. Lilvufir l 1.. .. ,Jl..d:r. Ir .a v ..0 \ .3115...” 'SJ.:I .. .n 9 ’ z: t . u . gmvwfl 1 {l- ., :19 '< I. anxid‘ vllf a. .. ”.4 I), .5.‘ 41’. 1: sun.“ ... .1 s. :1" ft . ll l.v .2”: it . .u; .v i. 2min La AMJWWT.‘ 2.2; I~ . .4: .9); {A t. «I b... 6‘ . 1": [ THEE}? 7000 This is to certify that the dissertation entitled COMPOSITES FROM MALEATED POLYOLEFlN-GRAFTED WOOD PARTICLES PRODUCED VIA REACTIVE EXTRUSION presented by KARANA CARLBORN has been accepted towards fulfillment of the requirements for the Doctoral degree in Forestry (/[Ltet‘re bug; in aiua ’vz Ck Major Piofessor’s Signature 0313 l 1 4100 g; Date MSU is an Ailirmative Action/Equal Opportunity Institution LIBRARTD “ " ‘‘‘‘ Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE B 07 C 2 CM); 01‘ 1 1 I C C. 2/05 p:/ClRC/DateDue.indd-p.1 COMPOSITES FROM MALEATED POLYOLEFIN-GRAFTED WOOD PARTICLES PRODUCED VIA REACTIVE EXTRUSION By Karana Carlbom A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Forestry 2006 ABSTRACT COMPOSITES FROM MALEATED POLYOLEFIN GRAF TED WOOD PARTICLES PRODUCED VIA REACTIVE EXTRUSION By Karana Carlbom This study examined the concept of using a reactive extrusion process to develop a new, formaldehyde-free binding system for wood composite products. The surfaces of wood particles were modified by grafting maleated polyethylene (MAPE) and maleated polypropylene (MAPP) compounds through a continuous reactive extrusion process. MAPE content was varied to study the effect of material composition on grafting efficiency, while extruder barrel temperatures and rotational screw speeds were varied to evaluate the effects of processing conditions on the modification of wood particles. Polymer molecular weight effects were followed using MAPP with different molecular weights. Efficiency of the modification was assessed using FTIR, 13C NMR and XPS surface analysis techniques, along with a titrimetric analysis to verify the esterification reaction between the wood particles and maleated polyolefins. Composite panels were made from wood particles modified with MAPE and MAPP binding agents under two different manufacturing methods. Specific contrasts of (i) base resin type, PE vs. PP, (ii) molecular weight/maleic anhydride content in MAPP binding agents, and (iii) the manufacturing methods (reactive extrusion vs. hot press) were investigated to determine the effects of these factors on the physico-mechanical properties of the composites. Finally, a response surface method using a Box-Behnken design was constructed to statistically model and optimize the material compositions-processing conditions- mechanical property relationships of formaldehyde-free wood composite panels. FTIR, 13C NMR, XPS and titration data confirmed the grafting of maleated polyolefins onto the surface of wood particles through an esterification reaction, while the level of grafting of MAPE onto wood particles was determined to be a function of the MAPE concentration. However, there was no significant difference found in grafting efficiency at different extrusion processing conditions; rather all of the conditions resulted in adequate grafting. Similarly, there was no difference in grafting efficiency with the molecular weight of MAPP. Reactive extrusion was found to be a suitable technique for the modification of wood particles with maleated polyolefins for all of the material compositions and processing conditions studied. Mechanical property test results indicated that most composite panels met or even exceeded the standard requirements for particleboard of medium density. While extruding the particles before panel pressing gave better internal bond (13) strength, superior bending properties were obtained through compression molding alone. MAPP-based panels outperformed MAPE-based panels in stiffness. Conversely, MAPE increased the 18 strength of the panels compared to MAPP. Relationships between material compositions, processing conditions and both flexural strength (MOR) and 18 strength of the panels were described by linear models. Increasing any of the manufacturing variables resulted in greater MOR and IB strength. Flexural stiffness (MOE) was described by a quadratic regression model. Increased MOE was obtained through higher pressing times, binding agent concentrations and/or pressing temperatures, although binding agent concentration had less effect on MOE at higher pressing temperatures. Numerical optimization showed that panels with desired mechanical properties could be made under a range of manufacturing conditions. ACKNOWLEDGEMENTS I would first like to thank my major professor, Dr. Laurent Matuana, for his guidance through my PhD program. He taught, me the difference between working hard and working smart, and gave me the opportunity to develop both my research and scientific writing skills. I am certain these skills will serve me well in my career. I appreciate my committee members Dr. Koelling, Dr. Selke and Dr. Bix for their helpful suggestions and advice. They spent a lot of time reading my dissertation and provided good comments in the editing process. My work was supported by the USDA—CSREES Grant—Advanced Technology Applications to Eastern Hardwood Utilization, and the McIntire-Stennis Cooperative Forestry Research Program. I am grateful to American Wood Fibers and Eastman Chemical Company for generously donating the materials that were used in this work. Many thanks go to my co-workers in the wood composites group at MSU. I have been very lucky to work with such smart people and to benefit from their knowledge and experience. I was also fortunate to work with numerous undergraduate lab assistants over the course of this project. Finally, I want to thank my family and friends who helped and supported me throughout this process. I appreciate my family for always encouraging me during my graduate studies. My friends helped me through tough days and celebrated small victories along the way. In particular, my friends here at MSU have been invaluable in getting through my Ph.D. program with a smile on my face. iv TABLE OF CONTENTS LIST OF TABLES ........................................................................................................... viii LIST OF FIGURES ........................................................................................................... ix LIST OF ABBREVIATIONS .......................................................................................... xiii CHAPTER 1 INTRODUCTION .............................................................................................................. 1 Objectives ............................................................................................................... 6 References ............................................................................................................... 7 CHAPTER 2 BACKGROUND AND LITERATURE REVIEW ........................................................... 10 Chemical Composition of Wood .......................................................................... 10 Cellulose ................................................................................................... lO Hemicellulose ........................................................................................... 1 1 Lignin ........................................................................................................ 11 Extractives ................................................................................................. 13 Wood Anatomical Structure ................................................................................. l3 Moisture Content .................................................................................................. 17 Wood Composites ................................................................................................. 17 Medium Density Fiberboard (MDF) ......................................................... 17 Particleboard ............................................................................................. 18 Adhesives for Wood Composites .............................................................. 18 Phenolic Resins ............................................................................. 19 Amino Resins ................................................................................ 21 Wood-Plastic Composites ..................................................................................... 25 Maleated Polyolefins ............................................................................................ 26 Wood Modification with Anhydrides ................................................................... 26 Wood Modification with Maleated Polyolefins .................................................... 28 Surface Analysis Techniques ................................................................................ 31 X-Ray Photoelectron Spectroscopy (XPS) ............................................... 31 Fourier Transform Infrared (F TIR) Spectroscopy .................................... 33 Nuclear Magnetic Resonance (N MR) Spectroscopy ................................ 34 Titrimetric Analysis .................................................................................. 35 Physical and Mechanical Property Testing ........................................................... 36 Density ...................................................................................................... 36 Flexural Properties .................................................................................... 37 Internal Bond Strength .............................................................................. 37 References ............................................................................................................. 39 CHAPTER 3 COMPOSITE MATERIALS MANUFACTURED FROM WOOD PARTICLES MODIFIED THROUGH A REACTIVE EXTRUSION PROCESS ................................ 43 Abstract ................................................................................................................. 44 Introduction ............................................................................................... 45 Experimental ............................................................................................. 48 Materials ....................................................................................... 48 Reactive Extrusion of Wood Particles .......................................... 50 Extraction of Wood Particles ........................................................ 52 Surface Characterization of Wood Particles ................................. 52 Panel Manufacturing and Mechanical Property Testing ............... 53 Statistical Analysis ........................................................................ 54 Results and Discussion ............................................................................. 55 Surface Characterization of Wood Particles ................................. 55 Mechanical Properties ................................................................... 66 Conclusions ............................................................................................... 71 References ................................................................................................. 72 CHAPTER 4 FUNCTIONALIZATION OF WOOD PARTICLES THROUGH A REACTIVE EXTRUSION PROCESS .................................................................................................. 74 Abstract ..................................................................................................... 75 Introduction ............................................................................................... 76 Experimental ............................................................................................. 80 Materials ....................................................................................... 80 Surface Modification of Wood Particles with Maleated Polyolefins in Reactive Extrusion ................................................. 82 Extraction of Wood Particles ........................................................ 84 Surface Characterization of Wood Particles ................................. 85 Results and Discussion ............................................................................. 89 Effect of Maleated Polyethylene (MAPE) Content ...................... 89 Effect of Extrusion Processing Conditions ................................... 99 Effect of Maleated Polypropylene (MAPP) Molecular Weight ......................................................................................... 103 Conclusions ............................................................................................. 1 10 References ............................................................................................... l 12 CHAPTER 5 INFLUENCE OF PROCESSING CONDITIONS AND MATERIAL COMPOSITIONS ON THE PERFORMANCE OF FORMALDEHYDE-F REE WOOD-BASED COMPOSITES ................................................................................................................ 114 Abstract ................................................................................................... 115 Introduction ............................................................................................. 1 16 Experimental ........................................................................................... 1 20 Materials ..................................................................................... 120 Surface Characterization of Wood Particles ............................... 122 vi Panel Manufacture ...................................................................... 124 Panel Property Testing ................................................................ 125 Statistical Analysis ...................................................................... 126 Results and Discussion ........................................................................... 127 Surface Characterization of Wood Particles ............................... 127 Physico-Mechanical Properties ................................................... 135 Density ............................................................................ 135 Effects of Processing Conditions .................................... 137 Effects of Binding Agent Compositions ......................... 140 Comparison with Standard ANSI A2081 ...................... 143 Conclusions ................................................................................. 144 References ................................................................................... 1 46 CHAPTER 6 MODELING AND OPTIMIZATION OF FORMALDEHYDE- FREE WOOD COMPOSITES USING A BOX- BEHNKEN DESIGN ................................................. 148 Abstract ................................................................................................... 149 Introduction ............................................................................................. 1 50 Experimental ........................................................................................... 1 52 Materials ..................................................................................... 152 Experimental Design ................................................................... 152 Compounding and Panel Manufacture ........................................ 155 Property Testing .......................................................................... 156 Results and Discussion ........................................................................... 157 Mechanical Properties ................................................................. 157 Statistical Analysis of the Model ................................................ 159 Modulus of Rupture (MOR) and Internal Bond (IB) Strength ........................................................................... 159 Modulus of Elasticity ...................................................... 162 Numerical Optimization of Mechanical Properties .................... 168 Conclusions ............................................................................................. 1 72 References ............................................................................................... 1 73 CHAPTER 7 SUMMARY OF FINDINGS .......................................................................................... 175 Future Work ............................................................................................ 180 CHAPTER 8 APPENDIX ..................................................................................................................... 183 vii Table 3.1. Table 3.2. Table 3.3. Table 3.4. Table 4.1. Table 4.2. Table 4.3. Table 4.4. Table 4.5. Table 4.6. Table 4.7. Table 4.8. Table 5.1. Table 5.2. Table 5.3. Table 5.4. LIST OF TABLES Characteristics of Maleated Polyolefins ................................................... 49 Formulation Used for Surface Modification of Wood Particles ............... 51 High-Resolution C15 Peaks and Elemental Surface Compositions of Wood Particles Determined by XPS ......................................................... 65 High-Resolution C1, Peaks and Elemental Surface Compositions of Wood Particles Determined by XPS ......................................................... 69 Characteristics of Maleated Polyolefins ................................................... 82 Formulations Used for Modified Wood Particles ..................................... 84 F TIR Absorption Bands and Assignments for Unmodified Wood Particles, Pure MAPE, Pure MAPP and Modified Wood Particles .......... 92 Elemental Surface Compositions and High-Resolution Cl, Peaks of Wood Particles Determined by XPS. ........................................................ 96 Hydroxyl Value, Acid Value and Saponification Value Determined by Titrimetric Analysis. ............................................................................ 97 Effect of Extruder’s Rotational Screw Speed on Surface Chemistry of Wood Particles Modified with 20% MAPE at 160°C. ........................... 103 Effect of Extruder’s Barrel Temperature on Surface Chemistry of Wood Particles Modified with 20% MAPE at 60 rpm. .......................... 103 Elemental Surface Compositions and High-Resolution C15 Peaks of MAPP, Wood Particles, and Wood Particles Modified with MAPP Determined by XPS. ............................................................................... 109 Characteristics of the Maleated Polyolefins Used as Binding Agents.... 122 Grafting Index for Peaks near 2900 cm'1 and 1740 cm'l for Unmodified and Modified Wood Particles with Various Maleated Polyolefin Compounds ............................................................................................. 133 High-Resolution C1, Peaks of Wood Particles Determined by XPS ...... 135 Density Data for Experimental Panels Bound with Maleated Polyolefins .............................................................................................. 137 viii Table 5.5. Table 5.6. Table 6.1. Table 6.2. Table 6.3. Table 6.4. Table A]. Table A2. Table A3. Table A.4. LIST OF TABLES (CONT’D) Effects of Processing Methods and Material Compositions on the Mechanical Properties of Particleboard Panels Bound with Maleated Polyolefins .............................................................................................. 1 39 Effect of Molecular Weight/Maleic Anhydride Content of MAPP on the Mechanical Properties of Particleboard Panels Bound with Maleated Polypropylenes ........................................................................................ 142 Box-Behnken Design Matrix in terms of Both Actual and Coded Factor Levels Generated by Design Expert Software ........................................ 155 Standard Property Requirements for Various Grades of Particleboard of Medium Density (640-800 kg/m3) .......................................................... 159 Analysis of Variance (ANOVA) for Response Surface Quadratic Model ...................................................................................................... 164 Numerical Optimization Settings and Results ........................................ 171 Standard Property Requirements for Particleboard and Medium Density Fiberboard ............................................................................................... 1 84 Mechanical Property Data for Experimental Maple Panels Bound with MAPE ..................................................................................................... 185 Mechanical Property Data for Experimental Maple Panels Bound with MAPP ...................................................................................................... 186 Mechanical Property Data for Experimental Panels Bound with 10.5% MAPP, Pressed for 6 Minutes at 180°C and 3.4 MPa Pressure .............. 189 ix Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. Figure 2.6. Figure 2.7. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 3.5. Figure 4.1. Figure 4.2. LIST OF FIGURES Chemical structures of (a) cellulose and (b) lignin. .................................. 12 Cross section of a ponderosa pine log showing growth rings ................... 14 SEM micrographs of (a) hardwood red oak at 100X magnification and (b) softwood white pine at 150X ........................................................ 16 The chemistry of PF resin (a) resole and (b) novolac. .............................. 20 Chemistry of urea-formaldehyde resin ..................................................... 22 Chemistry of melamine-formaldehyde ..................................................... 24 Reaction scheme for the modification of wood particles with maleated polyolefins ................................................................................................. 29 Modification scheme for esterification reaction between wood particles and maleated polyolefins: (a) mono-ester and (b) di-ester formation ....... 56 FTIR spectra of unmodified wood particles (A), maleated polyethylene- MAPE (B), MAPE-modified wood particles without extraction (C), with 24 hour-Soxhlet extraction (D), and after a second 24 hour-Soxhlet extraction (E) in the region 4000-400 cm". .............................................. 58 FTIR spectra of unmodified wood particles (A), maleated polypropylene- MAPP (B), MAPP-modified Wood particles without extraction (C), with 24 hour-Soxhlet extraction (D), and after a second 24 hour-Soxhlet extraction (E) in the region 4000-400 cm‘l ............................................... 59 Solid-state '3 C NMR spectra of (A) MAPE, (B) unmodified wood particles, and (C) wood particles modified with MAPE ........................... 63 Panel manufacture scheme for modified wood particles where R is an ethylene or propylene repeat unit and R’ is hydrogen or a methyl group ......................................................................................................... 67 Modification scheme for esterification reaction between wood particles and maleated polyolefins: (a) monoester and (b) diester formation. ........ 77 Diagram of the extruder showing the three heating zones ........................ 84 Figure 4.3. Figure 4.4. Figure 4.5. Figure 4.6. Figure 4.7. Figure 4.8. Figure 5.1. Figure 5.2. Figure 5.3. Figure 5.4. LIST OF FIGURES (CONT’D) F TIR spectra of unmodified wood particles (spectrum A), pure maleated polyethylene-MAPE (spectrum B), wood particles modified with 5% MAPE (spectrum C), 10% MAPE (spectrum D), 15% MAPE (spectrum E), and 20% MAPE (spectrum F) after a second 24-hour Soxhlet extraction in the region4000-400 cm". ..................................................... 91 Grafting index for FTIR absorbance bands near 2900 cm'l and 1740 cm’1 for unmodified wood particles and wood particles modified with 5-20% MAPE ....................................................................................................... 94 Effect of processing conditions on grafting index for unmodified and modified wood particles over the range of extruder barrel temperature and rotational screw speed combinations: (a) band near 2900 cm'1 and (b) band near 1740 cm". ......................................................................... 101 FTIR spectra of pure MAPP compounds with various molecular weights: 11,200 g/mol or E-43 (spectrum A), 39,000 g/mol or G-3216 (spectrum B), 47,000 g/mol or G-3015 (spectrum C) and 52,000 g/mol or G-3003 (spectrum D) in the region 4000-400 cm". ..... ' ........................................ 105 F TIR spectra of unmodified wood particles (spectrum A), wood particles modified with MAPP compounds of various molecular weights: 11,200 g/mol or E-43 (spectrum B), 39,000 g/mol or G-3216 (spectrum C), 47,000 g/mol or G-3015 (spectrum D), and 52,000 g/mol or G-3003 (spectrum E) in the region 4000-400 cm". ............................................. 106 Effect of molecular weight on grafting index for unmodified wood particles and wood particles modified with 20% MAPP compounds ..... 108 Modification scheme for esterification reaction between wood particles and maleated polyolefins: (a) monoester and (b) diester formation ................................................................................................. 129 Panel manufacturing scheme for the two-step method showing the bonding of pre-reacted wood particles .................................................... 129 Example FTIR spectra of unmodified wood particles (A), pure MAPEG- 2608 (B), and wood particles modified with MAPEG-2608 (C) in the region 4000 to 400 cm" .......................................................................... 131 X-ray density profile of a sample made from unextruded MAPP G-3003, illustrating the face and core regions of a typical sample ......... 137 xi Figure 6.1. Figure 6.2. Figure 6.3. Figure A. 1. Figure A.2. LIST OF FIGURES (CONT’D) Cube graphs of the linear relationship between mechanical property results and press temperature, pressing time, and binding agent concentration for (a) MOR and (b) 18 strength ...................................... 162 Perturbation plot of square root of MOE against pressing temperature (A), pressing time (B) and binding agent concentration (C) ................... 166 Interaction plots of the variation in square root of MOE as a function of the interaction between pressing temperature and binding agent concentration at (a) low press time (3 minutes) and (b) high press time (9 minutes) .............................................................................................. 168 SEM images of fracture surfaces of panels manufactured from maple with maleated polypropylene, G-3003 at 250x magnification. (a) 3%, (b) 10.5% and (c) 18% MAPP ...................................................................... 187 SEM images of fracture surfaces of panels manufactured from maple with maleated polypropylene, G-3003 at 500X magnification. (a) 3%, (b) 10.5% and (c) 18% MAPP ...................................................................... 