanism of permanent fixation, Holzforschung, 52, 217-221.
Juneja, S. C. and Hodgins, J. W. 1970. The properties of thermo-catalytically prepared wood-polymer composites, Forest Prod. J.,
20(12), 24-28.
Khan, M. A.; Idriss Ali, K. M. and Ahmad, M. U. 1992a. Radiation-induced wood plastic composites under combinations of monomers, J
Appl Polym Sci. 45, 2113-2119.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/apa.html

Khan, M. A. and Idriss Ali, K. M. 1992b. Effect of urea on the mechanical strength of wood-plastic composites, Radiat. Phys. Chem.
40(1), 69-70.
Khan, M. A. and Idriss Ali, K. M. 1994. Dose effect in the wood plastic composite formation in the presence of additives and
co-additives, Radiation Physics and Chemistry. 44(4), 415-420.
Kim, M. G. 2001. Examination of selected synthesis parameters for wood adhesive-type urea-formaldehyde resins by

13 C

NMR

spectroscopy. III, J Appl. Polym. Sci. 80, 2800-2814.


Kim, M. G.; Nieh, W. L. S. 1991. Study on the curing of phenol-formaldehyde resol resins by dynamic mechanical analysis. Ind. Eng.
Chem. Res. 30, 798-803.
Kininmonth, J. A. 1972. Permeability and fine structure of certain hardwoods and effects on drying II. Differences in fine structure of
Northofagus fusca sapwood and hard wood. Holzforschung, 26, 32-38.
Kissinger, H. E. 1957. Reaction kinetics in differential thermal analysis. Anal. Chem., 29 (11), 1702-1706.
Kojima, Y.; Usuki, A.; Kawasumi, M.; Okada, A.; Kurauchi, T. and Kamigaito, O. 1993a. Sorption of water in nylon 6-clay hybrid.J.
Appl. Polym Sci., 49, 1259-1264.
Kojima, Y.; Usuki, A.; Kawasumi, M.; Okada, A.; Fukushima, Y.; Kurauchi, T. and Kamigaito, O., 1993b. Mechanical properties of nylon
6-clay hybrid. J. Mater. Res. 8, p. 1185-1189.
Kornmann, X.; Lindberg, H. and Berglund, L.A. 2001. Synthesis of epoxy-clay nanocomposites. Influence of the nature of the curing
agent on structure, Polymer, 42. 4493-4499.
Kumar, S. 1994. Chemical modification of Wood, Wood and Fiber Science, 26(2), 270-280.
Kurauchi, T.; Okada, A.; Nomura, T.; Nishio, T.; Saegusa, S. and Deguchi, R. 1991. SAE Technical Paper Ser., 910584.
Lan, T. and Pinnavaia, T. J. 1994. Mechanism of clay tactoid exfoliation in epoxy-clay nanocomposites. Chem. Mater. 6, 2216-2219.
Lan, T.; Kaviratna, P. D. and Pinnavaia, T. J. 1995. Mechanism of clay tactoid exfoliation in epoxy-clay nanocomposites, Chem. Mater.,
7, 2144-2150.
Langwig, J. E.; Meyer, J. A. and Davidson, R. W. 1968. Influence of polymer impregnation on mechanical properties of basswood,
Forest Prod. J., 18(7), 33-36.
Langwig, J. E.; Meyer, J. A. and Davidson, R. W. 1969. New monomers used in making wood-plastics, Forest Prod. J., 19(11), 57-61.
Lee, D. C. and Jang L. W. 1996. Preparation and characterization of PMMA-clay hybrid composite by emulsion polymerization, J. Appl.
Polym. Sci. 61, 1117-1122.
Li, J-Z; Furuno, T. and Katoh, S. 2001. Preparation and properties of acetylated and propionylated wood-silicate composites,
Holzforchung, 55(1), 93-96.
Loo, L. S. and Gleason, K. K. 2003. Fourier transforms infrared investigation of the deformation behavior of montmorillonite in nylon
6/nanoclay nanocomposite. Macromolecules, 36: 2587-90.
Loos, W. E.; Walters, R. E. and Kent, J. A. 1967, Impregnation of wood with vinyl monomers, Forest Prod. J., 17(5), 40-49.
Loos, W. E. and Robinson, G. L. 1968. Rates of swelling of wood in vinyl monomers, Forest Prod. J., 18(9), 109-112.
Mathias,

L.

J.

and Wright,

J.

R.

1989.

New wood-polymer

composites:

Impregnation

and

in

situ

polymerizaiton

of

hydroxymethylacrylates, Polymer Reprints, 30(1), 233-234.


Messersmith, P.B. and Giannelis, E. P. 1994. Synthesis and characterization of layered silicate epoxy nanocomposites. Chem.
Mater. 6, 1719-1725.
Meyer, J. A. 1965. Treatment of wood-polymer systems using catalyst-heat techniques, Forest Prod. J., 15(9), 362-364.
Meyer, J. A. 1981. Wood-polymer materials: state of the art, Wood Science, 14(2), 49-54.
Meyer, J. A. 1982. Industrial use of wood-polymer materials: state of the art, Forest Prod. J. 32(1), 24-29.
Miroy, F.; Eymard, P. and Pizzi, A. 1995. Wood hardening by methoxymethyl melamine. Holz Roh Werkst.53, 276-276.
Murmanis, L. and Chudnoff, M. 1979. Lateral flow in beech and birch as revealed by the electron microscope. Wood Sci. Technol. 13,
79-87.
Nobutaka, N. and Yukio, S. 1977. Preparation of densified wood impregnated with phenolic resins, US Patent: 4031276.
Nascimento, G. M.; Constantino, V.R.L. and Temperini, M. L. A. 2002. Spectroscopic characterization of a new type of conducting
polymer-clay nanocomposite. Macromolecules, 35:7535-7537.
Noah, J. N. and Foudjet, A. 1988. Wood-polymer composites from some tropical hardwoods, Wood Sci. Technol. 22, 115-119.
Norimoto, M. and Gril, J. 1989. Wood bending using microwave heating. J. Microwave Power and Electromagnetic Energy. 24(4),
203-212.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/apa.html