188 xii AV EPA DRIFT FTIR GI HI IB MAPE MAPP MDF MF MFI MOE MOR MW NMR OSB PB PE PP PF SV UF WPC XPS LIST OF ABBREVIATIONS Acid Value Environmental Protection Agency (US) Diffuse Reflectance Fourier Transform Fourier Transform Infrared Spectroscopy Grafting Index Hydroxyl Index Internal Bond Maleated Polyethylene Maleated Polypropylene Medium Density Fiberboard Melamine Formaldehyde Melt Flow Index Modulus of Elasticity Modulus of Rupture Molecular Weight Nuclear Magnetic Resonance (Spectroscopy) Oriented Strandboard Particleboard Polyethylene Polypropylene Phenol Formaldehyde Resorcinol Formaldehyde Saponification Value Urea Formaldehyde Wood-Plastic Composite X-Ray Photoelectron Spectroscopy xiii CHAPTER 1 INTRODUCTION Wood-based composites are a multi-billion dollar industry in North America (1). Manufacturing plants produce vast quantities of reconstituted wood products such as particleboard, oriented strandboard (OSB), medium density fiberboard (MDF) and many others. These composite products are commonly made using formaldehyde-based adhesives, including urea-formaldehyde, melamine-formaldehyde, and phenol- formaldehyde (1-3). In 1998, 1,780 kilotons of adhesive resin solids were used to produce primary glued wood products (excluding the adhesive used to bond fumiture and other secondary wood products). Of this amount, nearly 92% were formaldehyde-based adhesives (1). Wood-based composite products find wide use in building construction, where they are utilized as roof sheathing, wall board, floor underlayment, wood I-joists and a number of other applications (1 -3). The furniture industry is also a major consumer of wood composites, notably particleboard and medium density fiberboard, which are commonly overlaid with veneer to make cabinets, shelving, tables and more (1 -3). The use of wood composites bound with formaldehyde-based adhesives in indoor applications is a concern because these adhesives are known to release formaldehyde both during panel pressing and service life (1, 4-8). Plants that produce wood composites using formaldehyde-based adhesives emit harmfiil chemicals to the environment. These include phenol, formaldehyde, ketones, and other volatile organic compounds (1, 4-7), which are classified as hazardous air pollutants (HAPs) by the United States Environmental Protection Agency (EPA) (7). Wood composites bound with formaldehyde-based adhesives also release formaldehyde over time (1, 7, 9). Although wood composites made today emit far less formaldehyde than those made 20 years ago, the problem has not been eliminated (1). Formaldehyde and other toxic compounds may be present in large amounts in both indoor and outdoor air as a consequence of the use of these adhesives (4-8). As a result of public concern about the environment, the US EPA enacted new emission standards for facilities that manufacture plywood and composite wood products in September of 2004 (7). These rules affect both new and existing plants that produce at least 10 tons of any one HAP per year, or any combination of 25 tons of HAPs per year. These regulations may spur industry to find new ways to bind composite products without the use of formaldehyde-based adhesives. In recent years, there have been several studies into environmentally friendly wood composites (1, 9). Some of the areas that have been investigated include urea- formaldehyde adhesives with low formaldehyde-to-urea molar ratios, and the development of natural adhesives such as tannin, lignin, soybean and cornstarch adhesives (l, 9), phenol-formaldehyde resins modified with either lignin (10-12) or tannin (13) and adhesive made from decayed wood (14). Additives that reduce formaldehyde release during composite pressing and during board use have also been developed (1, 9). Different types of fiberboard have been made without synthetic adhesives using steam explosion (15), wood surface activation using Fenton’s reagent (16) and wood fiber binding via lignin activation (17). A cellulose-based composite using allyl glycidyl ether grafted polyethylene as a binder was recently reported (18). Esterification of wood particles or other cellulosic material with anhydrides in a solvent system has also been used as a basis for formaldehyde-free composites (19-25). Although several approaches have been developed to reduce formaldehyde emissions, some of them cannot be implemented in the industrial production of wood composite products because the processes are not cost-effective, not environmentally friendly, or lead to products with undesirable properties such as long press time, dark coloration of the panel, low resistance to moisture, moderate strength, etc. Control of emissions from wood composite mills and glued wood products is still one of the major challenges facing the wood composites industry since the government is continuously developing and implementing stringent regulations to eliminate formaldehyde emission into the environment. A new approach to reduce and/or eliminate formaldehyde emissions from wood composite products will be addressed in this research project. The technology proposed in this project is based upon the reaction between the carboxylic acid groups of maleated polyolefins and hydroxyl groups on the surface of wood. This reaction results in the grafting of a pendent polyolefin chain to the wood surface through an ester linkage. Several authors have studied this reaction with maleated compounds and wood particles (or other cellulosic materials), usually in an attempt to evaluate maleated compounds as compatibilizers between hydrophilic wood fibers and the hydrophobic polymer matrix in wood-plastic composites (30-35). However, most of this previous work focused on fibers modified through a solvent-based process where the maleated compounds were dissolved in heated xylene, toluene, N,N- dimethylformamide (DMF) or other organic solvents, followed by the addition of cellulosic particles (30-34). Cellulosic fibers that have been grafted with maleated polyolefins have to be filtered and dried after modification, which makes this process both cost-ineffective and time consuming (30-34). Solvent disposal is also an issue with this modification process because some solvents may be harmful to workers and the environment. In order to produce modified wood particles that could be used to make alternative composite panels, it is necessary to change from a solvent-based process to a dry process. A solvent-free process that has been reported in the literature utilized a therrnokinetic mixer to modify wood particles with maleated polypropylene (33, 35). In that work, a heated, high-intensity mixer was used to graft maleated polypropylene onto wood fibers. However, the esterification reaction between wood particles and maleated polypropylene could not be confirmed through surface characterization techniques (33, 35). An effective dry process for grafting maleated polyolefins onto the surface of wood particles is still needed in order to produce the large volume of modified wood particles required to manufacture and commercialize formaldehyde-free wood composite panels. This study will investigate the use of reactive extrusion as a novel dry process for the surface modification of wood particles needed to manufacture formaldehyde-free wood composites. Reactive extrusion is a technique that can be used for the chemical modification of compounds, usually polymers that can melt during processing. Because no solvent is involved during processing, reactive extrusion is an environmentally friendly technique for surface modification. It also allows for control of reaction times and temperatures through variation of rotational screw speeds and extruder barrel temperatures. Reactive extrusion offers the advantage of intensive mixing of components in a short time, especially in a twin-screw extruder, which ensures a more thorough reaction of components (36). This process also has the advantage of low retention times in the extruder, so a series of different reaction conditions can be tested in a relatively short period of time. Examples of reactive extrusion for wood fiber surface modification were not found in the open literature, likely because the wood component does not melt and flow in the extruder. However, the ease of processing and control of reaction conditions makes this technique attractive for wood fiber surface modification. We hypothesized that maleated polyolefins would chemically react with hydroxyl groups on the wood particle surfaces during reactive extrusion (a dry process), resulting in wood particles with surface grafted polyolefin chains. During panel manufacturing, the pendent polyolefin chains attached to the wood particles are expected to melt and flow under heat and pressure in the hot press, forming entanglements. The entangled polymer chains will then lock together after cooling, forming direct particle-particle bonds. This process would result in a formaldehyde-free wood composite panel manufactured without any additional adhesive. Objectives The main goal of this work is to study the concept of using a reactive extrusion process as a means of developing a new, formaldehyde-free binding system for wood composite products. The following specific objectives must be accomplished to achieve the main goal of this project: 1. Evaluate the effects of material composition (binding agent types and content) and extrusion processing conditions (temperature profile and rotational screw speed) on the level of grafting (surface properties) of wood particles after the reactive extrusion process; 2. Characterize the surface of unmodified and modified wood particles in terms of chemical compositions (both elemental and functional groups); 3. Manufacture composites and evaluate their physico-mechanical properties; 4. Establish the relationships between material composition, processing conditions, and physico-mechanical properties in order to identify the major factors that govern the performance of formaldehyde-free wood composite products manufactured through a reactive extrusion process. References 10. 11. 12. 13. Sellers, T., Jr., "Wood adhesive innovations and applications in North America," Forest Products Journal, 51 (6): 12-22 (2001). Maloney, T.M., "The family of wood composite materials," Forest Products Journal, 46 (2): 19-26 (1996). Guss, L.M., "Engineered wood products: the future is bright," Forest Products Journal, 45 (7/8): 17-24 (1995). Barry, AD. and Comeau, D., "Volatile organic chemicals emissions from OSB as a function of processing parameters," Holzforschung, 53 (4): 441-446 (1999). Barry, A., Comeau, D., and Lovell, R., "Press volatile organic compound emissions as a function of wood particleboard processing parameters," Forest Products Journal, 50 (10): 35-42 (2000). Barry, A., Lepine, R., Lovell, R., and Raymond, S., "Response surface methodology study of VOCs in plywood press emissions," Forest Products Journal, 51 (1): 65-73 (2001). US-EPA, "National emission standards for hazardous air pollutants: plywood and composite wood products," Federal Register, 69 (146): 45943-46046 (2004). Maloney, T.M., Modern Particleboard and Dry-Process F iberboard Manufacturing, Updated Ed., Miller Freeman: San Francisco (1993). Pizzi, A., Wood Adhesives : Chemistry and Technology. Marcel Dekker: New York (1983). Matuana, L.M., Riedl, B., and Barry, A.O., "Kinetic characterization by DTA of lignosulfate-based phenol-formaldehyde resins," European Polymer Journal, 29 (4): 483-90 ( 1993). Kazayawoko, J.S.M., Riedl, B., Poliquin, J ., Barry, A.O., and Matuana, L.M., "A lignin-phenol-formaldehyde binder for particleboard. Part 1. Thermal characteristics," Holzforschung, 46 (3): 257-62 (1992). Sellers, T., Jr., McGinnis, G.D., Ruffin, T.M., and Janiga, E.R., "Lignin-modified phenol-formaldehyde resin development for fiberboard," Forest Products Journal, 54 (9): 45-51 (2004). Sellers, T., Jr. and Miller, G.D., Jr., "Laboratory manufacture of high moisture southern pine strandboard bonded with three tannin adhesive types," Forest products journal, 54 (12): 296-301 (2004). 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. Li, K. and Geng, X., "Formaldehyde-free wood adhesives from decayed wood," Macromolecular Rapid Communications, 26 (7): 529-532 (2005). Velasquez, J .A., Ferrando, F., Farriol, X., and Salvado, J ., "Binderless fiberboard from steam exploded Miscanthus Sinensis," Wood Science and Technolog, 37 (3-4): 269-278 (2003). Widsten, P., Qvintus-Leino, P., Tuominen, S., and Laine, J .E., "Manufacture of fiberboard from wood fibers activated with F enton's reagent (H202/FeSO4)," Holzforschung, 57 (4): 447-452 (2003). Yelle, D., Goodell, B., Gardner, D.J., Amirbahman, A., Winistorfer, P., and Shaler, S., "Bonding of wood fiber composites using a synthetic chelator-lignin activation system," Forest Products Journal, 54 (4): 73-78 (2004). Casarano, R., Matos, J .R., Fantini, M.C.A., and Petri, D.F.S., "Composites of allyl glycidyl ether modified polyethylene and cellulose," Polymer, 46 (10): 3289-3299 (2005) Matsuda, H., "Preparation and utilization of esterified woods bearing carboxyl groups," Wood Science and Technology, 21 (1): 75-88 (1987). Matsuda, H., Ueda, M., and Mori, H., "Preparation and crosslinking of oligoesterified woods based on maleic anhydride and allyl glycidyl ether," Wood Science and Technology, 22 (1): 21-32 (1988). Matsuda, H., "Thermal plasticization of lignocellulosics for composites," in Emerging Technologies for Materials and Chemicals from Biomass R.M. Rowell, T.P. Schultz, and R. Narayan, Editors. American Chemical Society: Washington, DC. p. 98-114 (1992). Clemons, C., Young, RA, and Rowell, R.M., "Moisture sorption properties of composite boards from esterified aspen fiber," Wood and Fiber Science, 24 (3): 353-63 (1992). Timar, M.C., Mihai, M.D., Maher, K., and Irle, M., "Preparation of wood with thermoplastic properties. Part 1. Classical synthesis," Holzforschung, 54 (1): 71- 76 (2000). Timar, M.C., Maher, K., Irle, M., and Mihai, M.D., "Preparation of wood with thermoplastic properties. Part 2. Simplified technologies," Holzforschung, 54 (1): 77-82 (2000). Timar, M.C., Maher, K., Irle, M., and Mihai, M.D., "Thermal forming of chemically modified wood to make high-performance plastic-like wood composites," Holzforschung, 58 (5): 519-528 (2004). 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Hon, D.N.S. and Xing, L.M., "Thermoplasticization of wood. Esterification," in Viscoelasticity of Biomaterials, W.G. Glasser and H. Hatakeyama, Editors. American Chemical Society: Washington DC. p. 118-132 (1992). Marcovich, N.E., Reboredo, M.M., and Arangguren, M.I., "Sawdust modification. Maleic anhydride chemical treatment," Holz als Roh- und Werkstojf 54 (3): 189- 193 (1996). Marcovich, N.E., Aranguren, M.I., and Reboredo, M.M., "Modified woodflour as thermoset fillers Part I. Effect of the chemical modification and percentage of filler on the mechanical properties," Polymer, 42 (2): 815-825 (2001). Marcovich, N.E., Reboredo, M.M., and Aranguren, M.I., "Modified wood flour as thermoset fillers. 11. Thermal degradation of wood flours and composites," T hermochimica Acta, 372 (1-2): 45-57 (2001). Felix, J .M. and Gatenholm, P., "The nature of adhesion in composites of modified cellulose fibers and polypropylene," Journal of Applied Polymer Science, 42 (3): 609-20 (1991). Matuana, L.M., Balatinecz, J .J ., Sodhi, R.N.S., and Park, C.B., "Surface characterization of esterified cellulosic fibers by XPS and FTIR Spectroscopy," Wood Science and Technology, 35 (3): 191-201 (2001). Li, Q. and Matuana, L.M., "Surface of cellulosic materials modified with functionalized polyethylene coupling agents," Journal of Applied Polymer Science, 88 (2): 278-286 (2003). Kazayawoko, M., Balatinecz, J .J ., and Woodhams, R.T., "Diffuse reflectance Fourier transform infrared spectra of wood fibers treated with maleated polypropylenes," Journal of Applied Polymer Science, 66 (6): 1163-1173 (1997). Kazayawoko, M., Balatinecz, J .J ., and Matuana, L.M., "Surface modification and adhesion mechanisms in wood fiber-polypropylene composites," Journal of Materials Science, 34 (24): 6189-6199 (1999). Kazayawoko, M., Balatinecz, J .l ., and Sodhi, R.N.S., "X-ray photoelectron spectroscopy of maleated polypropylene-treated wood fibers in a high-intensity therrnokinetic mixer," Wood Science and Technology, 33 (5): 359-372 (1999). Cheremisinoff, N.P., Guidebook to Extrusion Technology. Prentice Hall: Englewood Cliffs, NJ. (1993). CHAPTER 2 BACKGROUND AND LITERATURE REVIEW Chemical Composition of Wood Wood is a complex, multi-component material composed of several compounds. Wood is primarily composed of cellulose, hemicellulose and lignin, with minor components extractives and ash. Although amounts of these components vary between species, wood composition is approximately 40-45% cellulose, 15-30% hemicellulose, 20-30%1ignin, 1-10% extractives and 0-1% ash (1-4). Cellulose Cellulose is the primary component of wood. Carbon, hydrogen and oxygen are arranged into sugar molecules during photosynthesis in the living tree. These B-linked glucose units are further linked together to form long, linear chains. Cellulose chains are generally oriented in one direction, and give strength to the cell wall. The orientation of cellulose is largely due to the presence of a large number of hydroxyl groups. This allows the chains of cellulose to interact and hydrogen bond with adjacent chains, forming crystalline regions within the wood. However, not all of the cellulose is ordered into crystalline domains. Some portion of cellulose is amorphous, and in these regions the hydroxyl groups are free to form hydrogen bonds with water molecules. This accounts for the hygroscopic nature of wood, as water is held in the amorphous portions of cellulose. Water cannot easily enter the crystalline portions of wood as the cellulose chains there are tightly packed and held together by hydrogen bonding across their 10 hydroxyl groups. There are essentially no free hydroxyl groups available to hold water in the crystalline regions of wood. The amorphous region of cellulose is also the reactive site for wood modification as the hydroxyl groups are accessible for reaction with other compounds (1-4). Cellulose structure is illustrated in Figure 2.1a. Hemicellulose Hemicellulose is structurally similar to cellulose, as it is also composed of sugar molecules. However, hemicellulose is formed from various 5 and 6 carbon sugars, connected in short, often branched chains. Hydroxyl groups are also present in the hemicellulose, accounting for some of the hygroscopic nature of wood. Because of the short chain structure, most of the hemicellulose is soluble in water, and can be removed by water extraction (1-4). Lignin Like cellulose and hemicellulose, lignin is formed from carbon, hydrogen and oxygen. However, instead of taking the form of a polysaccharide, it is a phenolic compound which exists in a variety of forms. Phenylpropane units form the basic structure, linked together and branched in various ways. The structure of lignin is illustrated in Figure 2.1b. Lignin is primarily concentrated in the outer portions of wood cells and between the wood cells, where it fimctions as a binding agent to hold cells together. It is a brittle, stiff material that strengthens the cell. There are fewer hydroxyl groups in lignin than cellulose, which makes lignin less reactive (1-4). 11 3 /°\ :13 O l:\ O M 3 0—0 5° :1: o’ \a;_e 230:1: O m ‘2 o :1:/\ 6 :1: 0—0 nocrr2 Figure 2.1. Chemical structures of (a) cellulose and (b) lignin (3) 12 Extractives Extractives are a class of organic compounds that includes resins, carbohydrates, waxes, tannins, fats, oils, acid, etc. They are called extractives because they are not strongly bound to the wood structure and thus can be easily removed or extracted through processes such as steam distillation and solvent extraction. Extractives are largely responsible for the smell, color, density, flammability and fungal resistance of wood (1- 4). Wood Anatomical Structure Growth rings are one of the most distinctive features of wood. They are noticeable on a cross section of wood or on the stump remaining when the tree is cut. Figure 2.2 shows an example of a tree cross section with clearly defined growth rings. The age of a tree can be determined by counting the rings, as each ring represents one year of growth. The light-colored portion of the ring is called earlywood and is composed of large, conductive cells that form early in the growing season. In contrast, the dark portion of the ring is latewood, which is composed of cells with thicker walls and more strength. The thickness of the rings and the proportion of earlywood to latewood varies between species, and is also affected by growing conditions (3, 4). 13 Figure 2.2. Cross section of a ponderosa pine log showing growth rings (4) Opt 0f V65 Ba 31 The cells of wood are long and narrow with hollow centers called lumens. Openings called pits connect the cells along their length (3, 4). There are two main types of cells within wood: (i) fibers and (ii) vessels. Fibers give the wood strength, while vessels conduct liquids. Both the size and quantity of these cells differ within a tree. Based on these differences, wood is categorized as softwood or hardwood (3, 4). Sofiwoods, or coniferous trees, are characterized by a porous structure formed by a large quantity of fiber tracheids. There are no vessels present, although some softwood has resin canals that transport materials in the tree (3, 4). Softwood fibers are rectangular and very long, approximately 3-8 mm, and are oriented in nearly straight rows (3, 4). The fiber tracheids lend strength to the tree, while also conducting liquids in the vertical direction. Groups of cells called rays are also present to move sap horizontally in the living tree (3, 4). Softwood structure is illustrated in Figure 2.3. Hardwoods, or deciduous trees, contain both fibers and vessels, randomly oriented within the wood. An example of hardwood structure is illustrated in Figure 2.3. Hardwood fibers are circular and shorter in length than those in softwood (3, 4). A typical hardwood fiber averages only 1 mm in length (3, 4). Vessels form vertical canals for the transport of materials within the tree. Hardwoods also have rays to move materials horizontally through the tree (3, 4). 15 Figure 2.3. SEM micrographs of (a) hardwood red oak at 100X magnification and (b) softwood white pine at 150x (3) Mo Moisture Content Wood is a hygroscopic material (1-4). As a result, wood moisture content varies depending on the moisture content and temperature of its surroundings. The fiber saturation point is defined as the point where the lumens do not contain any free water but the cell wall is fully saturated (3). Below this point, moisture content of wood plays a large role in the dimensional stability, as wood shrinks and swells in response to moisture changes (3). Shrinking and swelling of wood elements can cause problems when wood is used in composites, as these changes weaken glue bonds and reduce interfacial adhesion between fibers and matrix components. Wood Composites Glued wood composites can be made from wood of varying geometry, including products based on lumber, sheets of veneer, strips of wood, flakes, particles, fibers and even wood flour (3-7). In the case of this research, wood particles were used to produce formaldehyde-free composites via reactive extrusion. The extrusion process limits the size of the wood material that can be used, as larger fibers or particles would be broken down by the shearing action of the screws in the twin-screw extruder. Glued wood composites made from wood particles or fibers would be most similar to the panels produced via reactive extrusion, thus these will be described in detail. Medium Density F iberboard (MDF) MDF is a panel product composed of wood fibers and a thermoset adhesive. Urea-formaldehyde (UP) is the most common adhesive used in MDF, although some 17 isocyanate-p-MDI is used for certain applications. Panels are hot pressed to cure the resin and consolidate the fibers to the medium density range, between 500 and 800 kg/m3. MDF panels are often covered with veneer and used in cabinetry and moldings (6, 8). Particleboard Particleboard is a panel product made from discrete particles of wood and UP adhesive. The wood and adhesive mixture is formed into a mat and then pressed in a heated press to consolidate the panel. Particleboard is used in some structural applications, such as flooring and wallboard, although it finds a majority of its use in furniture applications such as shelving, in which it is covered with a layer of veneer or laminate (6, 8). Adhesives for Wood Composites Two categories of adhesives dominate in the production of glued wood composite products. These are phenolic resins and amino resins, which combined made up over 90% of the adhesives used in wood composites in 1998 (7). Phenolic resins include phenol-formaldehyde (PF) and resorcinol-formaldehyde (RF), whereas amino resins include urea-formaldehyde (UP) and melamine-formaldehyde (MF) (3, 7, 9). Polymeric diphenyl methylene diisocyanate (PMDI) is also used for some specialty applications (3, 7). Phenolic and amino resins are discussed in detail. 18 Phenolic Resins Phenol-formaldehyde (PF) resins are made by the reaction of phenol with formaldehyde. There are two basic types of PF resin, resole and novolac (3, 9). The structures of these resins are shown in Figure 2.4. Resole PF is made through the reaction of phenol with excess formaldehyde in the presence of an alkaline catalyst. This results in a resin with a branched structure. The presence of reactive methylol groups allows this type of PF to self-cure with the application of heat, and thus it does not need an added catalyst (3). Resoles are dark yellow to dark brown in color. Novolac PF resins are formed from the reaction of excess phenol with formaldehyde under acidic conditions (3). This produces a resin with a linear structure. Novolacs cannot self-cure because there are no residual methylol groups in the structure. As a result, a curing agent must be added to these resins in order to form crosslinks. These resins are lighter in color than the resole PF and also have good moisture resistance. Resorcinol-formaldehyde (RF) is another type of phenolic resin used in the wood industry. This adhesive is produced by the reaction of resorcinol with formaldehyde. RF resins have two reactive hydroxyl groups in their structure, which allows them to cure rapidly at room temperature with the addition of a catalyst (3). RF is expensive to produce, but is still desirable for high moisture resistance of products made with this adhesive (3). 19 OH Phenol Formaldehyde Monomethylolphenol :—n—: 5 0-2 $-61 2 Hon/— Methylene bridge OH 110 """ alk '40 _________ ©—E—Ofl__ Z—OH + HOH O" P: F 1 Heat and Novolac formaldehyde donor Molecular grth 0H 1" OH I“ (RH I"! OH I. OH I" PH Oi at: cou * Figure 2.4. The chemistry of PF resin (a) resole and (b) novolac (3) 20 Am ad’: 8111 for C0- bi Amino Resins Urea-formaldehyde (UF) resins are the least expensive of the major wood adhesives. They are formed from the polymerization of excess formaldehyde with urea, and react as the four N-H bonds on the urea add across the carbonyl group on the formaldehyde (Figure 2.5). This produces methylol groups, which are suspended in water at about 65% resin solids when sold. The methylol groups crosslink by condensation when an acid catalyst is added, and after final curing the resin is hard, brittle and insoluble. However, cured UP is sensitive to moisture, which causes the resin to break down and release formaldehyde (3, 9). Formaldehyde release from UF bonded wood products is also a concern, especially in light of the new EPA regulations on formaldehyde emissions from wood composite manufacturing facilities (10). 21 iii-P * lr—riiif" Urea Formaldehyde Monomethylolurea H Ill—:1 i'iiom'f'iiJN-c-r't {mu 0 ll H m-itriri7 *"°" 9’“ Molecular growth H- H and cross-linking +30 l-H’ H-C-H iiiii li.i.i.. ......... “-c- - - - - - - "i Til it. 3 M43 T-H “‘30 + XflOl-l 02C 0 H at?" I! H ii"- -12-- 3.2-2- -2-.-,u-.-§-.~.-i... a ti 3 ti 1 H a ll- -ll H | :31! 0-3 fl-N ' i-0 ,3- Sin eel. 6-n 0-3 Figure 2.5. Chemistry of urea-formaldehyde resin (3) 22 Melamine-formaldehyde (MF) resins are also widely used in wood composites. These resins are produced through the reaction of melamine with formaldehyde as illustrated in Figure 2.6. This produces a resin that is somewhat water resistant because of the low solubility of melamine in cold water. MF glues can be used for interior and exterior applications, but their high cost is a large drawback. MF resins are more expensive than PF, and offer less durability for an increased price. Often, MF and UP resins are blended to produce a resin with lower cost than MF with improved properties compared to UP alone (3, 9). 23 Ii H l I " i / h/ \N r/ N Ii! 3 I I" T I! H J H II! Polymerizes and H—N— C— N—H + (II=O —~ H—N— —N—C—OH-—-° cross-links like I . \/ H \/ H urea resrn Melamine Formaldehyde Monomethylol melamine Figure 2.6. Chemistry of melamine-formaldehyde resin (3) 24 Wood-Plastic Composites Wood plastic composites (WPCs) are a versatile family of composite materials made from wood fibers, particles, or flour mixed with a thermoplastic polymer and formed into a variety of shapes (11, 12). WPCs are used in a variety of applications such as decking, fencing, docks, automotive applications, playground equipment, etc (12). The decking market has been the largest volume application for WPCs, consuming over 50% of the total WPC produced in 2002 (12). When well—made, WPCs have several advantages over wood or plastic alone such as resistance to fungal decay, high impact strength, low incidence of cracking, etc (1 l, 12). WPCs are usually manufactured by extrusion, where wood and plastic are melted together and compounded into a composite material (12). The composite can be extruded into lengths of material that resemble lumber, or into profiles with hollow interiors or various layers. More complex shapes can be produced through injection molding, in which the composite is injected into a mold while still melted. The compounded mixture cools in the mold and is removed once solid. Despite the advantages offered by WPCs, one significant drawback to their use is the lack of compatibility between wood, which is hydrophilic due to the large content of cellulose, and the hydrophobic plastic matrix (13-19). This incompatibility results in reduced strength properties and lower quality products. In order to overcome this drawback, coupling agents have been developed to compatibilize the wood and polymer components (13-19). The most successful examples of WPC coupling agents for polyolefin-based WPCs are maleated polyolefins, which bond with the wood component, while diffirsing into the plastic matrix. 