Ogata, N.; Kawakage, S. and Ogihara, T. 1997. Structure and thermal/mechanical properties of poly(ethylene oxide)-clay mineral
blends. Polymer, 38, 5115-5118.
Ogiso, K. and Saka, S. 1993. Wood-inorganic composites prepared by sol-gel process II. Effects of ultrasonic treatments on
preparation of wood-inorganic composites, Mokuzai Gakkaishi, 39(3), 301-307.
Paavo, M. and Markku P. 1994. Method for the hardening of wood material. EP0626240.
Parfitt, G. D. (ed.), 1981. Dispersion of Powder in Liquids, Applied Science Publisher, NJ.
Pizzi, A. 1994. Advanced wood adhesives technology, Pizzi ed., Marcel Dekker, 67-88.
Pizzi, A. and Panamgama, L. A. 1995. Diffusion hindrance vs. wood-induced catalytic activation of MUF adhesive polycondensation. J.
Appl. Polym. Sci. 55: 109-115.
Porter, R.S. and Casale, A. 2004. Recent studies of polymer reaction caused by stress. Polym. Eng. Sci. 25(3), 129-156.
Qutubuddin, S. and Fu, X. 2002. Polymer-Clay nanocomposites: Synthesis and Properties, Nano-Surface Chemistry, ed. Morton
Rosoff, Marcel Dekker, Inc. New York, 653-673.
Raff, R. A. V.; Herrick, I. W. and Adams M. F. 1965. Polymerization of styrene and styrene-divinylbenzene in wood, Forest Prod. J.,
15(7), 260-262.
Ramalingam, K. V.; Werezak, G. M. and Hodgins, J. W. 1963. Radiation-induced graft polymerization of styrene in wood. J. Polym. Sci.
Part C. 2, 153-167.
Rapp, A. O.; Bestgen, H.; Adam, W. and Peek, R-D. 1999. Electron energy loss spectroscopy (EELS) for quantification of cell-wall
penetration of a melamine resin, Holzforschung, 53(2), 111-117.
Ray, S. S. and Okamoto, M. 2003. Polymer/layered silicate nanocomposites: a review from preparation to processing, Prog. Polym.
Sci. 28, 1539-1641.
Rowell, R. M. 1975, Chemical modification of wood, advantages and disadvantages. Am. Wood Preservers Assoc. 71: 41-51.
Rowell, R. M.; Gutzmer, D. I.; Sachs, I. B. and Kinney, R. E. 1976. Wood Science, 9(1), 51-54.
Rowell, R. M. and Ellis, W. D. 1978, Determination of dimensional stabilization of wood using water soak method. Wood Fiber, 10(2),
104-111.
Rowell, R. M.; Moisuk, R.; Meyer, J. A. 1982. Wood-polymer composites: Cell wall grafting with alkylene oxides and lumen treatments
with methyl methacrylate, Wood Science, 15 (2), 90-96.
Rowell, R. M. 1983. Chemical modification of wood. For. Prod. Abstra., 6 (12), 363-382.
Rowell, R. M. 1983b, Controlled release delivery stystems. In: Bioactive polymer-wood composites. Ch. 23. Marcel Dekker Inc., New
York. Pp. 347-357.
Roussel, C.; Marchetti, V.; Lemor, A.; Wozniak, E.; Loubin, B. and Gerardin, P. 2001. Chemical modification of Wood by
polyglycerol/maleic anhydride treatment, Holzforschung, 55(1), 57-62.
Ruan, W. H.; Zhang, M. Q.; Rong, M. Z. and Friedrich, K. 2004. Polypropylene composites filled with in-situ grafting polymerization
modified nano-silica particles. J. Mater. Sci. 39(10), 3475-3478.
Rudman, P. 1966a. Studies in wood preservation, Pt. II*) Movement of aqueous solutions through the pits and cell walls of eucalypt
sapwoods, Holzforschung, 20(2), 57-60.
Rudman, P. 1966b. Studies in wood preservation, Pt. III*) The penetration of the fine structure of wood by inorganic solutions,
including wood preservatives, Holzforschung, 20(2), 60-67.
Saka, S.; Sasaki, M. and Tanahashi, M. 1992. Wood-inorganic composites prepared by sol-gel processing I. Wood-inorganic
composites with porous structure, Mokuzai Gakkaishi, 38(11), 1043-1049.
Saka, S. and Yakake, Y. 1993. Wood-inorganic composites prepared by sol-gel process III. Chemically-modified wood-inorganic
composites, Mokuzai Gakkaishi, 39(3), 308-314.
Salehi-Mobarakeh, H., Brisson, J. and Ait-Kadi, A. 1998. Ionic interphase of glass fiber/polyamide 6,6 composites.Polym. Compos. 19,
264-274.
Schaffer, E. 1971. Studies in Conservation. 16, 110-113.
Schadler L. S. 2003. Polymer-based and polymer-filled nanocomposites, In: Nanocomposite Science and Technology. Edited by Ajayan
P. M., Schadler L.S., Braun P. V. WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, pp.77-154.
Schmidt, D.; Shah, D. and Giannelis, E. P. 2002. New advances in polymer/layered silicate nanocomposites. Current Opinion in Solid
State & Mater. Sci. 6, 205-212.
Schneider, M. H. and Brebner, K. I. 1985. Wood-polymer combinations: the chemical modificaiton of wood by alkoxysilane coupling
agents, Wood Sci. Technol., 19, 67-73.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/apa.html

Schneider, M. H.; Brebner K. I. and Hartley I. D. 1991. Swelling of a cell lumen filled and a cell-wall bulked wood polymer composite
in water. Wood Fiber Sci., 23 (2), 165-172.
Schneider, M. H. 1994. Wood polymer composites, Wood and Fiber Science, 26(1), 142-151.
Schneider, M. H. 1995, New cell wall and cell lumen wood polymer composites, Wood Sci. Technol. 29, 121-127.
Schneider, M. H. 2000. Wood-polymer composites. The Encyclopedia of Materials: Science and Technology, Elsevier Science,
9764-9766.
Shi, H.; Lan, T. and Pinnavaia, T. J. 1996. Interfacial Effects on the Reinforcement Properties of Polymer-Organoclay Nanocomposites.
Chem Mater, 8, 1584-1587.
Siau, J. F.; Meyer, J. A.; Skaar, C. 1965. Wood-polymer combinations using radiation techniques, Forest Prod. J. 15(10), 426-434.
Siau, J. F. and Meyer J. A. 1966. Comparison of the properties of heat and radiation cured wood-polymer combinations, Forest Prod. J.
16(8), 47-56.
Siau, J. F.; Davidson, R. W.; Meyer, J. A. and Skaar, C. 1968. A geometrical model for wood-polymer composites, Wood Science, 1(2),
116-128.
Siau, J. F. 1969. The swelling of basswood by vinyl monomers, Wood Science, 1(4), 250-253.
Siau, J. F.; Smith. W. B. and Meyer, J. A. 1978. Wood-polymer composites from southern hardwoods, Wood Science, 10(3), 158-164.
Singh, A.; Dawson, B.; Franich, R.; Cowan, F. and Warnes, J. 1999. The relationship between pit membrane ultrastructure and
chemical impregnability of wood, Holzforschung, 53(4), 341-346.
Soulounganga, P.; Marion, C.; Huber, F. and Gerardin, P. 2003. Synthesis of poly(glycerol methacrylate) and its application to
dimensional stabilization of wood. J Appl Polym Sci. 88, 743-749.
Stamm, A. J. 1959. Effect of polyethylene glycol on the dimensional stability of wood. Forest Prod J. 9, 375-381.
Stamm, A. J. 1964. Factors affecting the bulking and dimensional stabilization of wood with polyethylene glycols. Forest Prod J. 14,
403-408.
Su, S. and Wilkie, C. A. 2003. Exfoliated poly(methyl methacrylate) and polystyrene nanocomposites occur when the clay contains a
vinyl monomer, J. Polym. Sci.: Part A: Polymer Chemistry. 41, 1124-1135.
Tanahashi, M.; Goto, T.; Hori, F.; Hirai, A. and Higuchi, T. 1989. Characterization of steam-exploded wood III: Transformation of
cellulose crystals and changes of crystallinity. Mokuzai Gakkaishi 35: 654-662.
Tanaka, Y.; Zhang, Q. and Saito, F. 2004. Mechanochemical dechlorination of chlorinated compounds. J. Mater. Sci. 39(16-17),
5497-5501.
Timmons, T. K.; Meyer, J. A.; Jr Cote, W. A. 1971. Polymer location in the wood-polymer composite, Wood Science, 4(1), 13-24.
Tsou, C. T. and Shiah, T. C. 2003. Reaction kinetics behavior of melamine-urea-formaldehyde wood adhesives by differential thermal
analysis and its application. Taiwan J. For. Sci. 18 (4), 339-48.
Usuki, A.; Kawasumi, M.; Kojima, Y.; Fujushima, A.; Okada, A. and Kamigaito, O. 1993a.Swelling behaviour of montmorillonite cation
exchanged for v-amino acids by -caprolactam. J. Mater. Re., 8, 1174-1178.
Usuki, A.; Kojima, Y.; Kawasumi, M.; Okada, A.; Fujushima, A.; Kurauchi, T. and Kamigaito, O. 1993b. Synthesis of nylon 6-clay
hybrid, J. Mater. Res., 8, 1179-1184.
Vaia, R.; Jandt, K.; Kramer, E. and Giannelis, E. 1995. Kinetics of polymer melt intercalation, Macromolecules, 28, 8080-8085.
Vaia, R.; Jandt, K.; Kramer, E. and Giannelis, E. 1996. Microstructural Evolution of Melt Intercalated Polymer-Organically Modified
Layered Silicates Nanocomposites. Chem. Mater. 8, 2628-2635.
VanderHart, D. L.; Asano, A. and Gilman, J.W. 2001. NMR measurements related to clay dispersion quality and organic-modifier
stability in nylon 6/clay nanocomposites. Macromolecules, 34, 3819-22.
Wang, M. S. and Pinnavaia, T. J. 1994. Clay-polymer nanocomposites formed from acidic derivatives of montmorillonite and an epoxy
resin. Chem. Mater. 6. 468-474.
Wang, Z. and Pinnavaia, T. J. 1998a. Hybrid organic-inorganic nanocomposites: Exfoliation of magadiite nanolayers in an elastomeric
epoxy polymer. Chem. Mater. 10, 1820-1826.
Wang, Z. and Pinnavaia, T. J. 1998b. Nanolayer reinforcement of elastomeric polyurethane. Chem. Mater. 10, 3769-3771.
Wellons, J. D.; Krahmer, R. L.; Sandoe, M. D. and Jokerst, R. W. 1983: Thickness loss in hotpressed plywood. Forest Prod. J. 33(1):
27-33.
Wright, J. R. and Mathias, L. J. 1993. New lightweight materials: Balsa wood-polymer composites based on ethyl a-(hydroxymethyl)
acrylate, J Appl. Polym. Sci. 48, 2241-2247.