25 Maleated Polyolefins Maleated polyolefins are usually polypropylene or polyethylene polymer that has been grafted with a small amount of maleic anhydride. These polymers are mainly used as coupling agents to promote interfacial adhesion in wood polymer composites between the non-polar plastic and polar wood components (13-19). The carboxylic acid portion of the maleated compound reacts with hydroxyl groups on the wood surface to form an ester linkage (13, 14, 17). The polymer chain is then free to diffuse into the polymer matrix, forming a physical bond upon cooling. This coupling mechanism is an effective method to promote compatibility between the phases in wood composites (13, 19-21). A small amount of maleated polyolefins (1-5%) is generally used to promote adhesion between the phases (13, 17, 19, 20). Wood Modification with Anhydrides Another method of compatibilizing wood and plastic in composite materials is to chemically modify the wood component. Chemical modification of wood with anhydride compounds has been studied extensively for several decades (22-32). Wood reacts with various anhydrides to form ester groups on the wood surface. Esterification has been shown to plasticize the wood, making it more versatile for wood composite applications. Several investigators have studied wood modification with anhydrides in order to determine thermal properties, moisture resistance, compatibility with polymeric matrices, etc. (22-32). A lower softening temperature, compared to unmodified wood, has been documented in esterified wood after reaction with various anhydrides (22, 27-29). Esterified wood has also been found to resist moisture (22-28, 30-32) and to be more 26 C01 C01 for C01 1hr ch \l' Ci Eli compatible than unmodified wood with polyester matrices in the manufacture of composites (22-28, 30-32). A second reaction of the esterified wood with epoxides formed oligoesters, which further thermoplasticized the wood and allowed for better control of structure and resulting properties of the modified wood (22-24, 26-28). While the thermoplastic nature of esterified and oligoesterified wood modified through these methods has been documented, few authors have exploited this characteristic to make wood composites without added adhesive. Matsuda and co- workers used three dicarboxylic acid anhydrides (maleic, phthalic and succinic anhydride) to esterify wood particles, which were then molded into sheets through compression molding (22, 24). Clemons and co-workers used similar methods to modify wood fibers with anhydrides to prepare fiberboards (25). In both studies, chemically modified wood was able to partially melt when heated and bond without additional adhesive. Formulations containing succinic anhydride were found to have the most thermoplastic character in both studies (22, 24, 25). Further work by Matsuda et a1. (22- 24) showed enhanced thermoplasticization of wood through the grafting of various types of epoxides onto the already esterified wood surface. The oligoesterified wood produced from these reactions was even more thermoplastic-like in structure than the esterified wood produced through reaction with the anhydrides, and could be easily molded. Using various types of epoxides allowed for different chemical structures in the thermoplasticized wood, and could also be used to form crosslinked wood composites which resembled plastic more than wood (22-24). Recently, Timar and co-workers demonstrated the use of maleic anhydride to esterify wood particles, followed by oligoesterification with glycidyl methacrylate and additional maleic anhydride as a 27 process to form thermoplastic-like wood particles (26-28). Panels formed through the compression molding of modified wood particles displayed mechanical properties that met or exceeded standard requirements for bending strength and internal bond strength of particleboard or fiberboard (28). These wood composites were also found to be resistant to fungal decay. Although esterifying and oligoesterifying the wood surface has been shown to be an effective method for manufacturing wood composite materials without additional adhesive, most methods require harmful organic solvents such as xylene, dimethyl sulfoxide, or N,N-dimethylformamide as part of the modification process (22, 24-32). Additionally, the oligoesterification reactions entail a two-step process (or more depending on the precise control of chemical structure desired) to functionalize the wood surface, which is both time consuming and complicated (22-24, 26-28). Removal of solvents and further drying of the wood particles is required before compression molding into panels (22. 24, 25, 28). Modification of Wood with Maleated Polyolefins Maleated polyolefins also react with wood particles through an esterification reaction (Figure 2.7). This reaction occurs between the maleated groups of the polyolefins and hydroxyl groups on the wood surface, forming a monoester (Figure 2.7.a) or a diester (Figure 2.7.b). This results in the grafting of a pendent polyolefin chain to the wood surface through the ester linkage. 28 //// ii / lFO‘C_ Cc? Monoester Polyolefin ; OH CH M 0 chain [— / a) / HO’CE l / —0H %CH2 1 / O / t 0 CIH or ; _0H % \/W \ C? Polyolefin I . O b) / O—~C—CH2 chain Wood Maleated / particle polyolefin / o— c —C II H / o Modified wood Diester particles Figure 2.7. Reaction scheme for the modification of wood particles with maleated polyolefins 29 Several authors have studied this reaction with maleated compounds and wood particles (or other cellulosic materials), usually in an attempt to evaluate maleated compounds as compatibilizers between hydrophilic wood fibers and the hydrophobic polymer matrix in wood-plastic composites (13-18). However, most of this previous work focused on fibers modified through a solvent-based process where the maleated compounds were dissolved in heated xylene, toluene, N,N-dimethylformamide (DMF) or other organic solvents, followed by the addition of cellulosic particles (13-17). As with wood particles esterified or oligoesterified with other anhydride compounds, cellulosic fibers that have been grafted with maleated polyolefins have to be filtered and dried after modification, which makes this process both cost-ineffective and time consuming (13- 17). Solvent disposal is also an issue with this modification process because some solvents may be harmful to workers and the environment. In order to produce modified wood particles that could be used to make alternative composite panels, it is necessary to change from a solvent-based process to a dry process. A solvent-free process that has been reported in the literature utilized a thermokinetic mixer to modify wood particles with maleated polypropylene (16, 18). In that work, a heated, high-intensity mixer was used to graft maleated polypropylene onto wood fibers. However, the esterification reaction between wood particles and maleated polypropylene could not be confirmed through surface characterization techniques (16, 18). 30 Surface Analysis Techniques Several techniques are employed in the surface analysis of wood-based materials. These techniques include X-ray photoelectron spectroscopy (XPS), Fourier transform infrared (FTIR) spectroscopy, and nuclear magnetic resonance (NMR) spectroscopy. These techniques are frequently used together to determine chemical composition, structure and information about reactions in materials. An additional chemical characterization technique for wood-based materials is titration analysis, which is used to determine bulk chemical changes through acid-base chemistry. Contact angle measurements are often used to determine wettability and surface tension of wood-based materials. Inverse gas chromatography can be also used to determine thermodynamic properties of modified wood particles. Techniques used to analyze chemical changes resulting from modification of wood particles in reactive extrusion with maleated polyolefins in this work will be discussed in more detail. X-Ray Photoelectron Spectroscopy (XPS) XPS is a technique that allows chemical elemental composition of samples to be analyzed in a quantitative manner. A beam of x-ray photons is aimed at the sample, which interacts with the surface and causes electrons to be emitted. The binding energy (Eb) of the emitted electrons is measured by the detector. Since each element has a characteristic Eb, XPS can be used to determine the elemental composition of a sample. This technique is surface sensitive, with a maximum sampling depth of approximately 10 nm, as a result of the limited escape depth of electrons through a solid (33). 31 Typically, a low resolution scan from 0 to 1100 eV binding energy is performed to determine the identity and concentration of each element present on the surface of a sample, along with atomic ratios of elements. For wood products and polymers, carbon and oxygen are often the primary elements of interest. The oxygen to carbon (O/C) atomic ratio is calculated from the low resolution scan to evaluate oxidative changes resulting from chemical reactions or weathering. To further analyze the chemical bonding of the carbon atoms present in a sample, a high resolution scan of the C], region from 280 to 300 eV is performed. F our carbon component peaks can be found from the high resolution scan of wood products. Carbon component C1 arises from carbon atoms bonded only to carbon and/or hydrogen atoms (C-C/C-H) and has a characteristic binding energy of 285 eV. Carbon atoms bonded to a single oxygen atom, other than a carbonyl oxygen (C-OH) constitute the C2 component, which has a binding energy of 286.5 eV. Carbon component C3 arises from carbon atoms bonded to two non-carbonyl oxygen atoms or to a single carbonyl oxygen atom (O-C-O, C=O), and C4 from carbon atoms which are linked to a carbonyl and a non-carbonyl group (O-C=O).9’”'l7 The binding energies of the C3 and C4 components of carbon are 288 and 289 eV, respectively (33). XPS has been used to determine the chemical changes resulting from grafting lignocellulosic fibers with maleated polyolefins (13-15, 17, 18) and to follow surface changes to wood plastic composites after accelerated weathering (34, 35). It has also been used to document surface chemistry changes in wood pulp after ozonation reactions (36, 37). 32 Fourier Transform Infrared (F T IR) Spectroscopy In FTIR spectroscopy, a beam of infrared radiation is used to excite the samples. As a result, infrared radiation is absorbed by organic compounds and converted into vibrational energy (38, 39). The mid IR region (4000 — 400 cm") is most commonly used for analysis of chemical compounds. FTIR can be used to determine the presence of functional groups in liquids or solids, using a variety of sampling methods. For solids, a quantitative method of sample analysis is diffuse reflectance infrared Fourier transform (DRIFT) spectroscopy (39, 40), in which the sample is finely ground and often mixed with a non-absorbing substance such as potassium bromide (40). The method produces a spectrum with bands corresponding to functional groups on the surface of the sample. The spectra are presented in Kubelka-Munk units, which relate sample concentration to the intensity of the bands in the spectra (40). The Kubelka—Munk equation is expressed as: _ 2 (1 R) :5 (1) KR) : 2R 3 Where R is the absolute reflectance of the layer, k is the molar absorption coefficient and s is the scattering coefficient (40). Within the last two decades, FTIR has been used to study wood structure (41). It has been applied to document the changes in the surface of wood after modification with maleic anhydride (23, 26, 27, 29, 30), and to study the esterification reaction between wood and maleated polyolefins (13, 15-17). FTIR has also been used extensively to 33 monitor changes in wood composites as a result of accelerated weathering (34, 35, 42- 44). Nuclear Magnetic Resonance flVMR) Spectroscopy NMR spectroscopy uses a magnetic field to energize the nuclei of selected atoms. Samples with certain nuclei absorb electromagnetic radiation in the radio frequency region (38). The absorption peaks for these nuclei are plotted versus peak intensity in the NMR spectrum. While several atoms can be used for NMR spectroscopy, the two most common forms, 1H and '3 C, are based on hydrogen and carbon nuclei, respectively (38). Both of these methods provide information about bonding and chemical connectivity in samples, though each has different advantages. For example, the integrated area under each peak in the 1H spectra correlates well with the number of protons associated with each peak (3 8). In ”C NMR, the peak areas do not typically correlate with the number of carbon atoms present, so this technique is not quantitative unless run under special conditions. However, each peak in the spectra usually represents a different form of carbon, that is, a different bonding structure (3 8). For example, a carbon atom with two attached hydrogen atoms will have a different chemical shift than carbon with an attached hydroxyl group. Solid state 13C NMR is a useful technique for analysis of solid materials. Solid state '3 C NMR has been used to study cellulose and other lignocellulosic materials and their compounds (45-47). This technique has also been applied to analyzing the carbon structure in wood (45, 46, 48). Solid state '3 C NMR was recently used to determine grafting mechanisms in maleic anhydride grafted polypropylene through reactive extrusion (49). 34 “ll: \\0( lllt‘ r T itrimetric Analysis Titrimetric analysis uses acid-base chemistry to determine the quantity of acidic or basic groups in a substance. Techniques have been developed for the detection of acid value (AV), saponification value (SV) and hydroxyl group content (HV) in alkyd resin samples (50). These analyses have been applied to wood and other cellulosic samples for the detection of changes in modified and grafted wood (1 3, 23, 26, 27, 29). Acid value determines the amount of free carboxylic acid groups present in a sample, while saponification value measures both acid and ester groups in a sample. Hydroxyl value quantifies the free and accessible hydroxyl groups present in the sample (50). These quantities were calculated using the following formulas: AV = 56.1 1‘1. (2) p sv = 56.1(b—a)E (3) p HV = [(b—a)N (56.1) + AV (4) where N is the normality of potassium hydroxide (KOH), p is the amount in grams of wood particles, b is the volume of KOH needed to neutralize the blank (solvent only) and a is the volume of KOH required to neutralize the sample. The factor 56.1 accounts for the molecular weight of KOH (50). 35 AV and SV can provide information about esterification in samples , which is desirable for wood modification studies (13, 23, 26, 27, 29). Hydroxyl group content is also important in analyzing the way samples react (13). If the OH value is reduced after modification, there is evidence that the reaction took place through these groups. Physical and Mechanical Property Testing Wood composite panels are classified according to their density as low, medium or high density products (4, 6). To determine whether formaldehyde-free panels made in . this study would conform to standard strength requirements for conventional particleboard, density of the panels was assessed and mechanical property data was compared to ANSI standard requirements for particleboard (51 ). Density Density of wood and wood-based samples is one of the most important parameters in assessing sample quality (3, 4, 6). Since mechanical property requirements vary with density, the density of experimental panels must be assessed to determine the appropriate range of standard property values for product comparison (51). A simple mass over volume calculation can be used to determine the average density. However, more information about the density profile can be collected using an X-ray density profiler. This instrument sends an x-ray beam through a sample as the beam travels across the sample thickness. Results of this analysis show the variation in density across the sample and can be used to find areas of high and low density in the sample, which may account for mechanical property variations. 36 5F de In. Sill F lexural Properties The flexural strength (modulus of rupture or MOR) and flexural stiffness (modulus of elasticity or MOE) of composites are determined via static bending tests. These tests are usually performed on dry samples, conditioned as per ASTM standards before testing. The MOR and MOE can be calculated from the load-deflection curves using the following equations: MOR=————-—3'P""=”"L (5) 2-b-d2 3 MOE=—-3—L3— (6) where Pmam is maximum load (N), P is the load at proportional limit (N), L is length of span (mm), b is specimen width (mm), d is specimen thickness (mm) and y is center deflection at proportional limit load (mm). Internal Bond (IB) Strength The internal bond (IB) strength measures the tensile strength perpendicular to the surface of the sample and indicates how well the particles are bonded to each other in the center of the panel. It is measured as the maximum load per unit area. In this test, the stress is applied perpendicular to the plane of the adhesive and the energy required to break the adhesive bond is measured. 18 strength is calculated according to Equation 7: 37 IB = —Pmax (7) where Pmax is maximum load (N), b is the width of the specimen (mm) and d is the depth of the specimen (mm). 38 J 14. 10. ll. 12. 13. 14. REFERENCES Haygreen, J.G. and Bowyer, J.L., Forest Products and Wood Science: An Introduction, 3rd Ed., Iowa State University Press: Ames (1996). Sjostrom, E., Wood Chemistry: fundamentals and applications, 2nd Ed., Academic Press: San Diego (1993). Marra, A.A., Technology of Wood Bonding: Principles in Practice. Van Nostrand Reinhold: New York (1992). Forest Products Laboratory, Wood Handbook: Wood as an Engineering Material. Forest Products Society: Madison, WI (1999). Guss, L.M., "Engineered wood products: the future is bright," Forest Products Journal, 45 (7/8): 17-24 (1995). Maloney, T.M., Modern Particleboard and Dry-Process F iberboard Manufacturing, Updated Ed., Miller Freeman: San Francisco (1993). Sellers, T., Jr., "Wood adhesive innovations and applications in North America," Forest Products Journal, 51 (6): 12-22 (2001). Maloney, T.M., "The family of wood composite materials," Forest Products Journal, 46 (2): 19-26 (1996). Pizzi, A., Wood Adhesives: Chemistry and Technology. Marcel Dekker: New York (1983). US-EPA, "National emission standards for hazardous air pollutants: plywood and composite wood products," Federal Register, 69 (146): 45943-46046 (2004). Clemons, C., "Wood-plastic composites in the United States: The interfacing of two industries," Forest Prod. J., 52 (6): 10-18 (2002). Principia Partners, Current and Emerging Applications for Natural and Wood Fiber Composites, presented at 7th International Conference on Woodfibre- Plastic Composites. 2003: Madison, WI. Felix, J .M. and Gatenholm, P., "The nature of adhesion in composites of modified cellulose fibers and polypropylene," Journal of Applied Polymer Science, 42 (3): 609-20 (1991). Matuana, L.M., Balatinecz, J.J., Sodhi, R.N.S., and Park, C.B., "Surface characterization of esterified cellulosic fibers by XPS and FTIR Spectroscopy," Wood Science and Technology, 35(3): 191-201 (2001). 39 ‘1‘) fib- 76. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. Li, Q. and Matuana, L.M., "Surface of cellulosic materials modified with functionalized polyethylene coupling agents," Journal of Applied Polymer Science, 88 (2): 278-286 (2003). Kazayawoko, M., Balatinecz, J.J., and Woodhams, R.T., "Diffuse reflectance Fourier transform infrared spectra of wood fibers treated with maleated polypropylenes," Journal of Applied Polymer Science, 66 (6): 1163-1173 (1997). Kazayawoko, M., Balatinecz, J.J., and Matuana, L.M., "Surface modification and adhesion mechanisms in wood fiber-polypropylene composites," Journal of Materials Science, 34 (24): 6189-6199 (1999). Kazayawoko, M., Balatinecz, J .J ., and Sodhi, R.N.S., "X-ray photoelectron spectroscopy of maleated polypropylene-treated wood fibers in a high-intensity thermokinetic mixer," Wood Science and Technology, 33 (5): 359-372 (1999). Keener, T.J. and Brown, T., "Epolene maleated polyethylene coupling agents," 7’” International Conference on Woodfiber-Plastic Composites, Madison, WI, May 19-20, 2003 : 23-27 (2004). Selke, SE. and Wichman, 1., "Wood fiber/polyolefin composites," Composites, Part A: Applied Science and Manufacturing, 35A (3): 321-326 (2004). Balasuriya, P.W., Ye, L., Mai, Y.W., and Wu, J ., "Mechanical properties of wood flake-polyethylene composites. II. Interface modification," Journal of Applied Polymer Science, 83 (12): 2505-2521 (2002). Matsuda, H., "Thermal plasticization of lignocellulosics for composites," in Emerging Technologies for Materials and Chemicals fiom Biomass R.M. Rowell, T.P. Schultz, and R. Narayan, Editors. American Chemical Society: Washington, DC. p. 98-114 (1992). Matsuda, H., "Preparation and utilization of esterified woods bearing carboxyl groups," Wood Science and Technology, 21 (1): 75-88 (1987). Matsuda, H., Ueda, M., and Mori, H., "Preparation and crosslinking of oligoesterified woods based on maleic anhydride and allyl glycidyl ether," Wood Science and Technology, 22 (1): 21-32 (1988). Clemons, C., Young, RA, and Rowell, R.M., "Moisture sorption properties of composite boards from esterified aspen fiber," Wood and Fiber Science, 24 (3): 353-63 (1992). Timar, M.C., Mihai, M.D., Maher, K., and Irle, M., "Preparation of wood with thermoplastic properties. Part 1. Classical synthesis," Holzforschung, 54 (1): 71- 76 (2000). 40 37 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. Timar, M.C., Maher, K., Irle, M., and Mihai, M.D., "Preparation of wood with thermoplastic properties. Part 2. Simplified technologies," Holzforschung, 54 (1): 77-82 (2000). Timar, M.C., Maher, K., Irle, M., and Mihai, M.D., "Thermal forming of chemically modified wood to make high-performance plastic-like wood composites," Holzforschung, 58 (5): 519-528 (2004). Hon, D.N.S. and Xing, L.M., "Thermoplasticization of wood. Esterification," in Viscoelasticity of Biomaterials, W.G. Glasser and H. Hatakeyama, Editors. American Chemical Society: Washington DC. p. 118-132 (1992). Marcovich, N.E., Reboredo, M.M., and Aranguren, M.I., "Sawdust modification. Maleic anhydride chemical treatment," Holz als Roh- und Werkstojf, 54 (3): 189- 193 (1996). Marcovich, N.E., Aranguren, M.I., and Reboredo, M.M., "Modified woodflour as thermoset fillers Part I. Effect of the chemical modification and percentage of filler on the mechanical properties," Polymer, 42 (2): 815-825 (2001). Marcovich, N.E., Reboredo, M.M., and Aranguren, M.I., "Modified wood flour as thermoset fillers. 11. Thermal degradation of wood flours and composites," Thermochimica Acta, 372 (1-2): 45-57 (2001). Ratner, BB. and Castner, D.G., "Electron Spectroscopy for Chemical Analysis," in Surface Analysis : The Principal Techniques, J.C. Vickerrnan, Editor. John Wiley & Sons: Chichester, England. p. 43-98 (1997). Matuana, L.M. and Kamdem, D.P., "Accelerated ultraviolet weathering of PVC/wood-flour composites," Polymer Engineering and Science, 42 (8): 1657- 1666 (2002). Stark, NM. and Matuana, L.M., "Surface chemistry changes of weathered HDPE/wood-flour composites studied by XPS and FTIR spectroscopy," Polymer Degradation and Stability, 86 (1): 1-9 (2004). Chtourou, H., Riedl, B., and Kokta, B.V., "Surface characterizations of modified polyethylene pulp and wood pulps fibers using XPS and inverse gas chromatography," Journal of Adhesion Science and Technology, 9 (5): 551-74 (1995) Chtourou, H., Riedl, B., Kokta, B.V., Adnot, A., and Kaliaguine, 8., "Synthetic pulp fiber ozonation: an ESCA and FTIR study," Journal of Applied Polymer Science, 49 (2): 361-73 (1993). Silverstein, RM. and Webser, F .X., Spectrometric Identification of Organic Compounds, 6th Ed., John Wiley & Sons, Inc. : New York (1998). 41 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. Pemble, M.E., "Vibrational Spectroscopy from Surfaces," in Surface Analysis: The Principal Techniques, J.C. Vickerman, Editor. John Wiley & Sons: Chichester, England. p. 267-311 (1997). Spectra-Tech, "The complete guide to FT—IR," 12.1-12.6 (1996). Marcovich, N.E., Reboredo, M.M., and Aranguren, M.I., "FTIR spectroscopy applied to wood flour," Composite Interfaces, 4 (3): 119-132 (1996). Stark, N.M., Matuana, L.M., and Clemons, C.M., "Effect of processing method on surface and weathering characteristics of wood-flour/HDPE composites," Journal of Applied Polymer Science, 93 (3): 1021-1030 (2004). Stark, NM. and Matuana, L.M., "Surface chemistry and mechanical property changes of wood-flour/high-density-polyethylene composites after accelerated weathering," Journal of Applied Polymer Science, 94 (6): 2263-2273 (2004). Matuana, L.M., Kamdem, D.P., and Zhang, J., "Photoaging and stabilization of rigid PVC/wood fiber composites," Journal of Applied Polymer Science, 80 (11): 1943-1950 (2001). Maunu, S.L., "NMR studies of wood and wood products," Progress in Nuclear Magnetic Resonance Spectroscopy, 40 (2): 151 -174 (2002). Hatfield, G.R., Sardashti, M., and Maciel, G.E., "Analysis of molecular orientational order in solid samples by nuclear magnetic resonance: application to lignin and cellulose in wood," Analytical Chemistry, 59 (13): 1659-64 (1987). Jandura, P., Kokta, B.V., and Riedl, B., "Fibrous Long-Chain Organic Acid Cellulose Esters and Their Characterization by Diffuse Reflectance FTIR Spectroscopy, Solid-State CP/MAS 13C-NMR, and X-Ray Diffraction," Journal of Applied Polymer Science, 78: 1354-1365 (2000). Solum, M.S., Pugmire, R.J., Jagtoyen, M., and Derbyshire, F., "Evolution of carbon structure in chemically activated wood," Carbon, 33 (9): 1247-1254 (1995) Bettini, S.H.P. and Agnelli, J.A.M., "Grafting of maleic anhydride onto polypropylene by reactive extrusion," Journal of Applied Polymer Science, 85 (13): 2706-2717 (2002). Anonymous, "Recommended methods for the analysis of alkyd resins," Pure and Applied Chemistry, 33 (2-3): 411-435 (1973). ANSI, A208.I-I999. Particleboard. The Composite Panel Association: Gaithersburg (1 999). 42 CHAPTER 3 COMPOSITE MATERIALS MANUFACTURED FROM WOOD PARTICLES MODIFIED THROUGH A REACTIVE EXTRUSION PROCESS This chapter is slightly modified from Polymer Composites, published in August 2005. 26 (4): 534-541. It is co-authored by K. Carlbom and L.M. Matuana. 43 ABSTRACT Wood-based composites such as particleboard and medium density fiberboard are currently made with formaldehyde-containing adhesives. Since the government is continuously developing and implementing very stringent regulations to eliminate formaldehyde emissions into the environment, alternative approaches must be developed to replace these adhesives. This study examined the concept of using a reactive extrusion process as a means of developing a new, formaldehyde-free binding system for wood composite products. The surfaces of wood particles were modified by grafting maleated polyolefins through a continuous reactive extrusion process. Chemical changes resulting from this treatment were followed by studying the FTIR, '3 C NMR and XPS spectra. The modified wood particles were compression-molded into panels, which were tested for mechanical properties. FTIR, 13C NMR and XPS data revealed that chemical reactions had taken place between the hydroxyl groups of wood particles and maleated polyolefins. The mechanical property test results indicated that the composite panels compared favorably with current standard requirements for conventional particleboard. 