08-Aug-15 00:15

Wood modifications for valued-added applications using nanotechnolog...

http://theses.ulaval.ca/archimede/fichiers/24888/apa.html

Wu, H. D.; Tseng, C.R. and Chang, F. C. 2001. Chain conformation and crystallization behavior of the syndiotactic polystyrene
nanocomposites studied using Fourier transform infrared analysis. Macromolecules, 34, 2992-2999.
Yalinkilic, M. K.; Tsunoda, K.; Takahashi, M.; Gezer E. D.; Dwianto, W. and Nemoto, H. 1998. Enhancement of biological and physical
properties of wood by boric acid-vinyl monomer combination treatment, Holzforschung, 52(6), 667-672.
Yalinkilic, M. K.; Imamura, Y.; Takahashi, M.; Demirci, Z. and Yalinkilic, A. C. 1999. Biological, mechanical, and thermal properties of
compressed-wood polymer composite (CWPC) pretreated with boric acid, Wood and Fiber Sci., 31(2), 151-163.
Yamaguchi, H. 1994a. Properties of silicic acid compounds as chemical agents for impregnation and fixation of wood, Mokuzai
Gakkaishi, 40(8), 830-837.
Yamaguchi, H. 1994b. Preparation and physical properties of wood fixed with silicic acid compounds, Mokuzai Gakkaishi, 40(8),
838-845.
Young, R. A. and Meyer, J. A. 1968, Heartwood and sapwood impregnations with vinyl monomers. Forest Prod. J., 18(4), 66-68.
Zerda, A.; Caskey, T. C. and Lesser, A. J. 2003. Highly concentrated, intercalated silicate nanocomposites: Synthesis and
characterization, Macromolecules. 36, 1603-1608.
Zhang, H. J.; Chui, Y. H. and Schneider, M. H. 1994. Compression control and its significance in the manufactue and effects on
properties of poplar LVL, Wood Sci. Technol. 28, 285-290.
Zollfrank, C. and Wegener, G. 2002. FTIR microscopy and ultrastructural investigation of silylated solid wood, Holzforschung, 56(1),
39-42.
Xiaolin Cai, 2007

Conclusions and Recommendations

APENDIX I Scanning Electron Microscope


Observation Imagines of Pure Aspen Wood, MUF
Treated Aspen Wood, and Montmorillonite
Clay/MUF/Wood Nanocomposites.

08-Aug-15 00:15

You might also like