44 INTRODUCTION Wood-based composites are commonly made using formaldehyde-based adhesives, including urea-formaldehyde, melamine-formaldehyde, and phenol- formaldehyde (1-3). In 1998, 1,780 kilotons of adhesive resin solids were used to produce primary glued wood products (excluding the adhesive used to bond furniture and other secondary wood products). Of this amount, nearly 92% were formaldehyde-based adhesives (3). Plants that produce wood composites using formaldehyde-based adhesives emit harmful chemicals to the environment. These include phenol, formaldehyde, ketones, and other compounds, which are known hazardous air pollutants (HAPs) (4). Common composite products such as plywood, oriented strandboard and particleboard are used in building construction and in furniture, which is a concern as these products also tend to release formaldehyde over time (5). Wood composites made today emit far less formaldehyde than those made 20 years ago, but the problem has not been eliminated. Formaldehyde and other toxic compounds may be present in large amounts in both indoor and outdoor air as a consequence of the use of these adhesives (6, 7). As a result of public concern about the environment, the Environmental Protection Agency (EPA) proposed new rules for facilities that manufacture plywood and composite wood products in August of 2002 (6). If adopted, these rules would affect both new and existing plants that produce at least 10 tons of any one HAP per year, or any combination of 25 tons of HAPs per year. The California Air Resources Board has gone even further; proposing a regulation that would eliminate urea-formaldehyde based wood composites from being sold in California, regardless of where they were made (7). 45 or. M 1hr 301' up 5X11) ll ()0 F00. SPuL These regulations will force industry to find new ways to bind composite products without the use of formaldehyde-based adhesives. In recent years, there have been several studies into environmentally friendly wood adhesives (3). Some of the areas that have been investigated include urea- formaldehyde adhesives with low forrnaldehyde-to-urea molar ratios, and the development of tannin, lignin, soybean and cornstarch adhesives (3, 8), and phenol- forrnaldehyde resins modified with lignin (9, 10). Additives that reduce formaldehyde release during composite pressing and during board use have also been developed (8). The regulation of formaldehyde emissions has also lead to some development of fiberboard without synthetic adhesives (11, 12). The binderless boards and those made with natural adhesives tend to have poorer properties than those made with synthetic adhesives. The approach of this work is to graft maleated polyolefins to wood particles in order to bond the wood particles together without the use of additional adhesive. Prior A work demonstrated the ability to graft maleated polyolefins to cellulosic materials through a wet process (13). However, the wet process had the drawback of using organic solvents, which had to be removed through drying. Therefore, the wet process is both expensive and time consuming on an industrial scale. The main objective of this study was to study the concept of using a reactive extrusion process as a means of developing a new, formaldehyde-free binding system for wood composite products. The effectiveness of the modification was followed by Fourier transform infrared (FTIR) spectroscopy, nuclear magnetic resonance (NMR) spectroscopy and X-ray photoelectron spectroscopy (XPS). FTIR is useful in 46 determining the presence of functional groups. NMR can determine bonding and chemical connectivity, while XPS can reveal the elemental composition on the surface of materials (13-15). Panels were pressed from the modified wood particles and mechanical properties of the resulting panels were tested and compared with current standard requirements for conventional particleboard (16). 47 EXPERIMENTAL Materials Maple wood particles of 425 micron (40-mesh) and 150 micron (100-mesh) size were supplied by American Wood Fibers (Schofield, WI) and were used as particles. The 150 micron particles were used for the analytical work because the diffuse reflectance IR technique required very small particles to minimize the effects of scattering and specular reflectance in the samples. However, these small particles were difficult to feed into the extruder. Since panel manufacturing required a large quantity of modified particles, larger (425 micron) particles, which were easier to process, were used in panel manufacturing and mechanical property testing. Hydrated zinc acetate, the catalyst, and xylene (99.9%, ACS Grade), the solvent used for Soxhlet extraction, were obtained from Baker Analytical Reagents (JT Baker Co., Phillipsburg, NJ). Maleated polyethylene (G- 2608 or MAPE) and maleated polypropylene (G-3003 or MAPP) supplied by Eastman Chemical Co. (Kingsport, TN) were used as the binding agents. Table 3.1 lists the characteristics of these binding agents. The wood particles were dried for 48 hours at 105°C to a final moisture content of less than one percent before processing. All other chemicals were used as received. 48 Table 3.]. Characteristics of Maleated Polyolefins Properties Maleated polyethylene Maleated polypropylene (MAPE or G-2608) (MAPP or G-3003) Maleic anhydride content 1.5% 1.5% by weight Melting point 122°C 156°C Weight average molecular weight (Mw) 51,700 g/mol 52,000 g/mol Melt flow indexI 8 g/ 10 min 12.7 g/ 10 min lMelt flow index measured at 190°C and 2.16 kg according to ASTM D1238. 49 Reactive Extrusion of Wood Particles A 10-liter high intensity mixer (Papenmeier TGAHKZO) was used for dry blending of the wood particles, binding agent, and catalyst. All components were combined in the mixer and blended for 10 minutes at room temperature. Amounts of all components used in the formulation are summarized in Table 3.2. The mixture was then fed into a 32 mm conical counter rotating twin-screw extruder (C. W. Brabender Instruments, Inc.) with a L/D ratio of 13:1, driven by a 7.5 hp Intelli-Torque Plasti- Corder Torque Rheometer®. Based on preliminary work, the barrel temperatures for the three zones inside the extruder were set at 160°C for maleated polyethylene and 165°C for maleated polypropylene, and the rotational speed of the screws was held at 60 rpm. No die was used to extrude these particles. 50 Table 3.2. Formulation Used for Surface Modification of Wood Particles % Total in Ingredients Composite Weight (g) Wood particles 79% 790 Maleated polyolefins 20% 200 Hydrated zinc acetate 0 (catalyst) 1 /° 10 51 E1 5111 M; U111 00 sec hm SEC Su 1(15 (Ni 63C. SUI): and 11)) Extraction of Wood Particles Modified and unmodified (unextruded) wood particles used for spectroscopic studies were Soxhlet extracted with xylene following the approach described by Li and Matuana (13). Particles were extracted for 24 hours after modification to remove any unreacted binding agent, oven-dried at 105°C until constant weight was achieved, and were then analyzed by FTIR. A second 24-hour Soxhlet extraction was then performed to make sure the removal of unreacted binding agent was complete from the surface of wood particles upon the first extraction. The infrared spectra of wood particles after this second extraction were collected for comparison with those collected after the first 24- hour extraction. All NMR and XPS analyses were performed on the particles after the second extraction. Surface Characterization of Wood Particles Unmodified and modified wood particles were dried to a constant weight at 105°C and analyzed by infrared spectrophotometry, using a Nicolet Protége’ 460 FTIR (N icolet Instrument Co., Madison, WI). Spectra were recorded in Kubelka-Munk units in the range of 4000-400 cm", with a resolution of 4 cm'1 and a coaddition of 128 scans for each spectrum. Pure powdered potassium bromide (KBr) was used as a reference substance while no dilution of powdered-wood particles in KBr was required to obtain a spectrum. Diffuse reflectance was used with the FTIR for transfer of infrared radiation and data analysis was performed using WinFIRST software (Thermo-Nicolet, Madison, WI). 52 obi. 0n 2 0pm prol amb optii ICCyi wig and surf; on C1111 The Pan PICS: large Press It’mpc Solid-state carbon 13 nuclear magnetic resonance (”C NMR) spectra were obtained using a 1H-13 C Cross-Polarization Magic Angle Spinning (CP-MAS) experiment on a Varian-Chemagnetics (Varian Inc., Palo Alto, CA) 400 MHz NMR Spectrometer operating at 100.529 MHz. The spectrometer was equipped with a double resonance probe. Samples were spun in a 6 mm rotor at 4 kHz, at the magic angle (54.7°) and at ambient temperature. Pulse widths, contact time and pulse amplitudes for the CP were optimized on an adamantane standard. The contact time in all cases was 1 ms. The recycle delay (pulse delay) was 2.0 s. Spectra were processed using exponential weighting (100 Hz). X-ray photoelectron spectroscopy (XPS) was used to determine the concentration and types of carbon atoms, as well as the oxygen-to-carbon atomic ratios present on the surface of the wood particles before and after modification. XPS analysis was carried out on a Physical Electronics Phi 5400 ESCA System, (Physical Electronics USA, Chanhassen, MN) using a non-monochromatic Mg source and a takeoff angle of 45°. The procedure for XPS data collection and analysis was detailed in other articles (13, 14). Panel Manufacturing and Mechanical Property Testing Panels were prepared from modified 425 micron wood particles using a hydraulic press from Eric Mill Co. (Erie, PA). Panel dimensions were 380 by 380 by 6 mm, with a target density of 720 kg/m3. Panels were pressed at 193°C for 7 minutes using 8 MPa of pressure. After pressing, panels were removed from the hot press and cooled at room temperature under compression for 15 minutes. 53 test test: D1 I of n with Stat dete b011- Ex; Three-point flexural tests and internal bond (IB) strength tests were performed on an Instron 4206 testing machine (using Series IX software). For three-point flexural tests, the crosshead speed was 3.05 mm/min, while the crosshead speed for 18 strength tests was 8.13 mm/min in accordance with the procedure outlined in ASTM standard D1037-99 (17). At least six samples were tested to obtain an average value for modulus of rupture (MOR), modulus of elasticity (MOE), and 1B strength, which were compared with the values listed in the standard ANSI A208 . l -1 999 Particleboard (16). Statistical Analysis A two-sample t-test was carried out with an a significance value of 0.05 to determine the effect of binding agent type (MAPE vs. MAPP) on the flexural and internal bond properties of the composites. Statistical analysis was performed using Design Expert software (Version 6) from Stat-Ease. Inc. Minneapolis, MN. 54 50 met thr0 char will RESULTS AND DISCUSSION Surface Characterization of Wood Particles Modification of the wood particles was expected to take place through the mechanisms proposed in Figure 3.1, where a monoester or a diester could be formed through reaction of the maleated polyolefin with hydroxyl groups on the wood surface (13-15). Various surface characterization techniques were then used to document changes in the chemistry of the unmodified wood particles and wood particles modified with maleated polyolefins through a reactive extrusion process. 55 Figt O fill—OH 0%?‘2 I “ (R). /' OH CH [— o>c \ / i‘R' (R)n Wood particle Maleated polyolefin *CH2_ CH2_)'_ Where R = or CH3 Figure 3.1. _ CH3 l 8 /|—O—C—CH2 3‘01'1 />3H /(Rln | HO’fi \C-R' Monoester o l (Rln | 8 /j—O—C—CH2 / I /l—O—E—CH (Rln /' 0 \C/—R' Di-ester Modified wood (Rln particle Modification scheme for esterification reaction between wood particles and maleated polyolefins: (a) mono-ester and (b) di-ester formation. 56 1118 sh0 sh0‘ strel \‘ibr £11101 C0111 The FTIR spectra of unmodified and modified wood particles, along with the maleated polyethylene are shown in Figure 3.2. Similarly, the FTIR spectra of unmodified and modified wood particles, along with the maleated polypropylene are shown in Figure 3.3. The unmodified wood particles (spectrum A in Figures 3.2 and 3.3) showed an absorbance band at 3400 cm", which was attributed to hydroxyl group stretching vibrations, and another at 2900 cm", which was associated with C-H stretching vibrations. A band near 1740 cm'1 was associated with C=O stretching vibrations, and another at 1122 cm“1 was likely due to C-0 stretching vibrations and C-C stretching from components of cellulose (13, 15). 57 rang-mx—oacx inn—3C Figure 3.2. N 3% N .9: D NN 8‘53 N N E? E 4000 3500 3000 2500 2000 1500 1000 500 Wavenumbers (cmA-1) FTIR spectra of unmodified wood particles (A), maleated polyethylene- MAPE (B), MAPE-modified wood particles without extraction (C), with 24 hour-Soxhlet extraction (D), and after a second 24 hour-Soxhlet extraction (E) in the region 4000-400 cm". 58 rang-mx—oacx mnmzc 4000 3500 3000 2500 2000 1500 1000 500 Wavenumbers (cmA-1) Figure 3.3. F TIR spectra of unmodified wood particles (A), maleated polypropylene- MAPP (B), MAPP-modified wood particles without extraction (C), with 24 hour-Soxhlet extraction (D), and after a second 24 hour-Soxhlet extraction (E) in the region 4000-400 cm". 59 Fig Silt 1168 171 and and m be ex P3. EhSG The spectra of maleated polyethylene and maleated polypropylene (spectrum B in Figures 3.2 and 3.3, respectively) showed four distinct absorption bands. The bands between 2960 cm'1 and 2840 cm'1 were due to symmetrical and asymmetrical C-H stretching vibrations of CH2 and CH3 in the polyolefin chain (13, 15). The small band near 1786 cm'l was attributed to anhydride C=O stretching. Bands found between 1720- 1710 cm'I arose from C=O stretching vibrations in the maleated polyolefins and the absorbance bands from 1463 to 1300 cm'1 were from C-H deformation vibrations of CH2 and CH3. The bands below 1250 cm'1 were associated with rocking vibrations of CH2 and CH3, or C-C stretching vibrations from the polyolefin chain (13, 15). The infrared spectra of wood modified with maleated polyolefins are also illustrated in Figures 3.2 and 3.3 (spectra C-E). For the interpretation of infrared spectra, it is important that all the unreacted maleated polyolefins are removed in the 24 hour Soxhlet extraction with solvent. It is seen from these figures that there was very little difference in the spectra of the modified wood particles after the first (spectrum D) and second extractions (spectrum E). This finding provides important evidence that the maleated polyolefins were chemically bonded to the wood particles. If the maleated polyolefins had not grafted to the particles, a decrease in peak intensity or a loss of the band with distinct peaks between 2960 and 2840 cm'1 would have been expected after the extraction, which would remove any polyolefins not chemically bonded to the wood particles (13). The modified wood particles (spectra C-E in Figures 3.2 and 3.3) also showed absorption bands at 3400 cm", but the intensity of this peak decreased compared to the unmodified wood particles, indicating that there were less OH groups on the surface of 60 [1101‘ 11111 6511 163 5111' rep the rib: incr rear in c imci 1680 0:110 llj’dri surfai SPCCII lSpeQi Who did 11 modified samples. This was expected based on the reaction scheme shown in Figure 1, where the maleated polyolefin reacted with the OH groups of wood particles forming an ester link (13). The reaction may form either a single ester link or a diester where both reactive groups on the maleated polyolefin bond with hydroxyl groups on the wood surface. A distinct change was found near 2900 cm'l, where a large band between 2960 and 2840 cm], having distinct peaks similar in appearance to the maleated polyolefin, replaced the single peak in the unmodified wood particle spectrum. This feature is characteristic of the maleated polyolefins (spectrum B), and is due to C-H stretching vibrations (13, 15). Another indication of grafting of the maleated polyolefin was an increased intensity in the band at around 1740 cm", possibly due to esterification reaction. Although not clearly seen in Figures 3.2 and 3.3 due to stacking of five spectra in one figure, the individual spectra of modified wood particles showed an increased intensity in the peaks at 1462 and 1376 cm'1 which is also indicative of the grafting reaction, suggesting more C-H character in the modified samples (13, 15). It should also be mentioned that the intensity of the band at 1122 cm'I decreased, likely due to less cellulose being detected on the surface because of the grafting of the long aliphatic hydrocarbon chain of maleated polyolefin (13-15). The evidence supporting the chemical bonding of maleated polyolefins to the surface of wood particles obtained by infrared was further supplemented by NMR spectra, as illustrated in Figure 3.4. The solid state '3 C NMR maleated polyethylene (spectrum A) showed only one distinct peak at 30 ppm, which was due to CH2 chain carbons of the maleated polyethylene (18). Oxidized carbons from the maleated group did not appear, likely due to their low concentration in the maleated polyethylene 61 C011 “’01 1‘1)n 3101 (1‘1 mm was wet spec compounds. By contrast, spectra of both unmodified wood (spectrum B) and modified wood particles (spectrum C) showed several characteristic peaks. The carbon peak at 152 ppm is attributed to lignin compounds in the wood. Bands from different types of carbon atoms in cellulose appeared at 106 and 73 ppm. Wood particles also showed a small peak near 55 ppm attributed to methoxy groups found in wood lignin and hemicelluloses (19). As expected, a large increase in the peak intensity at 30 ppm was observed in the modified wood particles (spectrum C), clearly indicating that the maleated polyethylene was chemically bonded to the wood particles. Any unbound maleated polyethylene would have been removed by the Soxhlet extraction process, and would not appear in the spectrum. 62 Figu 3O 55 152 B _ __ 73 106 c —t 300 250 200 150 100 50 0 PPM Figure 3.4. Solid-state 13C NMR spectra of (A) MAPE, (B) unmodified wood particles, and (C) wood particles modified with MAPE. 63 FTIR signi male signi \l'lllC surfa 511113 0.47 pols was ‘ The XPS data summarized in Table 3.3 also confirmed the findings of the FTIR analysis. As expected from compounds rich in carbon-containing groups, a significant increase in C1 was observed after surface modification of wood particles with maleated polyolefins. In addition, the content of oxidized carbon atoms (C2-C4) significantly decreased in the modified wood particles, along with the O/C atomic ratio, which was also expected due to the large increase in aliphatic carbon atoms on the surface of the modified wood particles (13, 14). The O/C atomic ratios showed that the surface of the wood particles changed dramatically with modification, decreasing from 0.47 to 0.03 with maleated polyethylene and from 0.47 to 0.06 with maleated polypropylene. Changes this large likely indicate that the reactive extrusion procedure was very successful in grafting maleated polyolefins to the wood particles. 64 Am: AOHUAVV 938w 38083-80: m was 388.88 a 2 B8: 08 £023 V803 conned 80¢ $88 cab 8088 v0 88 AOHU .O-U-Ov 89m nowxxo 3:088 Bwfim a 8 do .0808 cowxxo 78383 -80: 95 8 causes m88m 82:8 80¢ momma MU .AEOOV samba 38588 a 82: 550 .803 cowtmxo Bwfim a 8 Becca .58on 8588 80¢ m0 AIOOOV m88m come??— cohoce 8580 9 Eco wagon m89e 85:3 89a memes _ D 88388 :onBUN 823868 628m s cm @883 a Home 38890 39: @803 com: 8383 too? uofiuoE van 3588:? . . . . . . . . 2:22 8 o 2 ea 3 m o o g N 2 e so as as, BEBE 83:8 895 8d See mam ed 33 EN whee 6.23: 828283; Base: . . . . . . . $22 8 o 8 3 8 N o o mm o S e we we as) eeceee 3.8.8 895 :.o 0.8 2: o.o am I Em GEE 822338 Base: . . . . . . . 8825.23 5. o 2 we a x mm o 8 m cm a me a 83:8 noes eeeeeea: 8:2 o o 3 8 8 6 383a Ae\ev 8352—83 Viv—3:2 US .3585 ”as; 9.8.. .6 he were: max 43 8880on $35.82* woo? mo 80989800 888m 380805 use 33m 20 cots—cmomhwi .m.m 2an 65 )lecl bind" as H belu 11001 expe as 51 woul Ipor path 311) at Mechanical Properties Surface analysis techniques provided evidence that the maleated groups of the binding agent had chemically bonded with hydroxyl groups on the wood particle surface, as illustrated in Figure 3.1. This process resulted in the formation of an ester linkage between the wood surface and the maleated group of the binding agent, which produced a wood particle with a pendant polyolefin chain. Bonding between wood particles was expected to occur through diffusion and entanglement of the pendant polyolefin chains, as shown in Figure 3.5. During compression molding, the pendant polyolefin chains would melt and flow under the influence of heat and pressure, causing them to entangle. Upon cooling, the entangled polyolefin chains would be locked together, forming a direct particle-to-particle bond between the modified wood particles, without the addition of any adhesive. 66 Figiu 0 I I (II ll /j—O—C—CH2 Hzc—C—O—‘I\ él 0H CH /(R)n (Rln\ CH H0 I? / ' HO’fi \C—R' + R—C/ I‘OH ' O l l o (R)n (Rln Modified wood particle 2 Modified wood particle i Panel Figure 3.5. Panel manufacture scheme for modified wood particles where R is an 67 ethylene or propylene repeat unit and R’ is hydrogen or a methyl group. 5116 C011 6% C311 C011 0111C To determine whether panels made by this process would conform to standard strength requirements for conventional particleboard, mechanical property data was compared to ANSI standard requirements for particleboard. Table 3.4 lists the MOR, MOE, and 18 strength requirements for particleboard of medium density, ranging from 640-800 kg/m3. There are four grades of particleboard of medium density, all of which can be made with either interior or exterior adhesives. Grades M-1 and M-S are commercial grade boards, while M-2 and M-3 are intended for industrial use. Panels for outdoor use must also be labeled exterior, according to the ANSI standard for particleboard (16). The composite panels manufactured in this study were within this range, with an average density value of 778 i 12 kg/m3 for panels made with maleated polyethylene and 775 i 8 kg/m3 for panels made with maleated polypropylene. 68 Table SIC 111 11 l Fion‘ 3 'Ihei denou Chhcr Table 3.4. Requirements for Various Grades of Particleboard of Medium Density (640-800 kg/m3) Gradesl Experimental Values2 Properties M-l M-S M-2 M-3 Wood/MAPE Wood/MAPP MOR (N/mmz) 11.0 12.5 14.5 16.5 20.7 2: 3.4 A 22.9 i 4.3 A MOE (N/mmz) 1725 1900 2250 2750 1296 i 195 A 2870 i 320 B [B (N/mmz) 0.40 0.40 0.45 0.55 2.06 i 0.7 A 1.50 2: 0.3 A lFrom Standard ANSI A208.1-1999 Particleboard (16). 2The capitalized letters represent the ANOVA results. The means with the same letter denote that the difference between these two treatments is not statistically significant. Otherwise, the difference is statistically significant at values of or = 0.05. 69 of pa been signi 1101.“ 513110 bindi T1101 iOb‘i 101.11 Mi 10111 [381111 form; 6010 ANS Regardless of the maleated polyolefin type, the MOR and IB strength results of panels manufactured in this study indicated that the standard requirements have been met and surpassed for all grades of particleboard of medium density. No significant difference in both MOR and 18 was observed between maleated polyethylene and maleated polypropylene. However, the MOE data are below the standard requirements for stiffness when maleated polyethylene was used as a binding agent for wood particles. By contrast, the experimental panels exceeded the MOE requirements for all grades of particleboard of medium density when maleated polypropylene was used as binding agent. Panels bonded with maleated polypropylene were statistically superior in MOE compared to the maleated polyethylene counterparts. This difference is due to the higher stiffness of polypropylene when compared to polyethylene. These results are significant because particleboard is currently manufactured with formaldehyde-based adhesives. The formaldehyde-free wood composites manufactured in this study are more environmentally friendly and often outperform the requirements listed in the standard ANSI A208.]. 70 p0ly0? based wood could withm emim wood perior requiic CONCLUSIONS This study examined the possibility of modifying wood particles with maleated polyolefins in a reactive extrusion procedure in order to make formaldehyde-free wood— based composite panels. FTIR, 13 C NMR and XPS results verified the reaction between wood particles and maleated polyolefins. This proved that the maleated polyolefins could be successfully grafted to wood particles using a reactive extrusion process, without the use of any solvents. The study also showed that a new type of environmentally friendly wood composite product could be formed from the modified wood particles. This composite contained no formaldehyde-based adhesive, but still performed very well in mechanical tests, in some cases exceeding the standard requirements for particleboard of medium density. 71 Ix.) SJ) 10. 11. 12. REFERENCES Maloney, T.M., "The family of wood composite materials," Forest Products Journal, 46 (2): 19-26 (1996). Guss, L.M., "Engineered wood products: the future is bright," Forest Products Journal, 45 (7/8): 17-24 (1995). Sellers, T., Jr., "Growing markets for engineered products spurs research," Wood Technology, May/June: 40-43 (2000). Barry, A., Comeau, D., and Lovell, R., "Press volatile organic compound emissions as a function of wood particleboard processing parameters," Forest Products Journal, 50 (10): 35-42 (2000). Maloney, T.M., Modern Particleboard and Dry-Process F iberboard Manufacturing, Updated Ed., Miller Freeman: San Francisco (1993). Anonymous, “Summary of working draft of proposed rule for plywood and composite wood products,” National Emission Standards for Hazardous Air Pollutants (N ESHAP), Rule Development Project Lead: Greg Nizich ( nizichgrgg@epa. gov), US. EPA, Technology Transfer Network-Air Toxics Website, August 2002, http://wvwv.epagov/ttn/atw/plypart/plypart.html Anonymous, “Fact Sheet: Composite Wood Products,” California Air Resources Board Website, March 2003, htth/wwwarbca. gov Pizzi, A., Wood Adhesives .' Chemistry and Technology. Marcel Dekker: New York (1983). Matuana, L.M., Riedl, B., and Barry, A.O., "Kinetic characterization by DTA of lignosulfate-based phenol-formaldehyde resins," European Polymer Journal, 29 (4): 483-90 (1993). Kazayawoko, J .S.M., Riedl, 13., Poliquin, J ., Barry, A.O., and Matuana, L.M., "A lignin-phenol-formaldehyde binder for particleboard. Part 1. Thermal characteristics," Holzforschung, 46 (3): 257-62 (1992). Velasquez, J .A., Ferrando, F., Farriol, X., and Salvado, J ., "Binderless fiberboard from steam exploded Miscanthus Sinensis," Wood Science and Technology, 37 (3-4): 269-278 (2003). Widsten, P., Qvintus-Leino, P., Tuominen, S., and Laine, J .E., "Manufacture of fiberboard from wood fibers activated with F enton's reagent (H202/FeSO4)," Holzforschung, 57 (4): 447-452 (2003). 72 13. 14. 15. 16. 17. 18. 19. Li, Q. and Matuana, L.M., "Surface of cellulosic materials modified with functionalized polyethylene coupling agents," Journal of Applied Polymer Science, 88 (2): 278-286 (2003). Matuana, L.M., Balatinecz, J.J., Sodhi, R.N.S., and Park, C.B., "Surface characterization of esterified cellulosic fibers by XPS and FTIR Spectroscopy," Wood Science and Technology, 35 (3): 191-201 (2001). Kazayawoko, M., Balatinecz, J .J ., and Woodhams, R.T., "Diffuse reflectance Fourier transform infrared spectra of wood fibers treated with maleated polypropylenes," Journal of Applied Polymer Science, 66 (6): 1163-1173 (1997). ANSI, A208.1-1999, Particleboard. The Composite Panel Association: Gaithersburg (1999). ASTM, D 103 7-99, Standard Methods of Evaluating the Properties of Wood- Based Fiber and Particle Panel Materials. ASTM: West Conshohocken (1999). Heinen, W., Rosenmoller, C.H., Wenzel, C.B., de Groot, H.J.M., Lugtenburg, J ., and van Duin, M., "13C NMR study of the grafting of maleic anhydride onto polyethylene, polypropene and ethene-propene copolymers," Macromolecules, 29: 1151-1157 (1996). Solum, M.S., Pugmire, R.J., Jagtoyen, M., and Derbyshire, F., "Evolution of carbon structure in chemically activated wood," Carbon, 33 (9): 1247-1254 (1995) 73 FI‘ This auth CHAPTER 4 F UNCT IONALIZATION 0F WOOD PARTICLES THROUGH A REACTIVE EXTRUSION PROCESS This chapter is in press for the Journal of Applied Polymer Science (2006). It is co- authored by K. Carlbom and L.M. Matuana. 74 P01)" MA] durii were parti dil‘fe XPS ester 0f 1r. ll}'dri poly deter Signi cond no d extru male studi ABSTRACT Wood particles were modified in a reactive extrusion process with maleated polyethylene (MAPE) and maleated polypropylene (MAPP) compounds. Contents of MAPE were varied to study the effect of material composition on grafting efficiency during reactive extrusion, while extruder barrel temperatures and rotational screw speeds were varied to evaluate the effects of processing conditions on the modification of wood particles. Polymer molecular weight effects were investigated using MAPP with different molecular weights. Efficiency of the modification was assessed using FTIR and XPS surface analysis techniques, along with a titrimetric analysis to verify the esterification reaction between the wood particles and maleated polyolefins. The grafting of maleated polyolefins onto the surface of the wood particles through a reaction of hydroxyl groups on the wood surface with the maleated groups of the maleated polyolefins was confirmed, while the level of grafting of MAPE onto wood particles was determined to be a function of the MAPE concentration. However, there was no significant difference found in grafting efficiency at different extrusion processing conditions; rather all of the conditions resulted in adequate grafting. Similarly, there was no difference in grafting efficiency with the molecular weight of MAPP. Reactive extrusion was found to be a suitable technique for the modification of wood particles with maleated polyolefins for all of the material compositions and processing conditions studied. 75 fhnn pohi pn0r modi smne ihnn. noth High hydn thegi SeVei (0r0 fiber 05s 50h fille. lime Duh 00m INTRODUCTION In response to the need for new adhesives for wood composite products, a formaldehyde-free binding system for wood particles which used only maleated polyolefins to create direct wood particle to wood particle bonding was developed in our prior work (1, 2). Wood composite panels were manufactured from maleated polyolefin- modified wood particles, without using any additional adhesive. These panels met and in some cases even exceeded standard requirements for particleboard made with formaldehyde-based adhesive. However, the mechanism for the adhesive bonding was not firlly developed. Maleated polyolefins react with wood particles through an esterification reaction (Figure 4.1). This reaction occurs between the maleated groups of the polyolefins and hydroxyl groups on the wood surface, forming a monoester or a diester. This results in the grafting of a pendant polyolefin chain to the wood surface through the ester linkage. Several authors have studied this reaction with maleated compounds and wood particles (or other cellulosic materials) (3-14). However, most of this previous work focused on fibers modified through a solvent-based process where the maleated compounds were dissolved in heated xylene, toluene, N,N-dimethylformamide (DMF) or other organic solvents and cellulosic particles were added (3-10, 12-14). Cellulosic fibers have to be filtered and dried after modification, which makes this process both cost-ineffective and time consuming, and the often harmful organic solvents must be properly disposed of. In order to produce modified wood particles that could be used to make alternative composite panels, it is necessary to change from a solvent-based process to a dry process. 76 o I /I_OH %CH2 A! OH + (3)/0'11 (R)" f— 0 \ / C-R' (Rln Wood particle Maleated polyolefin —(CH2— CH2-)— Where R = or ——(CH2— l confl conu inoz uas< hadti speec thee‘ toth each zone {roe Surface Modification of Wood Particles with Maleated Polyolefins in Reactive Extrusion A 10-1iter high intensity TGAHK20 mixer (Papenmeier, Germany) was used for dry-blending of the wood particles, binding agent, and catalyst. All ingredients were combined in the mixer and blended for 10 minutes at room temperature. Amounts of all components used in the formulation are listed in Table 4.2. These mixtures were then fed into a 32 mm conical counter-rotating twin-screw extruder (C. W. Brabender Instruments, Inc., South Hackensack, NJ) with a length to diameter (L/D) ratio of 13:1. The extruder ®. The extruder was driven by a 5.6 kW Intelli-Torque Plasti-Corder Torque Rheometer had three temperature zones on the barrel (Figure 4.2) and an adjustable rotational screw speed. Various processing temperatures and rotational screw speeds were used to study the effects of processing conditions on the effectiveness of grafting maleated polyolefins to the surface of the wood particles. In all cases, a uniform barrel temperature was set for each test. For example, when the desired processing temperature was 130°C, all three zones were set at that temperature. This was done similarly for the other studied processing temperatures of 140, 160 and 180°C. 82 Table Bind Table 4.2. Formulations Used for Modified Wood Particles Ingredients in the formulations Binder content (%) Maple(particles MAPE or MAPP Hytil’caettidezinc g) (g) (g) 5 94 5 1 10 89 10 1 15 84 15 l 20 79 20 1 Hopper Motor Gear 1 Zone 1 Zone 2 Zone 3 Figure 4.2. Diagram of the extruder showing the three heating zones. 83 Inoc ban Hnn [em] Vafit Stud: COIIC Spec Exn rent Wei The effect of material composition on the grafting of MAPE onto wood particles in a reactive extrusion process was investigated by holding the extruder barrel temperature and rotational screw speed constant at 160°C and 60 rpm, respectively. Whereas, the MAPE contents were varied from 5 to 20% of the total batch weight (Table 4.2). To determine the effects of processing conditions on grafting efficiency of modified wood particles, the MAPE concentration was fixed at 20% and the extruder barrel temperatures and rotational screw speeds were varied. Barrel temperatures were limited to the range above the melting point of MAPE and below the degradation temperature of wood (130, 140, 160 and 180°C). The rotational speed of the screws was varied from 20 to 80 rpm in 20-rpm increments at each barrel temperature profile. The influence of molecular weight on the efficiency of the grafting reaction was studied using four different MAPP compounds (Table 4.1). Extrusion processing conditions were held constant with the barrel temperature at 160°C and the rotational speed of the screws at 60 rpm. Extraction of Wood Particles Unmodified wood particles were Soxhlet extracted with acetone for 24 hours to remove impurities, air dried for 60 hours, and then oven dried at 105°C to a constant weight before using for further analysis. Modified wood particles were also Soxhlet extracted with xylene for 24 hours to remove any unreacted maleated polyolefins which had not been linked to the wood particles, followed by air drying for 60 hours, and oven 84 dryii by i first dne Sur usin dete mod cm' spec Pure of t perf mm “01 drying to a constant weight at 105°C. These modified wood particles were then analyzed by FTIR. A second 24-hour Soxhlet extraction was performed to ensure the removal of unreacted maleated polyolefin was complete from the surface of wood particles upon the first extraction. Following the second extraction, modified wood particles were again air dried for 60 hours and then dried to a constant weight at 105°C, after which FTIR, XPS and titrimetric analyses were performed. Surface Characterization of Wood Particles FTIR spectra of both unmodified and modified wood particles were obtained using a Nicolet Prote’gé 460 FTIR spectrometer (N icolet Instrument Co., Madison, WI) to determine the functional groups present at the surface of the samples before and after modification. Spectra were recorded in Kubelka-Munk units in the range of 4000-400 cm", with a resolution of 4 cm'1 and a coaddition of 128 scans for each spectrum. All spectra were collected using diffuse reflectance with FTIR for transfer of IR radiation. Pure powdered potassium bromide (KBr) was used as a reference substance. No dilution of the wood particles in KBr was required to obtain the spectra. Data analysis was performed using WinFIRST sofiware from Thermo Nicolet (Madison, WI). No baseline modification was done before performing data analysis of the spectra. Based on prior work (1), the regions of interest in the FTIR spectra of modified wood particles with maleated polyolefins were the absorbance bands near 2900 cm'1 and 1740 cm“, for CH stretching of aliphatic carbon chains and carbonyl group stretching suggesting the formation of ester linkages, respectively. Using the integrated area under these peaks, a grafting index (G1) was calculated using the following equation: 85 when repre: thein Elect] a non 10w 1 conce OXF'Q‘ from aIOms hydrc (nher Garbo carbo 11). taktsl A . GI X _ X (Modified) — (1) Ax (Unmodified) where x represents the absorbance band at either 2900 cm'l or 1740 cm", Ax (Modified) represents the integrated area of the peak after modification and Ax (Unmodified) represents the integrated peak area of the unmodified wood particles. X-ray photoelectron spectroscopy (XPS) analysis was carried out on a Physical Electronics Phi 5400 ESCA System, (Physical Electronics USA, Chanhassen, MN) using a non-monochromatic Mg source and a takeoff angle of 45° relative to the detector. A low resolution scan from 0 to 1100 eV binding energy was used to determine the concentration of each element present on the surface of the samples, along with the oxygen to carbon (O/C) atomic ratio, whereas a high resolution scan of the C1, region from 280 to 300 eV was performed to further analyze the chemical bonding of the carbon atoms. Carbon components C1 arise from carbon atoms bonded only to carbon and/or hydrogen atoms (C-C/C-H), C2 from carbon atoms bonded to a single oxygen atom, other than a carbonyl oxygen (C-OH), C3 from carbon atoms bonded to two non- carbonyl oxygen atoms or to a single carbonyl oxygen atom (O-C-O, C=O), and C4 from carbon atoms which are linked to a carbonyl and a non-carbonyl group (O-C=O) (9, ll, 17). The procedure for XPS data collection and analysis was detailed in other articles (9, 11). As mentioned, the chemical reaction of maleated polyolefins and wood particles takes place between the maleate groups of the polyolefin and hydroxyl groups on the 86 surface of wood. Since C2 component arises from atoms bonded to a single oxygen atom, other than a carbonyl oxygen (C-OH), the change in the content of C2 component before and after modification can be used to monitor the occurrence of the esterification between the maleated polyolefins and the wood particles. To quantify this change, a hydroxyl index (HI) was calculated from C2 component of C1, data as follows: C2 Unmodified where CZModjficd and CZUnmodified represent C2 after modification and in the unmodified wood particles, respectively. A complimentary technique to FTIR and XPS analyses, titrimetric analysis was also performed on unmodified and modified wood particles to provide additional proof of the esterification reaction, and to elucidate the mechanism of the chemical reaction between wood and MAPE. Titrations were carried out following a procedure described elsewhere (17), using an Oakton pH CON 510 pH meter. Accurate endpoints were difficult to determine when a visual indicator was used because wood particles significantly darkened the solution color. To insure greater accuracy, potentiometric titrations were used with a first derivative method for endpoint determination (18). Four samples each of unmodified and modified wood particles were analyzed for acid value (AV), saponification value (SV), and hydroxyl value (HV). Acid value accounts for the amount of free carboxylic acid groups present in the sample, while saponification value 87 measures both acid and ester groups in the sample. Hydroxyl value quantifies the free and accessible hydroxyl groups present in the sample (17). These quantities were calculated using the following formulas (17): AV = 56.1 1‘1— (3) p sv = 56.1(b--a)E (4) p HV = [(b — a)N (56.1) + AV (5) where N is the normality of potassium hydroxide (KOH), p is the amount in grams of wood particles, b is the volume of KOH needed to neutralize the blank (solvent only) and a is the volume of KOH required to neutralize the sample. The factor 56.1 accounts for the molecular weight of KOH. 88 RESULTS AND DISCUSSION Effect of Maleated Polyethylene (MAPE) Content FTIR spectroscopy was used to monitor and quantify changes that occurred on the surface of wood particles after reactive extrusion with MAPE. Infrared spectra of unmodified wood particles, pure MAPE, and wood particles modified with various concentrations of MAPE are shown in Figure 4.3. Table 4.3 lists the wavenumbers of peaks found in these spectra, along with assignments of corresponding functional groups. Regardless of the MAPE content, evidence of the grafting of MAPE compound to the surface of wood particles was apparent in the spectra of modified wood particles (spectra C-F). A distinct change was clearly seen near the absorption band at 2900 cm", where a large band between 2920 and 2850 cm", having two distinct peaks similar in appearance to pure MAPE (spectrum B), replaced the single peak in the unmodified wood particle (spectrum A). This change was most noticeable at MAPE concentrations above 10% where the integrated area under the absorption band near 2900 cm'1 increased in the modified wood particles, compared to the unmodified ones, implying the grafting of the pendant polyethylene chain of MAPE to the surface of wood particles. Similarly, the integrated area under the absorption band near 1740 cm'1 has significantly increased for wood particles extruded with MAPE compound. Several authors have correlated the increased integrated area or peak height of this band with the esterification reaction between wood particles and maleated compounds, since absorbance in the range of 1725- 1750 cm'I is characteristic of ester carbonyl stretching (5-10, 12-14, 19). 89 93 2900 1740 8 1 6 1710 723 D 2919 285 In 3400 can—3c rang-mx—accx -n o 3400 29 1740 1740 ‘ 4 1122 4000 3500 3000 2500 2000 1 500 1000 500 Wavenumbers (cmA-1) Figure 4.3. FTIR spectra of unmodified wood particles (spectrum A), pure maleated polyethylene-MAPE (spectrum B), wood particles modified with 5% MAPE (spectrum C), 10% MAPE (spectrum D), 15% MAPE (spectrum E), and 20% MAPE (spectrum F) after a second 24-hour Soxhlet extraction in the region4000-400 cm". 90 Am A 5 ............... 2 SEE; moose a $8 ..... MS a cone? wide :0 .......... 83% .......... . . Smashed“. :o ”2 EN: . - ..... . a 2 S 922% odd Ba 06 as: a: $2 8: 3.: mm: 82 2 maocafieoo £0 £2 ..... 8.2 ..... 82-8: 2 223553 £0 83.8: 83-3: 3: S: 23.3: 2.2 9.2an ma: 2582 as. _ do: as. _ do: .......... 8: -82 a a a a A—NAF—onao ”momv lllllllllll o. 2 2 do msgaeaob a: E: . . . . 2.8380 53.8 ........ o. 2 E d o messed Ono 8: 8: - - 2.: 2.2 9222:. 9:0 £635 .......... fl: WE ...... 2.2.2 £0 .foao @222; :0 Sam; mom $3-83. 3088 33.2% 8% 2.3.2 3232:. :0 82.82 82.82 .......... 83-8mm p.83 .552 9.53 $22 nae p-83 A as o8: 082%: ado—Swag 0.3.— 5? .358... a? .552... $22 $22 _. 35.55:: 8.39::— coog 8.3:»:— vooB 9:..— 9:5 838E coo? BEBE Ea .722 25 .mdfiz 8.5 gonad 8o? Homage: é acoaowama. 93 95m ooueomo< a: .3 2.3 91 The grafting efficiency of wood particles with various MAPE contents through a reactive extrusion process was quantified by calculating the grafting index (Equation 1) from the integrated area under the peaks near 2900 cm'1 and 1740 cm'1 for unmodified and modified wood particles. Figure 4.4 clearly illustrates that the grafting index increased with MAPE content up to 15%, independent of the absorption peak used. There appeared to be some leveling off between 15% and 20% MAPE concentration, which may indicate a level of maximum grafting efficiency has been reached. These results suggest that the esterification reaction was a function of the MAPE concentration used. 92 3.0 8 2.5 E 3 E 2.0 I: :3. S 1.5 U w E 1.0 g - -o--2900 cm" x 0.5 < —C}—-1740cm" 0.0 0 5 10 15 20 25 MAPE Content (°/o) Figure 4.4. Grafting index for FTIR absorbance bands near 2900 cm'1 and 1740 cm" for unmodified wood particles and wood particles modified with 5-20% MAPE. 93 XPS and titrimetric data listed in Tables 4.4 and 4.5, respectively, corroborated the conclusions drawn from FTIR analysis. As expected, reactive extrusion of wood particles with MAPE caused an increase in the concentration of unoxidized carbon atoms (Cl component) and decreased the contents of oxidized carbon atoms (C2-C4 components). Consequently, a significant decrease of the O/C atomic ratios was observed due to the presence of aliphatic carbons of polyethylene chains of MAPE. Furthermore, both the hydroxyl index (HI in Table 4.4) and the free and accessible hydroxyl groups (HV in Table 4.5) significantly decreased after modification with MAPE compounds through a reactive extrusion process. The decrease in both hydroxyl index and free and accessible hydroxyl groups after modification implied that esterification reaction took place through the hydroxyl groups on the surface of the wood particles. Unlike FTIR, which clearly showed an increase in ester carbonyl as a function of MAPE content, C3 component (carbonyl groups) showed a decreasing trend as MAPE content increased, while C4 component (ester groups) was not detected at all on the surface of the modified wood particles. This apparent difference can be explained by the higher surface sensitivity of XPS, which has a probing depth of only a few nanometers (l l, 16). Unoxidized carbon from the pendant polyolefin chain of MAPE was concentrated on the wood particle surface, as evidenced by the high content of C1 component (Table 4.4), and may have obstructed the detection of C3 and C4 components during XPS analysis. 94 .N corms—om 80¢ toga—=28 22:: :9 .mD—n—Eflm 050m cm COHOOHDV mag “Tr—8&0 H—Omumowtuofio or: EOHM CNN docombxo HoExom __ vm vacuum a cote 8:350 805 803 com: 8.253 uooB BEBE .28 35608:? mod 36 _m.o omd 26 co; ME: 59: IO . . . . . . . . med: 2% a? 8 o 2 c 3” a we 3 8 o 8 o 8 8 8 N 8288 8328 825 . . . . . . . . max: 22 £3 8 o a _ om m ow No B o 3. o 3 mo 2 o 8285 8322 825 . . . . . . . . .232 $2 a? 8 c K m 8 2 NM 3 2 o E c an a a : 8288 3323 253 . . . . . . . . med: .2 as, 8 c we 4 cm 2 S a 2 o a o 2 mm 3. 2 B285 3323. Boa 85 2a Ks 2.2 :.o cod 5.2 8.2 $22 and ”no 2; cm. 2 m 3.3 So 85 m _ .3 a. _ m 8328 ES, 8288:: 5 8 8 6 8:8 New u o 28on GE MECBNE g 38o 5.6 $38.... 20 .. mcocmmomfioo 3:082“; max .3 358869 mac—omtam coo? mo 38¢ .20 :ouEOmoy—éwm: use muoEmanoU ooflBm Esau—m den £an 95 Table 4.5. Hydroxyl Value, Acid Value and Saponification Value Determined by Titrimetric Analysis Titration values2 Wood particles' Hydroxyl value Acid value Saponification value (HV) (AV) (SV) (eq/kg) (eq/kg) (err/kg) Unmodified 268 i 4 3.15 i 0.11 64 i 6 Modified with 5% MAPE 233 j: 8 3.37 i 0.23 65 i 4 Modified with10%MAPE 233i26 3.53i0.12 79i8 Modified with15%MAPE 217111 3.87i0.20 132:10 Modified with 20% MAPE 215:3 4.00:0.24 131:8 lUnmodified and modified wood particles used were those obtained after a second 24-h Soxhlet extraction. 2Titration values represent an average of four samples. 96 As shown in Figure 4.1, two possible reactions could occur between MAPE and wood particles in a reactive extrusion process: a single site reaction, which lead to the formation of monoester with carboxylic acid pendent groups (Figure 4.1a) and/or or diester formation without carboxylic acid pendent groups (Figure 4.1b) (9, 13). Acid value (AV) and saponification value (SV) determined by titrimetric analysis (Table 5) were used to elucidate which reaction had occurred. The AV, which accounts for the free carboxylic acid group content in the system, slightly increased in the samples after modification (Table 4.5). However, this increase was not significant compared to the unmodified wood particles. Thus, there were a negligible amount of free carboxylic acid groups in the system. By contrast, SV, which accounts for both ester and acid groups, increased significantly after modification (Table 4.5). Since there was a negligible amount of free carboxylic acid groups on the surface of the modified wood particles, the increased SV originated mainly from the formation of maleate ester with OH groups on the wood particle surface. This result suggests that most of the MAPE was attached to the wood particle surface via two acid groups from the cyclic anhydride of the MAPE, i.e., through diesterification reaction (Figure 4.1b). Moreover, FTIR spectra shown in Figure 4.3 support this mechanism. Two distinct peaks in the range 1700-1750 cm'l should have been detected if the esterification reaction had occurred through monoester reaction (6, 9, 13). The first peak at around 1705-1710 cm'1 is associated with the non- reacted carboxylic acid (monoester formation) whereas the second one near 1725-1750 cm'1 are caused by the ester carbonyl absorption of the reacted MAPE with the OH groups of wood (diester formation) (6,9,13). As seen in Figure 4.3, all spectra of modified wood particles showed only one peak at 1740 cm", which is characteristic of 97 ester carbonyl, and the absence of bands at 1705-1710 cm'1 in these spectra clearly proved that there were no free carboxylic acid groups in the samples, thus confirming diester formation. The relatively high temperature (160°C) used to modify the wood particles in the reactive extrusion process may also explain why the diester form was predominant in this system. Other authors have reported higher monoester formation in the reaction of maleated compounds with wood particles in solvents at room temperature (3-8). Additional heat was required to cause the second acid group of the maleated compound to react with wood particles, forming the diester (4,6). Hon and Xing determined that diesters began to increase in their system at temperatures above 140°C (4). Similarly, Felix and Gatenholm found more diester content than monoester in cellulosic particles modified with MAPP at 100°C (13). Since the temperature used for reactive extrusion of wood particles in this study was even higher than those reported by other authors, it was likely sufficient to cause the reaction of both acid groups in the MAPE with hydroxyl groups on the wood particle surface, resulting in diester as the primary ester form. Although AV and SV were calculated, it was not possible to quantify the total ester content in these samples because calculation of monoester, diester and total ester content require sample weight gain after modification (3, 4). During the reactive extrusion process, waste material is produced at the beginning and end of the run, so not all modified sample can be collected. Even without this information, evidence of the grafting reaction suggests that reactive extrusion is an efficient process for modifying wood particles with binding agents such as MAPE compounds. 98 Effect of Extrusion Processing Conditions In a prior part of this study, wood particles were modified at a constant extrusion temperature profile (160°C) and rotational screw speed (60 rpm). Since the chemical reaction between wood particles and maleated polyolefins can be affected by both temperature and residence time in the extruder, it is important to understand how these extrusion processing conditions would affect the grafting efficiency of wood particles modified through reactive extrusion. Figure 4.5 shows the grafting indices for the absorption bands near 2900 cm'1 (Figure 4.5a) and 1740 cm'1 (Figure 4.5b) plotted for each extruder barrel temperature over the range of rotational screw speeds investigated. There was a clear indication that the grafting indices increased after modification, irrespective of the extrusion temperature profile and rotational screw speed used. The increased grafting index values indicated that more CH2 and C=O groups had been attached to the wood particle surface, which implied that the esterification reaction had occurred between the wood particles and MAPE. However, surface modification was not a function of processing conditions used since no trend in grafting index was observed with increasing the extruder’s temperature profile or increasing rotational screw speed for either absorbance band. 99 AmMOdifiOd/Azgoounmodified Amomodifiedlmmunmodified Figure 4.5. 3.0 a 2.5 ( ) 1‘21 2.0 X . —-e--130°C o ---- 9 """ V ——r:}— 140°C 1.5 ---A--- 160°C - -o - —180°C 1.0 X )l( X $K +Control 0.5 0.0 . O 20 40 60 80 100 Rotational Screw Speed (rpm) 3.0 2.5 2.0 mew-130°C —D— 140°C 1.5 ----A--160°C - -e - —180°C 1.0 X )K >K X +Control 0.5 0.0 0 20 40 60 80 100 Rotational Screw Speed (rpm) Effect of processing conditions on grafting index for unmodified and modified wood particles over the range of extruder barrel temperature and rotational screw speed combinations: (a) band near 2900 cm'] and (b) band near 1740 cm". 100 XPS results agreed with the above conclusions obtained from the infrared spectroscopy (Tables 4.6 and 4.7). Changing the extruder’s rotational screw speeds while maintaining a constant temperature profile (Table 4.6) or vice versa (Table 4.7) did not affect the grafting efficiency. In both cases, it was apparent that the surfaces of modified wood particles had been chemically changed from the unmodified wood particles, as evidenced by the large increase in the concentration of unoxidized carbon atoms (C1 component) coupled with a significant reduction in both the O/C atomic ratios and hydroxyl indices (HI) under all processing conditions. However, no trend was observed in C1, H1 or the O/C atomic ratio as the extruder’s rotational screw speed or barrel temperature was increased. 101 Table 4.6. Effect of Extruder’s Rotational Screw Speed on Surface Chemistry of Wood Particles Modified with 20% MAPE at 160°C , Elemental $821231]; compositions O/C Analysis of C1, peaks (%) 0H (%) atomic index screw speed ratios (HI) (rpm) 0 C c1 C2 C3 C4 ”nexm‘ded 31.85 68.15 0.47 39.75 51.20 8.46 0.58 1.00 (control) 20 5.61 94.39 0.06 92.62 5.12 2.26 0.00 0.10 40 3.72 96.28 0.04 95.80 3.40 0.80 0.00 0.07 60 2.39 97.61 0.02 96.90 3.10 0.00 0.00 0.06 80 2.97 97.03 0.03 95.08 4.34 0.58 0.00 0.08 Table 4.7. Effect of Extruder’s Barrel Temperature on Surface Chemistry of Wood Particles Modified with 20% MAPE at 60 rpm , Elemental E123: s compositions O/C Analysis of C1, peaks (%) OH tern r atur e (%) atomic index pe ratios (HI) (°C) 0 C C1 C2 C3 C4 unem’ded 31.85 68.15 0.47 39.75 51.20 8.46 0.58 1.00 (control) 130 3.36 96.64 0.03 95.27 3.04 1.69 0.00 0.06 140 5.47 94.53 0.06 92.61 5.26 2.12 0.00 0.10 160 2.39 97.61 0.02 96.90 3.10 0.00 0.00 0.06 180 5.05 94.95 0.05 95.24 4.76 0.00 0.00 0.09 Effect of Maleated Polypropylene (MAPP) Molecular Weight Maleated polyolefins with a wide variety of molecular weights are available on the market. Four different MAPP compounds were used (Table 4.1) to determine how different molecular weights (from a low of 11,200 to a high of 52,000 g/mol) affect the surface modification of wood particles in reactive extrusion process. To achieve this objective, the MAPP content was held constant at 20% while the extrusion temperature profile and rotational speed of the screws were set at 160°C and 60 rpm, respectively. Figures 4.6 and 4.7 illustrate the FTIR spectra of pure MAPP compounds with various molecular weights and wood particles modified with these MAPPs, respectively. The peak assignments for these spectra are also listed in Table 4.3. Evidence of the grafting of MAPP to the wood particles through esterification is apparent in the modified wood spectra (Figure 4.7, spectra B-E) because the single absorption band near 2900 cm'1 in the spectrum of unmodified wood particles has been replaced by a large band between 2955 and 2840 cm", having distinct peaks similar in appearance to the pure MAPP counterparts. Another indication of grafting of the maleated polyolefin was an increased area of the absorption band at around 1740 cm", from C=O groups attached to the wood surface, likely from the esterification reaction (5-10, 12-14) Although not clearly seen in Figure 4.7 due to stacking of five spectra in one figure, the individual spectra of modified wood particles showed an increased intensity in the bands at 1462 and 1376 cm'1 which also indicated the grafting reaction had occurred and added more C-H to the surface of the modified samples (10, 12, 14). 103 xacg-mx—oacx “was: A 4000 3500 3000 2500 2000 1500 1000 500 Wavenumbers (cmA-1) Figure 4.6. FTIR spectra of pure MAPP compounds with various molecular weights: 11,200 g/mol or E-43 (spectrum A), 39,000 g/mol or G-32l6 (spectrum B), 47,000 g/mol or G-3015 (spectrum C) and 52,000 g/mol or G-3003 (spectrum D) in the region 4000-400 cm". 104 O O O) N (I) V- 05 N racgonr—wacx to ur-mla C: f 4000 3500 3000 2500 2000 1500 1000 500 Wavenumbers (cm"-1) Figure 4.7 FTIR spectra of unmodified wood particles (spectrum A), wood particles modified with MAPP compounds of various molecular weights: 11,200 g/mol or E-43 (spectrum B), 39,000 g/mol or G-3216 (spectrum C), 47,000 g/mol or G-3015 (spectrum D), and 52,000 g/mol or G-3003 (spectrum E) in the region 4000-400 cm' . 105 Moreover, all four MAPP compounds had significantly increased grafting indices as compared to unmodified wood particles (Figure 4.8). However, no significant difference in grafting index was observed between the four types of MAPP-modified wood particles. These results were in agreement with those obtained from the XPS analysis (Table 4.8). As expected, the surface of modified wood particles resulted in an increased CI content and a significant decrease in both O/C atomic ratio and hydroxyl index (HI), all of which provided evidence of the surface modification through esterification reaction. 106 3.0 '3 2.5 E 3 2 0 E . C :3. 5 1.5 8 E 10 ‘3 ' -—.<>-—2900 cm" E 4’? 0,5 —1:1——1740cm'1 0.0 0 11,200 39,000 47,000 52,000 Mw of MAPP (glmol) Figure 4.8. Effect of molecular weight on grafting index for unmodified wood particles and wood particles modified with 20% MAPP compounds. 107 Table 4.8. Elemental Surface Compositions and High-Resolution C15 Peaks of MAPP, Wood Particles, and Wood Particles Modified with MAPP Determined by XPS Elemental compositions O/C Analysis of Cl, peaks (%) OH Materials (%) atomic index 0 C mm C1 C2 C3 C4 (m) 5'43 4 89 95 11 0 05 94 92 4 35 0 74 0 0 O8 (Mw=11,200) ‘ ' ‘ ‘ ' ' ° ' G-3216 (Mw=39,000) 3.86 96.14 0.04 95.85 3.24 0.91 O 0.06 G-3015 (Mw=47,000) 3.77 96.23 0.04 96.01 2.62 1.37 O 0.05 G-3003 (Mw=52,000) 3.28 96.72 0.03 96.58 2.57 0.84 0 0.05 Unmodified Wood 31.85 68.15 0.47 39.75 51.20 8.46 0.58 1.00 Wood-E-43 7.22 92.78 0.08 93.39 4.75 1.86 0 0.09 Wood-03216 3.60 96.40 0.04 94.86 4.17 0.97 0 0.08 Wood-G-3015 3.36 96.64 0.04 95.38 4.20 0.42 O 0.08 Wood-G-3003 5.87 94.13 0.06 90.99 6.19 2.83 0 0.12 108 However, no distinct trend was observed between molecular weight of MAPP and grafting efficiency through a reactive extrusion process. Since both FTIR and XPS results confirmed that there were no significant differences in grafting of MAPP with various molecular weights, the molecular weight range studied (11,200-52,000 g/mol) may not have been large enough to observe differences in reactivity. Another possibility could be that at 20% MAPP content, the maximum level of grafting had already been achieved with each compound, regardless of its molecular weight. Thus the differences in the grafting efficiency were not detected. Moreover, Figure 4.4 clearly illustrates that the GI leveled off between 15 and 20% maleated polyolefin content, which may indicate a level of maximum grafting efficiency had been reached. Perhaps another study at lower MAPP concentrations would elucidate differences of MAPP molecular weight on the efficiency of the grafting reaction. 109 CONCLUSIONS This study examined the chemical reactions between maleated polyolefins and wood particles in a reactive extrusion process. The effects of maleated MAPE content, extrusion processing conditions (barrel temperature and rotational screw speed), and molecular weight of MAPP were studied, with the goal of determining the effects of each set of conditions on the grafting efficiency of modified wood particles in a reactive extrusion process. Efficiency of the modification was evaluated using FTIR, XPS and titrimetric analysis. From the experimental results, the following conclusions can be drawn: 1. The esterification reaction between wood particles and MAPE was a function of the MAPE concentration used to modify the wood particles. The grafting reaction produced mostly the diester form of the modified wood particle during reactive extrusion. 2. No significant difference was found in grafting efficiency of the modified wood particles at different extrusion processing conditions. Changing the extruder’s barrel temperature profile (130-180°C) and/or its rotational screw speed (20-80 rprn) resulted in adequate grafting, which indicated that the esterification reaction was not a function of processing conditions over the range studied. 3. Regardless of MAPP molecular weight (from a low of 11,200 to a high of 52,000 g/mol) all investigated MAPP compounds were effective in changing the surface of wood particles after modification, compared to unmodified wood particles. However, no distinct trend was observed between molecular weight of MAPP and grafting efficiency through a reactive extrusion process. Iit is believed that the 110 high content of MAPP (20%) used in this study prevented the detection of differences in the grafting efficiency because the maximum level of grafting reaction had already occurred with each component at 20% MAPP content. . The reactive extrusion process was found to be a suitable way to modify wood particles with maleated polyolefins as it worked quickly and without the use of solvents. This process would be industrially friendly, and in light of our prior success bonding the modified fibers together in a hot press, would allow large quantities of modified wood particles to be produced for the manufacture of a formaldehyde-free wood composite product. 111 10. 11. REFERENCES Carlbom, K. and Matuana, L.M., "Composite materials manufactured from wood particles modified through a reactive extrusion process," Polymer Composites, 26 (4): 534-541 (2005). Matuana, L.M. and Carlbom, K., US Provisional Patent # 60/592,918, July 30, (2004) Matsuda, H., "Preparation and utilization of esterified woods bearing carboxyl groups," Wood Science and Technology, 21 (1): 75-88 (1987). Hon, D.N.S. and Xing, L.M., "Thermoplasticization of wood. Esterification," in Viscoelasticity of Biomaterials, W.G. Glasser and H. Hatakeyama, Editors. American Chemical Society: Washington DC. p. 118-132 (1992). Marcovich, N.E., Reboredo, M.M., and Arangguren, M.I., "Sawdust modification. Maleic anhydride chemical treatment," Holz als Roh- und Werkstoff, 54 (3): 189- 193(1996) Aranguren, M.I., Marcovich, N.E., and Reboredo, M.M. "Sawdust and woodflour: Its esterification and use in the formulation of polymer composites," in Recent Advances in Biotechnology for Tree Conservation and Management, Proceedings of an IFS Workshop, Brazil, (1998). Timar, M.C., Mihai, M.D., Maher, K., and Irle, M., "Preparation of wood with thermoplastic properties. Part 1. Classical synthesis," Holzforschung, 54 (l): 71- 76 (2000). Timar, M.C., Maher, K., Irle, M., and Mihai, M.D., "Preparation of wood with thermoplastic properties. Part 2. Simplified technologies," Holzforschung, 54 (1): 77-82 (2000). Matuana, L.M., Balatinecz, J .J ., Sodhi, R.N.S., and Park, C.B., "Surface characterization of esterified cellulosic fibers by XPS and F TIR Spectroscopy," Wood Science and Technology, 35 (3): 191-201 (2001). Kazayawoko, M., Balatinecz, J .J ., and Woodhams, R.T., "Diffuse reflectance Fourier transform infrared spectra of wood fibers treated with maleated polypropylenes," Journal of Applied Polymer Science, 66 (6): 1163-1173 (1997). Kazayawoko, M., Balatinecz, J .J ., and Sodhi, R.N.S., "X-ray photoelectron spectroscopy of maleated polypropylene-treated wood fibers in a high-intensity thermokinetic mixer," Wood Science and Technology, 33 (5): 359-3 72 (1999). 112 12. 13. 14. 15. 16. 17. l8. l9. Kazayawoko, M., Balatinecz, J.J., and Matuana, L.M., "Surface modification and adhesion mechanisms in wood fiber-polypropylene composites," Journal of Materials Science, 34 (24): 6189-6199 (1999). Felix, J .M. and Gatenholm, P., "The nature of adhesion in composites of modified cellulose fibers and polypropylene," Journal of Applied Polymer Science, 42 (3): 609-20 (1991). Li, Q. and Matuana, L.M., "Surface of cellulosic materials modified with functionalized polyethylene coupling agents," Journal of Applied Polymer Science, 88 (2): 278-286 (2003). Cheremisinoff, N.P., Guidebook to Extrusion Technology. Prentice Hall: Englewood Cliffs, NJ. (1993). Ratner, 8D. and Castner, D.G., "Electron Spectroscopy for Chemical Analysis," in Surface Analysis .' The Principal Techniques, J .C. Vickerman, Editor. John Wiley & Sons: Chichester, England. p. 43-98 (1997). Anonymous, "Recommended methods for the analysis of alkyd resins," Pure and Applied Chemistry, 33 (2-3): 411-435 (1973). Harris, D.C., Quantitative chemical analysis, 4th Ed., W.H. Freeman: New York. Chapter 12 (1995). Socrates, G., Infrared characteristic group frequencies: tables and charts, 2nd Ed., John Wiley & Sons, Inc.: New York (1994). 113 CHAPTER 5 INFLUENCE OF PROCESSING CONDITIONS AND MATERIAL COMPOSITIONS ON THE PERFORMANCE OF F ORMALDEHYDE-F REE WOOD-BASED COMPOSITES This chapter is in press for Polymer Composites (2006). It is co-authored by K. Carlbom and L.M. Matuana. 114 ABSTRACT This study examined the differences between formaldehyde-free wood composite panels made with maleated polyethylene (MAPE) and maleated polypropylene (MAPP) binding agents. Specifically, the study investigated the contrasts of (i) base resin type, PE vs. PP, (ii) molecular weight/maleic anhydride content in MAPP binding agents, and (iii) the manufacturing methods (reactive extrusion vs. hot press) on the physico- mechanical properties of the composites. FTIR and XPS analyses of unmodified and modified wood particles after reactive extrusion with maleated polyolefins provided evidence of chemical bonding between the hydroxyl groups of wood particles and maleated polyolefins. While extruding the particles before panel pressing gave better internal bond strength, superior bending properties were obtained through compression molding alone. MAPP-based panels outperformed MAPE-based panels in stiffness. Conversely, MAPE increased the IB strength of the panels compared to MAPP. Polymer base resin had no effect on modulus of rupture or screw holding capacity. Differences between the two maleated polypropylene compounds were not significant for any of the mechanical properties tested. Formaldehyde-free wood composites manufactured in this study often outperformed standard requirements for conventional particleboard, regardless of material composition or manufacturing method used. 115 INTRODUCTION The chemical reaction between wood and anhydride compounds has been shown to plasticize the wood, making it more versatile for wood composite applications. Several investigators have studied wood modification with anhydrides in order to determine thermal properties, moisture resistance, compatibility with polymeric matrices, etc. (l-ll). A lower softening temperature, compared to unmodified wood, has been documented in esterified wood after reaction with various anhydrides (2, 8, 10, ll). Esterified wood has also been found to resist moisture (l, 3, 4-11) and to be more compatible than unmodified wood with polyester matrices in the manufacture of composites (4-6). A second reaction of the esterified wood with epoxides formed oligoesters, which further thermoplasticized the wood and allowed for better control of structure and resulting properties of the modified wood (1, 7-11). While the thermoplastic nature of esterified and oligoesterified wood modified through these methods has been documented, few authors have exploited this characteristic to make wood composites without added adhesive. Matsuda and co- workers used three dicarboxylic acid anhydrides (maleic, phthalic and succinic anhydride) to esterify wood particles, which were then molded into sheets through compression molding ( 1, 8). Clemons and co-workers used similar methods to modify wood fibers with anhydrides to prepare fiberboards (3). In both studies, chemically modified wood was able to partially melt when heated and bond without additional adhesive. Formulations containing succinic anhydride were found to have the most thermoplastic character in both studies (1 , 3, 8). Further work by Matsuda et a1. (1 , 7, 8) showed enhanced thermoplasticization of wood through the grafting of various types of 116 epoxides onto the already esterified wood surface. The oligoesterified wood produced from these reactions was even more thermoplastic-like in structure than the esterified wood produced through reaction with the anhydrides, and could be easily molded. Using various types of epoxides allowed for different chemical structures in the thermoplasticized wood, and could also be used to form crosslinked wood composites which resembled plastic more than wood (1, 7, 8). Recently, Timar and co-workers demonstrated the use of maleic anhydride to esterify wood particles, followed by oligoesterification with glycidyl methacrylate and additional maleic anhydride as a process to form thermoplastic-like wood particles (9-11). Panels formed through the compression molding of modified wood particles displayed mechanical properties that met or exceeded standard requirements for bending strength and internal bond strength of particleboard or fiberboard (11). These wood composites were also found to be resistant to fungal decay. Although esterifying and oligoesterifying the wood surface has been shown to be an effective method for manufacturing wood composite materials without additional adhesive, most methods require harmful solvents such as xylene, dimethyl sulfoxide, or N,N-dimethylformamide as part of the modification process (1-6, 8-11). Additionally, the oligoesterification reactions entail a two-step process (or more depending on the precise control of chemical structure desired) to functionalize the wood surface, which is both time consuming and complicated (1, 7-11). Removal of solvents and further drying of the wood particles is required before compression molding into panels (1, 3, 8, 11). A simpler approach of functionalizing wood particles to produce wood composite panels without additional adhesive has been developed in our prior work, which 117 demonstrated the concept of using a reactive extrusion process as a means of developing a new, formaldehyde-free binding system for wood composite products (12). It has been shown that the surface modification of wood particles with maleated polyolefins as the binding agent could be achieved through a continuous reactive extrusion process without the use of solvents (i.e., dry process). Surface analysis of the modified wood particles indicated that the maleate groups of the binding agent were attached to the surface of the wood particles via an esterification reaction, exposing polyolefin chains on the surface of the particles (12, 13). The modified wood particles were then compression molded to produce formaldehyde-free wood composite panels. During panel manufacturing, the pendant polyolefin chains attached to the wood particles melted and flowed under heat and pressure in the hot press, forming entanglements. The entangled polymer chains locked together after cooling, forming direct particle-particle bonds. The physico- mechanical properties of these composite panels, which contain no formaldehyde-based adhesive, compared favorably with or even exceed current standard requirements for particleboard of medium density. Although formaldehyde-free wood composite panels were successfully manufactured through a reactive extrusion process (12), the process required two different steps: (i) reaction of the wood particles with the maleated polyolefin, and (ii) hot pressing the modified particles into panels. A method that could reduce this process to one step would save time and reduce production costs. Therefore, we hypothesized that both the esterification reaction and binding into panels could occur simultaneously during compression molding, which would greatly simplify the panel manufacturing process. 118 Furthermore, there currently are many types of maleated polyolefins, differing in molecular weight, base resin type, acid value, percentage of maleic anhydride, etc. Research has not been carried out to examine the effects of these maleated polyolefin variables on the performance of formaldehyde-free wood composites. Thus, the objective of this work was to investigate the effects of material compositions as (i) base resin type of the maleated polyolefin, either polyethylene or polypropylene, and (ii) maleic anhydride content/molecular weight of maleated polypropylene, as well as manufacturing methods (reactive extrusion vs. hot press) on the physico-mechanical properties of the composites. Emphasis was also placed on determining the effectiveness of the modification after reactive extrusion of wood particles with maleated polyolefins using Fourier transform infrared (FTIR) spectroscopy and X-ray photoelectron spectroscopy (XPS). 119 EXPERIMENTAL Materials Maple wood particles of 425 micron (40-mesh) and 150 micron (100-mesh) size were supplied by American Wood Fibers (Schofield, WI) and were used as particles. The smaller particles (150 micron) were used for surface characterization to minimize the effects of scattering and specular reflectance in the diffuse reflectance IR analysis. However, panel manufacturing required a large quantity of modified particles, so larger (425 micron) particles, which were easier to process, were used in panel manufacturing and property testing. Hydrated zinc acetate, the catalyst, and xylene (99.9%, ACS Grade), the solvent used for Soxhlet extraction, were obtained from Baker Analytical Reagents (JT Baker Co., Phillipsburg, NJ). Maleated polyethylene-MAPE (G-2608) and two maleated polypropylenes-MAPP (G-3003 and G-3015) supplied by Eastman Chemical Co. (Kingsport, TN) were used as binding agents. Characteristics of the maleated compounds are listed in Table 5.1. All chemicals were used as received. 120 Table 5.1. Characteristics of the maleated polyolefins used as binding agents Properties MAPE MAPP MAPP G-2608 G-3003 G-3015 Weight % maleic anhydride 1.5 1.5 2.5 Melting point (°C) 122 156 155 Average molecular weight (g/mol) 51,700 52,000 47,000 lMelt flow index at 190°C 8 12.7 Viscosity at 190°C --- 60,000 25,000 lMelt flow index measured at 190°C and 2.16 kg according to ASTM D1238. 121 Surface Characterization of Wood Particles For batch compounding, 150 micron size wood particles were dried for 48 hours at 105°C to a final moisture content of less than one percent before processing. A 10-1iter high intensity mixer (Papenmeier TGAHK20) was used for dry blending of the wood particles, binding agent, and catalyst. The woodzbinding agentzcatalyst weight ratio was 79:20:l. All components were combined in the mixer and blended for 10 minutes at room temperature. Reactive extrusion of wood particles was achieved by feeding the compounded wood particles into a 32 mm conical counter rotating twin-screw extruder (C. W. Brabender Instruments, Inc.) with an L/D ratio of 13:1, driven by a 7.5 hp lntelli-Torque Plasti-Corder Torque Rheometer®. The barrel temperatures for the three zones inside the extruder were set at 160°C for maleated polyethylene and 165°C for maleated polypropylene, and the rotational speed of the screws was held at 60 rpm during the experiments. Following the procedures used in prior work, unmodified wood particles and wood particles that had been modified through reactive extrusion were Soxhlet extracted, dried to a constant weight at 105°C and analyzed by FTIR (12, 13). Based on previous studies (12, 13), the regions of interest in the FTIR spectra of wood particles modified with maleated polyolefins were the absorbance bands near 2900 cm'1 and 1740 cm", for CH stretching of aliphatic carbon chains and carbonyl group stretching suggesting the formation of ester linkages, respectively. Using the integrated area under these peaks, 3 grafting index (G1) was calculated using the following equation: 122 A x (Modified) Ax (Unmodified) GIx = (1) where x represents the absorbance band at either 2900 cm'I or 1740 cm", Ax (Modified) represents the integrated area of the peak after modification and AK (Unmodified) represents the integrated peak area of the unmodified wood particles (13). X-ray photoelectron spectroscopy (XPS) was performed on both unmodified wood and wood modified with maleated polyolefin binding agents following the procedure used in prior work (12, 13). A high resolution scan of the C1, region from 280 to 300 eV was run to elucidate the chemical bonding of the carbon atoms. Curve fitting of this high resolution area resolved four different forms of carbon for unmodified wood particles, whereas three different forms were resolved for modified wood particles. Carbon components C1 arise from carbon atoms bonded only to carbon and/or hydrogen atoms (C-C/C-H), C2 from carbon atoms bonded to a single oxygen atom, other than a carbonyl oxygen (C-OH), C3 from carbon atoms bonded to two non-carbonyl oxygen atoms or to a single carbonyl oxygen atom (O-C-O, C=O), and C4 from carbon atoms which are linked to a carbonyl and a non-carbonyl group (O-C=O). Particular attention was given to the C2 component from the high resolution scan because C2 component arises from atoms bonded to a single oxygen atom, other than a carbonyl oxygen (C-OH). Therefore, change in the content of C2 component before and after modification can be used to monitor the occurrence of the esterification between the maleated polyolefins and the wood particles, which occurs at the hydroxyl groups on the wood surface. To 123 quantify this change, hydroxyl index (HI) was calculated from C2 component of C], data as follows: C2 Unmodified where C2M0dified and CZUnmodified represent C2 after modification and in the unmodified wood particles, respectively (13). Panel Manufacture Larger wood particles (425 micron size) were compounded at room temperature using the same procedure described above for the surface analysis. Following the mixing step, two different methods were used to manufacture the composite panels from the compounded wood particles as follows: The first method was a one-step process where the compounded wood particle mixtures were directly hot pressed without the reactive extrusion step. Compression molding was performed using a hydraulic press from Erie Mill Co. (Erie, PA). Panels were pressed at 193°C for 7 minutes using 8 MPa pressure. After pressing, panels were removed from the press and cooled at room temperature under compression for 15 minutes. Panel dimensions were 380 by 380 by 6 mm, with a target density of 720 kg/m3 . 124 The second method was a two-step process where wood particles were modified with maleated polyolefins in a reactive extrusion process, and then compression molded in a hot press. Reactive extrusion of wood particles followed the same procedure as previously described for the modification of wood particles for surface characterization. Once extruded, the wood particles were compression molded into panels using the above- described pressing conditions. Panel Property Testing Density was measured by two different methods for all panels: (i) a simple mass over volume calculation for three panels of each type and (ii) internal density profile (X- ray density analysis) using a Quintek QMS Density Profiler, model QDP-OIX, with 5 replicates per panel type. Three-point flexural, internal bond (IB) strength and screw holding capacity tests were performed on an Instron 4206 testing machine (using Series IX software) in accordance with procedures outlined in ASTM standard D1037-99 (14). The crosshead speeds were 3.05 mm/min, 8.13 mm/min, and 0.6 mm/min for flexural, 1B, and screw holding capacity tests, respectively. Screw holding capacity was carried out from the face of the panels. At least six samples were tested to obtain an average value for modulus of rupture (MOR), modulus of elasticity (MOE), IB strength and screw holding capacity, all of which were compared with values listed for particleboard of medium density in the standard ANSI A208.1-l999 (15). 125 Statistical Analysis A two-sample t-test was carried out with an or significance value of 0.05 to determine the effects of material compositions and manufacturing method on the density, flexural, internal bond and screw holding properties of the composites. Comparisons between binding agents’ base resin types and maleic anhydride contents/molecular weights were made under one manufacturing method. Whereas comparisons between manufacturing methods were performed under one base resin type or maleic anhydride content/molecular weight. All statistical analysis was performed using Design Expert software (Version 6) from Stat-Ease, Inc. Minneapolis, MN. 126 RESULTS AND DISCUSSION Surface Characterization of Wood Particles As mentioned, the reaction of wood particles with maleated polyolefins occurs between maleate groups of the maleated polyolefin and hydroxyl groups on the surface of wood, which lead to the formation of a monoester with carboxylic acid pendant groups (Figure 5.1a) and/or diester formation where both carboxylic acid groups react with the wood particle surface (Figure 5.1b). In order to form a wood composite panel, the pendant polyolefin chains grafted to the wood surfaces must melt and flow under heat and pressure during compression molding, forming entanglements. Upon cooling, these entangled polymer chains would be locked together, forming the panel (Figure 5.2). 127 11 0-0—CH2 / o / H 0’9 Monoester / a) ' OH CH, o / + 0 | CH or ; OH >/ M \ ('3' O b) / o—c—-CH2 / Wood / O—C—C part1cle Maleated / 8 H Di-ester polyolefin Modified wood particle Figure 5.1. Modification scheme for esterification reaction between wood particles and maleated polyolefins: (a) monoester and (b) diester formation. o 0 II II / l—O—C—CH2 + HZC—C—O— \ / \ / _O_C__S/\/\/\ Wfi—fi-O“§ / ('5 o h Modified wood particle l 2 Modified wood particle I Figure 5.2. Panel manufacturing scheme for the two-step method showing the bonding of pre-reacted wood particles. 128 F TIR spectroscopy was used to observe and quantify changes that occurred on the surfaces of wood particles after reactive extrusion with maleated polyolefins. FTIR spectra of unmodified wood, maleated polyethylene (MAPE G-2608) and wood particles modified with MAPE after reactive extrusion are shown in Figure 5.3. As stated earlier, the bands of interest in the FTIR spectra occurred near 2900 cm'1 and 1740 cm". In the spectrum of modified wood particles (spectrum C), two distinct peaks between 2927 cm'l and 2853 cm'1 similar in appearance to pure MAPE (spectrum B) replaced the single peak near 2900 cm'1 in the unmodified wood (spectrum A). The integrated area under this band increased after modification with MAPE (spectrum C), implying the grafting of the polyolefin chain of MAPE to the surface of the wood particles. Similarly, the integrated area under the absorption band near 1740 cm" significantly increased for wood particles extruded with MAPE. Several authors have correlated the increased integrated area or peak height of this band with the esterification reaction between wood particles and maleated compounds, since absorbance in the range of 1725-1750 cm'1 is characteristic of ester carbonyl stretching (1-6, 9, 10). Although not shown, similar infrared spectra were obtained for wood particles treated with maleated polypropylene binding agents, i.e., MAPP G-3003 and G-3015. 129 1710 racguur—aacx mama: 4000 3500 3000 2500 2000 1500 1000 500 Wavenu m bers (cm *-1) Figure 5.3. Example FTIR spectra of unmodified wood particles (A), pure MAPEG- 2608 (B), and wood particles modified with MAPEG-2608 (C) in the region 4000 to 400 cm". 130 The grafting efficiency of wood particles modified with MAPE (G-2608) and two types of MAPP (G-3003 and G-3015) through a reactive extrusion process was quantified by calculating the grafting index (Equation 1) from the integrated area under the peaks near 2900 cm'1 and 1740 cm'1 for unmodified and modified wood particles. Results listed in Table 5.2 show that grafting index increased after reactive extrusion with each maleated polyolefin, independent of the absorption peak used. These results suggest the esterification reaction had occurred during reactive extrusion. 131 Table 5.2. Grafting index for peaks near 2900 cm'1 and 1740 cm'1 for unmodified and modified wood particles with various maleated polyolefin compounds Grafting Index (GI)l Samples 1 1 2900 cm' 1740 cm' Unmodified wood particles 1.0 1.0 Wood modified with MAPE G-2608 2.3 2.1 Wood modified with MAPP G-3003 1.9 1.9 Wood modified with MAPP G-3015 1.8 1.9 1G1 is calculated using Equation 1. 132 XPS data listed in Table 5.3 supported the results of FTIR analysis. As expected, reactive extrusion of wood particles with each of the maleated polyolefins caused a significant increase in the concentration of unoxidized carbon atoms (Cl component) and corresponding decrease in the contents of oxidized carbon atoms (C2-C4 components) in the modified wood particles. Additionally, the hydroxyl index (HI) has significantly decreased after modification with MAPE or MAPP compounds through a reactive extrusion process. The reduction in hydroxyl index suggested that for all three maleated polyolefins, the esterification reaction took place through the hydroxyl groups on the surface of the wood particles. While the grafting index calculated from the F TIR analysis for the band near 1740 cm'1 indicated an increase in ester groups after modification, regardless of the type of maleated polyolefin used, C3 component (carbonyl groups) was reduced in all samples after reactive extrusion with maleated polyolefins, while C4 component (ester groups) was not detected at all on the surface of the modified wood particles (Table 5.3). This apparent difference can be explained by the higher surface sensitivity of XPS, which has a probing depth of only a few nanometers (l6). Unoxidized carbon from the pendant polyolefin chain of MAPE or MAPP was concentrated on the wood particle surface, as evidenced by the high content of C1 component (Table 5.3), and may have obstructed the detection of C3 and C4 components during XPS analysis (13). 133 Table 5.3. High-Resolution C15 Peaks of Wood Particles Determined by XPS Analysis of C15 Peaks (%) 011 index Materials C1 C2 C3 C4 (I-II)‘ uandlfied W°°d 39.75 51 .20 8.46 0.58 1,00 partrcles Pure MAPE G-2608 (Mw=51,700 g/mol) 87.13 7.71 5.16 0.00 0,15 Wood modified with MAPE G-2608 95.08 4.34 0.58 0.00 0.08 Pure MAPP G-3003 (Mw=52,000 g/mol) 96.58 2.57 0.84 0.00 005 Wood modified with MAPP G-3003 90.99 6.19 2.83 0.00 0.12 Pure MAPP G-3015 (sz47’000 g/mol) 96.01 2.62 1.37 0.00 005 Wood modified with MAPP @3015 95.38 4.20 0.42 0.00 0,08 lOH index calculated from Equation 2. 134 Physico-Mechanical Properties Density Table 5.4 summarizes the density (calculated and profile) for the panels manufactured in this study. An example of the x-ray density profile, illustrating the density of the sample at the face and core regions, is shown in Figure 5.4. Experimental panels were within the medium density range as specified in the ANSI standard A208.1, with average calculated density values ranging from 775-780 kg/m3 (Table 5.4). The calculated and overall (x-ray profile) density of the panels was nearly the same, regardless of the panel manufacturing method. However, the manufacturing method showed two distinct trends in the density profiles of the composite panels. Panels made from unextruded wood particles (one step process, i.e. hot press only) had a higher density in the face region than those containing extruded wood particles (two step process, e.g., reactive extrusion followed by hot press). This feature is typical of conventional glued particleboard, due to greater compaction and solidification of the faces of the panels which experience direct heat during the pressing cycle (17, 18). Conversely, panels manufactured from extruded wood particles had a higher density in the core region. 135 Table 5.4. Density data for experimental panels bound with maleated polyolefins X-ray Density Profile Calculated Panel Types Densi Overall Face Core (kg/m (kg/m5 (kg/m3) (kg/m3) MAPE G-2608 — unextruded 780 780 905 748 MAPE G-2608 — extruded 778 779 881 759 MAPP G-3015 — unextruded 780 781 906 745 MAPP G-3015 - extruded 782 783 833 783 MAPP G-3003 — unextruded 775 778 866 760 MAPP G-3003 — extruded 775 776 821 760 1000 ——-— - -+—~ -» - ------ g ’\ / . Core 5 600 Face Face :E‘ 8 400 0) o 200 o , 0 1 2 3 4 5 6 Thickness (mm) Figure 5.4. X-ray density profile of a sample made from unextruded MAPP G-3003, illustrating the face and core regions of a typical sample. 136 Mechanical properties of a material are strongly influenced by its density. Therefore, the above results suggest that the mechanical properties of the composites may be dependent on the processing conditions since density differences exist in the face and core region of the panels due to the manufacturing process. Since the overall density was relatively the same between the panels with extruded and unextruded wood particles, differences in mechanical properties associated with the processing conditions might be attributed to the density profile in the panels. Effects of Processing Conditions The modulus of rupture (MOR), modulus of elasticity (MOE), internal bond (IB) strength and screw holding capacity data for panels manufactured in this study, along with standard requirements for particleboard of medium density grades, ranging from 640-800 kg/m3, are presented in Table 5.5. To determine significant differences between processing conditions, comparisons were made by varying the processing conditions (rows 1 vs. 2 and 3 vs. 4 under experimental panels) while holding the maleated polyolefin type constant. 137 Table 5.5. Effects of processing methods and material compositions on the mechanical properties of particleboard panels bound with maleated polyolefins. Mechanical Properties1 Panel Types MOR MOE IB Strength Screw (MPa) (MPa) (MPa) Holding (N) Medium Density 11.0 — 16.5 1725 — 2750 0.40 - 0.55 900 — 1100 Grades1 Experimental Panels2 MAPE— 25.41 :30A 2068i233A 1.22:0.32A 1353 i184A unextruded MAPE“ 20.70i3.4B 1296:195B 2.07:0.69B 1563: 180A extruded MAPP - 30.04 i 6.6 A 3582 i 567C 0.43 : 0.19C 1469 i 330 A unextruded MAPP- 23.00:4.7B 2875 i347D 1.50:0.29D 1580:299A extruded lProperty requirement data is from standard ANSl A208.1- 1999-Particleboard. 2The means with different letters indicate significance between treatments at the a = 0.05 level, while the means with the same letter indicates no difference between treatments. 138 Mechanical properties of the composite panels differed depending on manufacturing processes. Panels had significantly higher MOR and MOE values when the wood particles were not extruded prior to pressing. This was likely due to localized melting of the unreacted maleated compounds and greater flow at the faces of the panels, causing compaction in the face region, due to the direct heat from the platens. The faces would see more heat throughout the pressing cycle, likely causing the reaction between the wood and binding agent as well. Greater compaction of the face region of the panels was supported by the higher density of that region, as determined through X-ray density profile analysis (Table 5.4). High face region density has been correlated to increased MOR and MOE in conventional particleboard made with formaldehyde-based adhesives (17, 18). In addition, increasing density is related to increased strength properties through a power law relationship (19). Since overall density was relatively the same between the panels with extruded and unextruded wood particles, the increased face density of panels with unextruded wood particles must be responsible for the enhanced bending properties of these panels. Although panels made with extruded wood particles had lower bending properties than their unextruded counterparts, these panels still exceeded the requirements for conventional particleboard in most cases. Table 5.5 also summarizes the IB strength of composite panels under different manufacturing methods. Internal bond strength is an indication of how well the particles are bonded in the panel, particularly at the core region. The experimental results indicated that the unextruded wood particles underwent both the grafting reaction and entanglement during compression molding since panels were successfully produced without pre-reacting wood particles in the extruder. Unlike the bending properties, 139 panels prepared from unextruded wood particles had lower IB strength, compared to those made from extruded wood particles. Therefore, the lesser lB strength of composite panels with untreated wood particles was attributed to the reduced density in the core region of these boards, due to the heat not flowing to the center of the panel fast enough to cause adequate polymer flow and panel compaction during the limited pressing time. Similar trends have been reported for conventional glued particleboard (17, 18). In addition, heat is required to drive the reaction between the wood particles and the maleated polyolefins. Since the unextruded wood particles experience heat only during the hot pressing step, they may not receive sufficient heat to complete the esterification reaction and form chemical bonds between the wood and maleated polyolefins, especially in the core region of the panels. By contrast, wood particles that were pre-reacted in the extruder likely had more extensive bonding due to the extra heat and mixing during the extrusion step. This accounts for the significantly higher IB strength in panels made with extruded wood particles. Processing methods had no effect on the screw holding capacity, which was higher than the requirements for particleboard of medium density (Table 5.5). Effects of Binding Agent Compositions Two comparisons were made to determine significant effects of maleated polyolefin base resin types (PE vs. PP) (Table 5.5), and maleic anhydride content in maleated polypropylene (1.5% vs. 2.5% by weight)/molecular weight (52,000 vs. 47,000 g/mol) on the mechanical properties of the composites (Table 5.6). 140 Table 5.6. Effect of molecular weight/maleic anhydride content of MAPP on the mechanical properties of particleboard panels bound with maleated polypropylenes Mechanical Properties1 Panel Types MOR MOE 1B Strength Screw (MPa) (MPa) (MPa) Holding (N) Medium Density 11.0 — 16.5 1725 - 2750 0.40 - 0.55 900 — 1100 Gradesl Experimental Panels2 MAPP 03003 30.04 i 6.6 A 3582 i 567 A 0.43 2: 0.19 A 1469 i 330 A unextruded MAPP 03015 30.24 i 8.6 A 3586 i 698 A 0.36 i 0.16 A 1445 i157 A unextruded MAPP A 3-3015 1986:49B 2663 :270B 1.60 :0.64B 1552 i193 extruded MAPP B A @3003 23.0 0: 4.7 B 2875 i 347 B 1.50 _+_ 0.29 1580 i 299 extruded lProperty requirement data is from standard ANSI A208.1- l999-Particleboard. 2The means with different letters indicate significance between treatments at the or = 0.05 level, while the means with the same letter indicates no difference between treatments. 141 For the comparison of PE vs. PP-based binding agents, the specific comparisons made in Table 5.5 were rows 1 vs. 3 and rows 2 vs. 4. In this way, processing condition was held constant, while the polymer base resin was varied. Composite panels with polypropylene-based binding agents outperformed their polyethylene counterparts in stiffness (MOE), regardless of manufacturing method used, mainly due to the higher stiffness of polypropylene in the binding agent (Table 5.5). However, the strength of the composites (MOR) was not affected by the type of base resin of the maleated polyolefin since both PE and PP behave similarly, regardless of the processing method. Conversely, panels made with maleated polyethylene outperformed those made with maleated polypropylene in IB strength, likely due to the lower melting temperature of polyethylene. Lower melting temperature would allow » the polyethylene-based compound to flow to a greater extent even into the center region of the panels, causing stronger internal bonding. The screw holding capacity of the panels was not affected by the base resin type in maleated polyolefin. Differences between the two maleated polypropylene compounds were not significant for any of the mechanical properties tested (Table 5.6). Specific comparisons made were row 1 vs. 2 and row 3 vs. 4, where processing condition was held constant while the molecular weight/maleic anhydride content of MAPP was varied. The weight average molecular weights of the two MAPPs differed only by 5,000 g/mol and the difference in maleic anhydride content was 1% between the two. These polymers may have been similar enough that they did not create significant differences in the composite panel properties. 142 Comparison with Standard ANSI A208.] The MOR, IB strength and screw holding capacity results for experimental panels manufactured in this study indicated that the standard requirements were met or surpassed for all grades of particleboard of medium density when the particles were extruded before pressing (Tables 5.5-5.6). Without the extrusion step, the 18 strength was within the required range with maleated polypropylene, and surpassed when PE- based binding agent was used. MOE data were below the standard requirements for stiffness when MAPE was used with extrusion, but the panels with unextruded wood particles bonded with MAPE surpassed the stiffness requirements. Additionally, when MAPP was used, the panels exceeded the requirements for all grades of particleboard of medium density, regardless of processing conditions. 143 CONCLUSIONS This study examined the differences between panels made with MAPE and MAPP binding agents, specifically the contrasts of: (i) base resin type, PE or PP, and (ii) molecular weight/maleic anhydride content in MAPP binding agents, along with the effects of the manufacturing process. Particles were either pre-reacted with maleated binding agents in the extruder and then compression molded in a two-step process, or directly compression molded without the extrusion step. Surface characterization of unmodified and modified wood particles after reactive extrusion with maleated polyolefins showed similar changes in the wood particle surfaces after reaction, regardless of maleated polyolefin used. This evidence suggested that the wood particles can undergo the same reaction in a one-step process, without any obvious differences in reactivity. Differences in mechanical properties of the panels were correlated to the type of base resin used in the binding agent and to panel density profile. Results showed that while extruding the particles before panel pressing gave better overall internal bond strength, superior bending properties were obtained through compression molding alone. MAPP based panels outperformed MAPE based panels in stiffness, likely due to the higher stiffness of the PP base resin. MAPE enhanced the IB strength compared to MAPP, attributed to better melting and flow of the polyethylene. Polymer base resin had no effect on MOR or screw holding capacity. Differences between the two maleated polypropylene compounds were not significant for any of the mechanical properties tested. 144 The study also showed that a new type of environmentally friendly wood composite product could be formed from modified wood particles, regardless of processing conditions. This composite contained no formaldehyde-based adhesive, but still performed very well in mechanical tests, in many cases exceeding the standard requirements for particleboard of medium density. 145 10. 11. REFERENCES Matsuda, H., "Preparation and utilization of esterified woods bearing carboxyl groups," Wood Science and Technology, 21 (1): 75-88 (1987). Hon, D.N.S. and Xing, L.M., "Thermoplasticization of wood. Esterification," in Viscoelasticity of Biomaterials, W.G. Glasser and H. Hatakeyama, Editors. American Chemical Society: Washington DC. p. 118-132 (1992). Clemons, C., Young, R.A., and Rowell, R.M., "Moisture sorption properties of composite boards from esterified aspen fiber," Wood and Fiber Science, 24 (3): 353-63 (1992). Marcovich, N.E., Reboredo, M.M., and Arangguren, M.I., "Sawdust modification. Maleic anhydride chemical treatment," Holz als Roh- und Werkstojf, 54 (3): 189- 193 (1996). Marcovich, N.E., Aranguren, M.I., and Reboredo, M.M., "Modified woodflour as thermoset fillers Part I. Effect of the chemical modification and percentage of filler on the mechanical properties," Polymer, 42 (2): 815-825 (2001). Marcovich, N.E., Reboredo, M.M., and Aranguren, M.I., "Modified wood flour as thermoset fillers. 11. Thermal degradation of wood flours and composites," T hermochimica Acta, 372 (1-2): 45-57 (2001). Matsuda, H., Ueda, M., and Mori, H., "Preparation and crosslinking of oligoesterified woods based on maleic anhydride and allyl glycidyl ether," Wood Science and Technolog, 22 (1): 21-32 (1988). Matsuda, H., "Thermal plasticization of lignocellulosics for composites," in Emerging Technologies for Materials and Chemicals from Biomass R.M. Rowell, T.P. Schultz, and R. Narayan, Editors. American Chemical Society: Washington, DC. p. 98-114 (1992). Timar, M.C., Mihai, M.D., Maher, K., and Irle, M., "Preparation of wood with thermoplastic properties. Part 1. Classical synthesis," Holzforschung, 54(1): 71- 76(2000) Timar, M.C., Maher, K., Irle, M., and Mihai, M.D., "Preparation of wood with thermoplastic properties. Part 2. Simplified technologies," Holzforschung, 54 (1): 77-82 (2000). Timar, M.C., Maher, K., Irle, M., and Mihai, M.D., "Thermal forming of chemically modified wood to make high-performance plastic-like wood composites," Holzforschung, 58 (5): 519-528 (2004). 146 12. 13. 14. 15. 16. 17. 18. 19. Carlbom, K. and Matuana, L.M., "Composite materials manufactured from wood particles modified through a reactive extrusion process," Polymer Composites, 26 (4): 534-541 (2005). Carlbom, K. and Matuana, L.M., "Functionalization of Wood Particles through a Reactive Extrusion Process," Journal of Applied Polymer Science: In press (2006) ASTM, D 1037-99, Standard Methods of Evaluating the Properties of Wood- Based Fiber and Particle Panel Materials. ASTM: West Conshohocken (1999). ANSI, A208.1-1999, Particleboard. The Composite Panel Association: Gaithersburg (1999). Ratner, 8D. and Castner, D.G., "Electron Spectroscopy for Chemical Analysis," in Surface Analysis: The Principal Techniques, J .C . Vickerman, Editor. John Wiley & Sons: Chichester, England. p. 43-98 (1997). Schulte, M. and F ruhwald, A., "Some investigations concerning density profile internal bond and relating failure position of particle board," Wood Science and Technology, 54: 289-294 (1996). Wong, E.D., Zhang, M., Wang, Q., and Kawai, 8., "Formation of the density profile and its effects on the properties of particleboard," Wood Science and Technology, 33 (4): 327-340 (1999). Forest Products Laboratory, Wood Handbook: Wood as an Engineering Material. Forest Products Society: Madison, WI (1999). 147 CHAPTER 6 MODELING AND OPTIMIZATION OF F ORMALDEHYDE-FREE WOOD COMPOSITES USING A BOX-BEHNKEN DESIGN This chapter has been accepted for publication in Polymer Composites (February 2006). It is co-authored by K. Carlbom and L. M. Matuana 148 ABSTRACT A response surface model using a Box-Behnken design was constructed to statistically model and optimize the material compositions-processing conditions- mechanical property relationships of formaldehyde-free wood composite panels. Three levels of binding agent content, pressing time and press temperature were studied and regression models were developed to describe and optimize the statistical effects of the formulation and processing conditions on the mechanical properties of the panels. Linear models best fit both the flexural strength (MOR) and internal bond (IB) strength of the panels. Increasing any of the manufacturing variables resulted in greater MOR and IB strength. F lexural stiffness (MOE) was best described by a quadratic regression model. Increased MOE could be obtained through higher pressing times, binding agent concentrations and/or pressing temperatures. However, binding agent concentration had less effect on increasing the MOE at higher pressing temperatures. Numerical optimization showed that formaldehyde-free panels with desired mechanical properties could be manufactured at pressing temperatures ranging from 187.18—199.97°C, pressing time from 3.31—8.83 minutes, and binding agent concentration from 7.66—11.86%. 149 INTRODUCTION The recent implementation of more stringent emissions regulations for manufacturing facilities that produce particleboard and other wood composite panels (1) has spurred research into alternative binding technologies that do not involve formaldehyde. Our prior work demonstrated the efficiency of the reactive extrusion process as a way to graft maleated polyolefins onto the surface of wood particles, which could then be bound together without any additional adhesive (2-4). A new type of environmentally-friendly wood composite product was manufactured from the modified wood particles. This composite contained no formaldehyde-based adhesive, but still performed very well in mechanical tests, in some cases exceeding the standard requirements for particleboard of medium density. In previous work, the effects of extrusion processing conditions on the grafting of maleated polyolefins to wood particles were investigated, but work is still needed to determine the relationships between the mechanical properties of the formaldehyde-free composite panels and the material composition and panel manufacturing variables (2-4). The quality of wood composite materials such as particleboard depends heavily on material composition and manufacturing condition variables such as resin content, pressing temperature and pressing time (5). When multiple variables are involved, it becomes difficult to study the system using the common approach of varying only one factor at a time, while holding the others constant. This approach is not only time- consuming, but can also be costly and does not easily identify all of the interactions between factors (6, 7). A more efficient way to investigate these systems is to develop a mathematical model describing the relationship between the response and independent 150 variables, in which the significance of individual factors and multi-factor interactions can be determined (6-9). A Box-Behnken design (BBD) is a versatile method to statistically model and optimize response variables that are affected by multiple independent factors. Compared to full factorial or central composite designs, the BBD requires fewer trials and can efficiently model quadratic or higher-order relationships. Because of these features, the BBD has been used to study a variety of wood composite products (10-13). In this study, the effects of binding agent content, pressing time and pressing temperature on the mechanical properties of the resulting panels were evaluated using a Box-Behnken design to statistically model the system. Mechanical properties studied included modulus of rupture (MOR), modulus of elasticity (MOE) and internal bond (IB) strength, all of which are key parameters for assessing panel quality. Numerical optimization was performed in order to determine the best conditions for the manufacture of formaldehyde-free wood composite panels. 151 EXPERIMENTAL Materials Maple wood particles of 425 micron (40-mesh) size used in this study were supplied by American Wood Fibers (Schofield, WI). Hydrated zinc acetate, the catalyst, was obtained from Baker Analytical Reagents (JT Baker Co., Phillipsburg, NJ). Maleated polypropylene (MAPP or G-3003) supplied by Eastman Chemical Co. (Kingsport, TN) was used as the binding agent. The MAPP had a weight average molecular weight (MW) of 52,000 g/mole, approximate viscosity of 60,000 cP at 190°C and maleic anhydride content of 1.5% by weight. Experimental Design The Box-Behnken design (BBD) is a three-level design based upon the combination of two-level factorial designs and incomplete block designs (6, 7). BBDs are spherical designs, with the design points for high and low levels located at an equal distance from the center of the design. These designs have excellent predictability within the spherical design space and require fewer experiments than full factorial designs or central composite designs (CCDs) with the same number of factors. For example, an investigation with three factors would require at least 27 experiments in a full factorial design, 15 experiments in a CCD, or 13 experiments in a BBD, with additional replicates of the center point as necessary in each design to estimate experimental error. Additionally, BBDs are rotatable or nearly rotatable regardless of the number of factors studied (6, 7). 152 All statistical analysis, modeling and numerical optimization was performed using Design Expert software, v.6 (Stat-Ease, Inc. Minneapolis, MN). The BBD matrix generated by Design Expert software displays factor levels in the experimental design in two ways (i) the actual factor levels, which are the values from the experiment, and (ii) the coded factor levels, +1, -1, and 0, for high levels, low levels, and center point, respectively. Coded factor levels are defined as: Coded factor levels = Actual value - Factor mean (1) (Range of factorlal values/2) The BBD experimental design matrix is shown in terms of both actual and coded factor levels in Table 6.1. Twelve replicates were run for each experiment. 153 Table 6.1. Box-Behnken Design Matrix in terms of Both Actual and Coded Factor Levels Generated by Design Expert Software. Factors Effifgaim goint A: Press B: Press Time C:CBinding Agent ype Temperature (°C) (min) onchrtlon 1 IBFact 160 (-1) 6 (0) 18 (+1) 2 Center 180 (0) 6 (0) 10.5 (0) 3 IBFact 180 (0) 3 (-l) 18 (+1) 4 IBFact 160 (-1) 9 (+1) 10.5 (0) 5 IBFact 160 (-1) 6 (0) 3 (-1) 6 IBFact 180 (0) 9 (+1) 18 (+1) 7 IBFact 180 (0) 3 (-l) 3 (-1) 8 IBFact 200 (+1) 9 (+1) 10.5 (0) 9 IBFact 200 (+1) 6 (0) 3 (-1) 10 IBFact 200 (+1) 6 (0) 18 (+1) 11 IBFact 160 (-l) 3 (-1) 10.5 (0) l2 IBFact 180 (0) 9 (+1) 3 (-1) 13 IBFact 200 (+1) 3 (-1) 10.5 (0) 154 Compounding and Panel Manufacture The wood particles were dried for 48 hours at 105°C to a final moisture content of less than one percent before processing. A 10-liter high intensity mixer (Papenmeier TGAHKZO) was used for dry blending of the wood particles, binding agent, and catalyst. Levels of binding agent were varied as indicated in Table 6.1, based on the oven dry weight of wood flour. Zinc acetate esterification catalyst was added at 1% of the binding agent weight in all cases. All components were combined in the mixer and blended for 10 minutes at room temperature. Following blending, reactive extrusion was used to induce the esterification reaction between maleated polypropylene and wood particles. This was achieved as follows: the compounded wood particles were fed into a 32 mm conical counter rotating twin-screw extruder (C. W. Brabender Instruments, Inc.) with an L/D ratio of 13:1, driven by a 7.5 hp Intelli-Torque Plasti-Corder Torque Rheometer®. The barrel temperatures for all three zones inside the extruder were set at 165°C, and the rotational speed of the screws was held at 60 rpm. Once extruded, the wood particles were compression molded into panels using 3.4 MPa of pressure at various pressing times and temperatures set at levels indicated in Table 6.1. After pressing, panels were removed from the press and cooled at room temperature under compression for 15 minutes. Panel dimensions were 380 by 380 by 6 mm, with a target density of 750 kg/m3. 155 Property Testing Three-point flexural and internal bond (IB) strength tests were performed on an Instron 4206 testing machine (using Series IX software) in accordance with procedures outlined in ASTM standard D1037-99 (14). The crosshead speeds were 3.05 mm/min and 8.13 mm/min for flexural and IB tests, respectively. Values for modulus of rupture (MOR), modulus of elasticity (MOE) and IB strength were compared with values listed for particleboard of medium density in the standard ANSI A208.1-1999 (15). 156 RESULTS AND DISCUSSION Mechanical Properties To determine whether experimental panels would conform to standard strength requirements for conventional particleboard, mechanical property data was compared to ANSI standard requirements for particleboard. Table 6.2 lists the MOR, MOE, and IB strength requirements for particleboard of medium density, ranging from 640-800 kg/m3. In previous studies, composite panels manufactured through this process were within this range, with average densities ranging from 775-782 kg/m3 (2, 4). There are four grades of particleboard of medium density, all of which can be made with either interior or exterior adhesives. Grades M-1 and M-S are commercial grade boards, while M-2 and M-3 are intended for industrial use. Results of MOR, MOE and IB strength tests revealed a large range of property values for the experimental panels (Table 6.2). At the lowest levels of pressing time, temperature and binding agent concentration, panels were below the standard property requirements. However, at the highest levels of these factors, most panels met the standard requirements for particleboard of medium density. Several panels even exceeded the standard requirements for various grades of particleboard. These results are significant because particleboard is currently manufactured with formaldehyde-based adhesives. The formaldehyde-free wood composites manufactured in this study are more environmentally friendly and often outperform the requirements listed in the standard ANSI A208.]. 157 Table 6.2. Standard Property Requirements for Various Grades of Particleboard of Medium Density (640-800 kg/m3) Gradesl Expleglrlzgtal Properties M-l M-S M-2 M-3 Property Range MOR (MPa) 11.0 12.5 14.5 16.5 0.48 — 17.91 MOE (MPa) 1725 1900 2250 2750 101.2 — 2556 IB (MPa) 0.40 0.40 0.45 0.55 0.04 — 2.69 ‘From Standard ANSI A208.1-1999 Particleboard. 2Experimental values are from panels manufactured under the various factor combinations specified in Table 1. Minimum values are from panels pressed at 160°C for 6 minutes with 3% MAPP, while maximum values are from panels pressed at 200°C for 6 or 9 minutes with 10.5 or 18% MAPP. 158 Statistical Analysis of the Model Regression analysis was performed on the mechanical property results in order to develop best-fit models for the experimental data. A separate regression analysis was run for each of the three mechanical properties studied to determine the relationships between mechanical properties and binding agent content, pressing time and pressing temperature. Modulus of Rupture (MOR) and Internal Bond (IB) Strength For both MOR and IB strength, regression analysis of the experimental data showed that a linear model best fit the relationship between the response (MOR or IE) and the binding agent content, pressing time and pressing temperature. In both cases, a square root transformation was applied to the data in order to normalize the variance of the residuals. These models, described in terms of coded factors, were described as: 4MOR = +2.75 + 0.78A + 0.23B + 0.91C (2) and 4/IB strength = +1 + 0.30A + 0.084B + 0.54C (3) where A is press temperature, B is press time, and C is binding agent concentration. For both MOR and IB strength, only the three main factors were significant, with no interactions found between them. Since the equations are displayed in terms of coded factors, the relative effect of each variable on the response can be evaluated by comparing the absolute value of its coefficient and its algebraic sign. As shown in 159 Equations 2 and 3, the effects of all of the factors had positive algebraic signs, which indicated that increasing any one would increase theMOR or IB strength. In both cases, concentration of binding agent (factor C) had the largest effect on the mechanical properties, followed by pressing temperature (factor A) and pressing time (factor B). This is illustrated by a cube graph in Figure 6.1(a) for MOR and 6.1(b) for IB. Increasing the amount of binding agent resulted in better adhesion between the wood particles as a result of having more polymer chains attached to the wood particles which would entangle to hold the panel together. Similarly, an increase in temperature would allow more thorough melting and flow of the polymer chains attached to the wood particles, thus increasing the efficiency of binding. Although longer pressing time did increase both MOR and IB, the effect was much smaller which indicated that pressing time was much less important in improving the MOR and IB than binding agent concentration or press temperature. 160 (a) Sqrt( MOR) 3.11 4.68 B+ 1.29 2.85 G) E l:- g 2.64 4.21 C+ o. in C: Concentration B- 0.82 2.38 C- A- A+ A'Temperature (b) Sqrt(lB) 1.32 1.92 B+ 0.24 0.85 (D E 1: 8 9 1.15 1.75 C+ a. 63' C: Concentration B- 0.08 0.68 C- A- A+ A‘ Temperature Figure 6.1. Cube graphs of the linear relationship between mechanical property results and press temperature, pressing time, and binding agent concentration for (3) MOR and (b) IB strength. 161 Modulus of Elasticity (MOE) Results of the regression analysis for MOE suggested that the data was best fit by a quadratic model. A. square root transformation was also applied to the data for MOE in order to normalize the variance of the residuals in the system. Table 6.3 shows the analysis of variance (ANOVA) results for the response surface quadratic model. As listed in the table, some of the factors were not significant (Probability > F greater than 0.0500), so the model was reduced by removing the insignificant factors. The reduced model, containing only significant terms and described in terms of coded factors is: JMOE = + 33.49 + 9.47A + 3.153 + 9.99C +1.14A2 + 2.7032 - 2.44C2 - 2.02AC (4) For MOE, all main effects (A, B and C) were significant, along with the second- order main effects (A2, B2 and C2), and the AC interaction. Because A, B and C had positive algebraic signs, an increase in temperature, pressing time or binding agent concentration would improve the MOE. By contrast, the two-factor AC interaction had a negative algebraic sign, which indicated a negative effect on MOE. 162 Table 6.3. Analysis of Variance (ANOVA) for Response Surface Quadratic Model Source 88:31:; lifrgefifslzf S1232: F-Value Prob. > Fl Quadratic Model 1671.35 9 185.71 379.34 < 0.0001 A 716.70 1 716.70 1463.99 < 0.0001 B 79.35 1 79.35 162.08 < 0.0001 C 799.04 1 799.04 1632.19 < 0.0001 A2 5.49 1 5.49 11.21 0.0123 B2 30.59 1 30.59 62.49 < 0.0001 C2 25.00 1 25.00 51.07 0.0002 AB 0.58 1 0.58 1.18 0.3137 AC 16.30 1 16.30 33.30 0.0007 BC 0.89 1 0.89 1.83 0.2187 1Values of “Prob. > P” less than 0.0500 indicate model terms are significant. In this case A, B, C, A2, B2, C2 and AC are the significant model terms. 163 The perturbation plot of the square root of MOE against temperature, pressing time and binding agent concentration shows the contribution of each factor to the MOE (Figure 6.2). The perturbation plot illustrates the changes in MOE as each factor moves from the chosen reference with all other factors held constant at the middle level of the design space (6, 7). Pressing temperature and binding agent content (factors A and C) are shown to have the largest effect on the MOE, while pressing time (factor B) shows only a small effect on MOE. Because two of the main factors (A and C) that increase the MOE are part of a significant interaction, it would not be appropriate to investigate these factors separately. The effect of one factor will depend upon the level of the other, since the interaction is significant. 164 Figure 6.2. Sqrt(MOE) 51— A 40— L 30— 20— 10 l l l l l -1.0 -05 0.0 0.5 1.0 Deviation from Reference Point Perturbation plot of square root of MOE against pressing temperature (A), pressing time (B) and binding agent concentration (C). 165 Figure 6.3 shows the interaction graphs of the change in MOE as a function of the temperature-concentration interaction. The effect of the temperature-concentration interaction is illustrated at the lowest pressing time, 3 minutes, in Figure 6.3(a) and at the highest pressing time, 9 minutes, in 6.3(b). A small increase in MOE resulted from increased pressing time, regardless of the temperature or binding agent concentration. However, this effect was quite small, resulting in gains of only 3-8 MPa to the MOE over the range studied. The binding agent concentration had greater effect on MOE when the pressing temperature was the lowest. At higher temperatures, the differences between low and high binding agent concentration were reduced. This can be verified by comparing the magnitude of the difference between the lines for high and low binding agent concentration. At the lowest temperature, the difference in MOE was large, but was reduced as the temperature increased. This reduced efficiency of binding agent concentration to increase the MOE at higher temperatures accounts for the negative effect of the temperature-binding agent interaction, which was shown by the negative algebraic sign on AC in Equation 4. The effect of this interaction is that less binding agent is required at higher temperatures to produce panels with sufficient MOE values. At lower temperatures, more binding agent would be required to manufacture composites with the same range of MOE values. A greater degree of melting and flow of the polymer attached to the wood particles likely occurs at higher temperatures, which would allow a smaller amount of binding agent to be as more efficient in creating particle-particle bonds. 166 Figure 6.3. 60- (a) 18% 47- \ ft? 3 34 E’ O' a) 21-1 3% 3.. l I T l l 160 170 180 190 200 Temperature °C 60- (b) 18% 47- 1’1? 2 3 4-1 E" U m a- l l I I I 160 170 180 190 200 Temperature °C Interaction plots of the variation in square root of MOE as a function of the interaction between pressing temperature and binding agent concentration at (a) low press time (3 minutes) and (b) high press time (9 minutes). 167 Numerical Optimization of Mechanical Properties The numerical optimization function in Design Expert software was used to determine combinations of binding agent concentration, panel pressing time and temperature that would result in the most favorable mechanical properties of the resulting composites. The numerical optimization function of Design Expert is based on the desirability function developed by Derringer and Suich (16), which transforms each response value to a desirability index (di). Each desirability index is defined by three parameters: goal, lower and upper, and the program allows the d, goal parameter to be to one of five options: minimum, maximum, target, in range, or equal to. Once these settings have been defined, d; varies between zero (worst case) and one (ideal case). Design Expert searches for the largest overall di and presents a series of solutions which best maximize the di. Table 6.4 lists the optimization criteria settings used to optimize the mechanical properties of the composite panels. In order to produce a composite panel that had adequate strength (MOR), stiffness (MOE), and IB strength, the goal for each property was initially set in range based upon the property requirements of the particleboard standard ANSI A208.] (Table 6.2). For example, square root of MOR was set in range 3.32—4.06 MPa, MOE was set in range 41 .53—52.44 MPa and IB strength was set in range 0.63—0.74 MPa (15). However, the data for IB strength showed that the standard property requirements were met and even exceeded at very low concentrations of binding agent. Since the IB would be higher than required at nearly all combinations of pressing temperature, panel pressing time and binding agent concentration, the IB criteria was set in range over all experimental values instead of in range of the property requirement 168 values. In this way, the MOR and MOE could be targeted to the necessary values, with the understanding that IE strength would likely be much higher than required. Additionally, because we wanted to determine which combination of material composition and processing conditions would produce panels with the desired properties, the optimization criteria for all of these factors were set in range over the entire set of studied values. For this analysis, the desirability functions of the mechanical properties were set as follows: (i) if JMOR < 3.32 or JMOR > 4.06 MPa, and x/MOE < 41.53 or J MOE > 52.44 MPa then (I, = 0 (worst case) (ii) if 3.32 S «lMOR S 4.06 and 41.53 S \lMOE S 52.44, then (1, = 1 (ideal case) All other optimization parameters were set to the default settings of one for the weights of upper and lower limits for the input factors, and three for the importance (3 relative scale that weights each of the resulting dis in the overall desirability product). Ten cycles were run per optimization, with the epsilon value for the minimum difference in eliminating duplicate results set at its default value. 169 Ff. l .11 iii. oo: o: 2.: Sam NEW cfiw :43— o co: em: ”5:4 nmd mam: m3. Neda— w co: em: 3.? owm 3.: no.» _méo— n co: mm: cart? ova 3.: 3.x 2.3— o co: m _ ._ code mm.m mod 36 mada— m co: om: mode 3am mm: _ Sam 9462 :0 co: mm: 3.? om.m 2.: _m.m 3.93 m co: N. _ ._ no.3 N.Mm obs 3.x 360— N oo: o: 3.34 mm.m cos 2.» 3.3. _ 3:6 :55 3:»: 3L 5:838:60 :55 Ge E328: Eat—em Eozvtcm 226:5 so? “seam 65.9 8o...— 23885oh 8:28 mucus—em 83330 m _ _ 8.. one owes 5 Es: 83:5 m _ _ 3.2 mm. _ a owes 5 3:6 Amozeem m _ _ 2:. mg once 5 And—2:102:23 m _ _ w: m awash E ax; 5:828:00 Eow< wEvEm m _ _ o m emcee E AEEV oEC. @535 m _ _ com of 098.. E Gov 23anth wfimmoi sue? an?» agar—3:: .59:— 333 :85 ~23: :85 33:4— 130 .885— mafiahmacv 8:27.3— mzamom can mwfifiom 5:23.530 _motofizz 66 03mm. 170 Numerical optimization results produced 9 optimum solutions, all with desirability of 1.00. These results are displayed in Table 6.4. As listed in the table, pressing temperatures ranged from 187.18—199.97°C, pressing time ranged from 3.31— 8.83 minutes, and binding agent concentration from 7.66—11.86%. These results showed that formaldehyde-free wood composite products could be made under a variety of conditions, many of which are not all that different than those used to manufacture conventional particleboard. In addition, a desired solution could be employed if the goal was to minimize a particular factor, such as pressing time, in order to reduce manufacturing costs. 171 CONCLUSIONS Three panel manufacturing variables (pressing temperature, pressing time and binding agent content) were analyzed by developing statistical models for their relationships with mechanical properties MOR, MOE and IB strength. Additionally, numerical optimization was used to determine the best conditions for manufacturing the panels in terms of the mechanical properties. The following conclusions were obtained: 1. Both MOR and IB were best fit with linear models. Increasing any of the manufacturing variables (pressing temperature, pressing time and binding agent content) resulted in greater MOR and IB strength. A quadratic model best fit the MOE. Increased MOE could be obtained through longer pressing times, but the effect of this variable was small. Larger effects on MOE were obtained by increasing the binding agent concentrations and pressing temperatures. However, the negative interaction between temperature and binding agent concentration indicated that binding agent concentration had less effect on increasing the MOE as the pressing temperature increased. Panels with mechanical property values in the range of the mechanical property requirements for particleboard of medium density could be produced using several pressing temperature-pressing time-binding agent concentration combinations with maximum desirability. This showed that formaldehyde-free panels could be produced at similar adhesive contents and processing conditions used for conventional glued particleboard made with formaldehyde-based adhesive. 172 10. 11. 12. REFERENCES US-EPA, "National emission standards for hazardous air pollutants: plywood and composite wood products," Federal Register, 69 (146): 45943-46046 (2004). Carlbom, K. and Matuana, L.M., "Composite materials manufactured from wood particles modified through a reactive extrusion process," Polymer Composites, 26 (4): 534-541 (2005). Carlbom, K. and Matuana, L.M., "Functionalization of Wood Particles through a Reactive Extrusion Process," Journal of Applied Polymer Science: In press (2006) Carlbom, K. and Matuana, L.M., "Influence of Processing Conditions and Material Compositions on the Performance of Formaldehyde-Free Wood-Based Composites," Polymer Composites: In press (2006). Maloney, T.M., Modern Particleboard and Dry-Process F iberboard Manufacturing, Updated Ed., Miller Freeman: San Francisco (1993). Myers, RH. and Montgomery, D.C., Response Surface Methodology: Process and Product Optimization Using Designed Experiments. John Wiley & Sons, Inc.: New York (1995). Montgomery, D.C., Design and Analysis of Experiments, 5th Ed., John Wiley & Sons, Inc.: New York (2001). Matuana, L.M. and Mengeloglu, F ., "Manufacture of rigid PVC/wood-flour composite foams using moisture contained in wood as foaming agent," Journal of Vinyl & Additive Technology, 8 (4): 264-270 (2002). Matuana, L.M. and Li, Q., "Statistical modeling and response surface optimization of extruded HDPE/wood-flour composite foams," Journal of Thermoplastic Composite Materials, 17 (2): 185-199 (2004). Barry, A., Lepine, R., Lovell, R., and Raymond, 8., "Response surface methodology study of VOCs in plywood press emissions," Forest Products Journal, 51 (1): 65-73 (2001). Barry, AD. and Comeau, D., "Volatile organic chemicals emissions from OSB as a function of processing parameters," Holzforschung, 53 (4): 441-446 (1999). Barry, A., Comeau, D., and Lovell, R., "Press volatile organic compound emissions as a function of wood particleboard processing parameters," Forest Products Journal, 50 (10): 35-42 (2000). 173 13. 14. 15. 16. Park, B.D., Riedl, B., Hsu, E.W., and Shields, J ., "Hot pressing process optimization by response surface methodology," Forest Prod. J., 49 (5): 62-68 (1999) ASTM, D 103 7-99, Standard Methods of Evaluating the Properties of Wood- Based Fiber and Particle Panel Materials. ASTM: West Conshohocken (1999). ANSI, A208.1-1999, Particleboard. The Composite Panel Association: Gaithersburg (1999). Derringer, G. and Suich, R., "Simultaneous optimization of several response variables," Journal of Quality Technology, 12 (4): 214-219 (1980). 174 CHAPTER 7 SUMMARY OF FINDINGS Public awareness about formaldehyde and other toxic chemicals being released to the environment has helped to drive new government standards for wood composite products. As environmental regulations become more stringent, the need for formaldehyde-free adhesive systems for wood composite products will likely increase. The main goal of this work was to study the concept of using a reactive extrusion process as a means of developing a new, formaldehyde-free binding system for wood composite products. The following specific objectives were accomplished to achieve the main goal of this project: 1. Evaluate the effects of material composition (binding agent types and content) and extrusion processing conditions (temperature profile and rotational screw speed) on the level of grafting (surface properties) of wood particles alter the reactive extrusion process; 2. Characterize the surface of unmodified and modified wood particles in terms of chemical compositions (both elemental and functional groups); 3. Manufacture composites and evaluate their physico-mechanical properties; 175 4. Establish the relationships between material composition, processing conditions, and physico-mechanical properties in order to identify the major factors that govern the performance of formaldehyde-free wood composite products manufactured through a reactive extrusion process. The following sections relate the findings of this work to these objectives. Objectives 1 and 2: Results of FTIR, 13C NMR, XPS and titration analysis verified the reaction between wood particles and maleated polyolefins. This proved that the maleated polyolefins could be successfully grafted to wood particles using a reactive extrusion process, without the use of any solvents. The esterification reaction between wood particles and MAPE was found to be a function of the MAPE concentration used to modify the wood particles. The grafting reaction produced mostly the diester form of the modified wood particle during reactive extrusion. However, no significant difference was found in grafting efficiency of the modified wood particles at different extrusion processing conditions. Changing the extruder’s barrel temperature profile (BO-180°C) and/or its rotational screw speed (20-80 rpm) resulted in adequate grafting, which indicated that the esterification reaction was not a function of processing conditions. Regardless of MAPP molecular weight (from a low of 11,200 to a high of 52,000 g/mol) all investigated MAPP compounds were effective in changing the surface of wood particles after modification, compared to unmodified wood particles. However, no distinct trend was observed between molecular weight of MAPP and grafting efficiency through a reactive extrusion process. Regardless of molecular weight, it is believed that 176 the high content of MAPP (20%) used in this study prevented the detection of differences in the grafting efficiency because the maximum level of grafting reaction had already occurred with each component at 20% MAPP content. The reactive extrusion process was found to be a suitable way to modify wood particles with maleated polyolefins as it worked quickly and without the use of solvents. This process would be industrially friendly and would allow large quantities of modified wood particles to be produced for the manufacture of a formaldehyde-free wood composite product. Objective 3: Modified wood particles made with MAPE were successfully formed into formaldehyde-free composite panels without the use of any additional adhesive. The composite contained no formaldehyde-based adhesive, but still performed very well in mechanical tests, in many cases exceeding the standard requirements for particleboard of medium density. Additional work focused on the differences between panels made with MAPE and MAPP binding agents, specifically the contrasts of: (i) base resin type, PE or PP, and (ii) molecular weight/maleic anhydride content in MAPP binding agents, along with the effects of the manufacturing process. Particles were either pre-reacted with maleated binding agents in the extruder and then compression molded in a two-step process, or directly compression molded without the extrusion step. 177 Differences in mechanical properties of the panels were correlated to the type of base resin used in the binding agent and to panel density profile. Results showed that while extruding the particles before panel pressing gave better overall internal bond strength, superior bending properties were obtained through compression molding alone. MAPP based panels outperformed MAPE based panels in stiffness, likely due to the higher stiffness of the PP base resin. MAPE enhanced the IB strength compared to MAPP, attributed to better melting and flow of the polyethylene. Polymer base resin had no effect on MOR or screw holding capacity. Differences between the two maleated polypropylene compounds were not significant for any of the mechanical properties tested. Objective 4: In a final phase of the study, three panel manufacturing variables (pressing temperature, pressing time and binding agent content) were analyzed by developing statistical models for their relationships with mechanical properties MOR, MOE and IB strength. Additionally, numerical optimization was used to determine the best conditions for manufacturing the panels in terms of the mechanical properties. Statistical modeling revealed that both MOR and IB were best fit with linear models. Increasing any of the manufacturing variables (pressing temperature, pressing time and binding agent content) resulted in greater MOR and IB strength. However, the MOE was best fit by a quadratic model. Increased MOE could be obtained through longer pressing times, but the effect of this variable was small. Larger effects on MOE were obtained by increasing the binding agent concentrations and pressing temperatures. 178 However, the negative interaction between temperature and binding agent concentration indicated that binding agent concentration had less effect on increasing the MOE as the pressing temperature increased. Panels with mechanical property values in the range of the mechanical property requirements for particleboard of medium density could be produced using several pressing temperature-pressing time-binding agent concentration combinations with maximum desirability. This showed that formaldehyde-free panels could be produced at similar adhesive contents and processing conditions used for conventional glued particleboard made with formaldehyde-based adhesive. 179 Future Work This work resulted in the development of a new type of formaldehyde-free wood composite material that could be produced through a dry process. The surface composition of unmodified and modified wood particles and maleated polyolefins were studied in detail in order to verify that the reaction had occurred, and to determine the efficiency of the modification under several different material compositions and processing conditions. However, hardwood maple was used as the wood component in nearly all of the work. Only one set of pine panels was manufactured in order to determine a baseline comparison between wood species (see appendix). The effect of wood species, specifically hardwood vs. softwood, should be studied in greater detail to determine whether the wood type has an effect on the grafting efficiency during reactive extrusion. The presence of naturally occurring resins in some softwood species might have an effect on the grafting reaction, which may cause differences in the quality of panels produced with this method. A large portion of particleboard is currently made with softwood species such as pine, so this study would be relevant for industrial application of the technology. No differences were found in between modified wood particles produced under various extrusion processing conditions when the concentration of MAPE was held constant at 20%. The results of that portion of the study suggested that 20% by weight might be too much binding agent to see any differences that may exist under these conditions. In order to determine whether differences could be observed at lower binding agent concentrations, a similar investigation should be run at a lower concentration, perhaps between 5-10% by weight. This would be below the level that was determined to 180 cause maximum grafting, around 15%. If differences in grafting efficiency were found with this lower concentration, the extrusion conditions could be modified for future work in order to produce the greatest amount of grafting in the modified particles. Since this work indicated that panels with properties within the standard requirement range could be made at much lower concentrations than the 20% used in early studies, this would ensure that the panel quality would be high when made with low binding agent concentrations. In addition, the moisture resistance of wood composite panels such as particleboard or MDF made with UF adhesive is generally very poor. Results of thickness swelling tests after a 24 hour cold soak (see appendix) indicated that formaldehyde-free composites had low swelling. Standard property requirement values for particleboard and medium density fiberboard vary depending upon grade and density, and not all grades specify thickness swelling requirements. Therefore, laboratory tests should be performed using commercial particleboard or MDF as a control to determine whether these formaldehyde-free panels have an advantage in moisture resistance. These panels may prove to be better at resisting moisture than the conventional glued particleboard and MDF, which would be another industrial selling point. Finally, the studies done to determine whether panels with suitable mechanical properties could be made through a one-step process (extrusion only) suggested that the IB strength of these panels was lower than those made through the two-step process. This was viewed as a drawback of this manufacturing method, even though the 1B strength was close to or within the standard property requirement range from ANSI. These panels were made at only one pressing condition, which had previously been determined to produce panels with good mechanical properties in the two-step method. 181 However, this pressing setup did not necessarily account for the extra time needed to properly melt the maleated polyolefin throughout the entire mat thickness and allow completion of the chemical reaction between the wood and maleated polyolefin during pressing. A study should be undertaken to optimize the pressing conditions for the one- step panels, as these are the most likely to be industrially viable considering start-up costs for producers that are currently making other wood-based products that are not extruded. If the IB strength could be improved through longer pressing, perhaps at lower temperatures or with a pre-pressing step under low pressure, these panels would be even more advantageous from an industrial perspective. 182 APPENDIX pommoonm Ho: n mam Ems: 688.er bacon 8:662 $2.38 < 52.4 unease. sec” 6882252358728 .8. 52.4 688% as”: 3o 8% 3A a: a: 83% av stoned: one 8: S: we we ovev xv 8855-3 85 84.8. in a: a: 82% av decoded: m; can 3% a: aa 8% a: 26:85-92 2.32 OVO mNh~ o.— — m..— we ma ma awakens—Hugh: mm m; an: on a: we oeev 8 NA: 26 cam on a: we oaev a: I: and 32 n2 8 a: 83% we 22 23 RS 2; we a: 83% me $2 23 82 as 8 a: 82% 2 mi 23 mm: 3 _ we we 83% we 72 nJ—uhacAo—UEQA— 3:5 3% Es: aamwwa aeafiwea. pane ass 25 .25 E mo: ~52 eofiefl. 55$ has: a85.3.; pfionconmm 3650 8:802 was Becca—omvam 8.: 3:688:93”: 43.22:on magnum .:.< 2an 183 3”.— moa new 90 a: v3. o 58 o .Xom o: mwom mdm Wm mm own 0 58 N. gem $3.2 cow—rang: me; am: m.w_ 0.5 E N: c E8 Q .o\oom 3d 33 32 md 2 m? o 58 a .o\com ooN 003 Wow 0m 9 wt. 0 58 n 3&8 on; omfl 5.: 5w mm o: o 58 m gem on; omfl 52 Ex mm o: o 58 n 6&3 mad 3: WQ 2 ow N: o 58 N. .o\°m_ $6 $3 M2: vm 8 NR o 58 5 £5— 3d 83 as cm 3. N: c 58 N. .32. omd mom 9N me 3 o: o 58 n .Xm mm<2 toe—Ema Ann—EV Ann—EV Ana—.6 9%“me netmmwmn< Ann—EV AEEV 25¢. .23.— m— HOE may: 3.05—93h. .333 57.59 305—3.; mg a? 230m 225 2%: _eEEcaxm é 98 €285 306282 .m.< osfl 184 _md 09% Sdm w mm own 0 23-0 med 23m nmdm o _m m: 0 83-6 ovd Sam No.2 fl _ m m w? o 0 ~ mmé No.0 72% o _ .S 3 f v2. 0 me .342 33:23:: am; mmom 0.2 c 3 NE. 0 23-0 om; chum odm n E mt. o 83-0 a; omom wm _ H M: o: o Smmé mvd wmmm 3 S E own 0 flim— mm<2 cue—flim— Aam—nzv WNW—RV ”Wu—WW wamwwsm aotmmwmn< ”WWW—a“ nah—“HM"; 33‘ .053 anus—93H .533 . . arm—<2 53> 950m 2053 2922 Equatoaxm Sm 3mm €3on $05382 .m.< 2an 185 200 um 200 um Figure A]. SEM images of fracture surfaces of panels manufactured from maple with maleated polypropylene, G-3003 at 250X magnification. (a) 3%, (b) 10.5% and (c) 18% MAPP. 186 100 um 100 um (C) 100 um Figure A.2. SEM images of fracture surfaces of panels manufactured from maple with maleated polypropylene, G-3003 at 500X magnification. (a) 3%, (b) 10.5% and (c) 18% MAPP. 187 Table A.4. Mechanical Property Data for Experimental Panels Bound with 10.5% MAPP, Pressed for 6 Minutes at 180°C and 3.4 MPa Pressure Wood S ecies Thickness Densi MOR MOE " 1mm) (kg/m (MPa) (MPa) Maple 6 774 i 6 7.6 i 2 1121 i 299 Pine 6 769 i 10 7.7 i 3 1445 i 264 188 111111111111111111