You are on page 1of 500

Adhesives for Wood

and Lignocellulosic Materials
Scrivener Publishing
100 Cummings Center, Suite 541J
Beverly, MA 01915-6106

Adhesion and Adhesives: Fundamental and Applied Aspects

of adhesion; modeling of adhesion phenomena; mechanisms of adhesion; surface


and interfacial analysis and characterization; unraveling of events at interfaces;

aspects in reinforced composites; formation, characterization and durability of


adhesive joints; surface preparation methods; polymer surface
biological adhesion; particle adhesion; adhesion of metallized plastics; adhesion of
adhesion;
superhydrophobicity and superhydrophilicity. With regards to adhesives, the Series
will include, but not limited to, green adhesives; novel and high-performance
adhesives; and medical adhesive applications.

Series Editor: Dr. K.L. Mittal


P.O. Box 1280, Hopewell Junction, NY 12533, USA
Email: usharmittal@gmail.com

Publishers at Scrivener
Martin Scrivener (martin@scrivenerpublishing.com)
Phillip Carmical (pcarmical@scrivenerpublishing.com)
Adhesives for Wood
and Lignocellulosic Materials

R. N. Kumar and A Pizzi


This edition first published 2019 by John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
and Scrivener Publishing LLC, 100 Cummings Center, Suite 541J, Beverly, MA 01915, USA
© 2019 Scrivener Publishing LLC
For more information about Scrivener publications please visit www.scrivenerpublishing.com.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or other-
wise, except as permitted by law. Advice on how to obtain permission to reuse material from this title
is available at http://www.wiley.com/go/permissions.

Wiley Global Headquarters


111 River Street, Hoboken, NJ 07030, USA

For details of our global editorial offices, customer services, and more information about Wiley prod-
ucts visit us at www.wiley.com.

Limit of Liability/Disclaimer of Warranty


While the publisher and authors have used their best efforts in preparing this work, they make no rep-
resentations or warranties with respect to the accuracy or completeness of the contents of this work and
specifically disclaim all warranties, including without limitation any implied warranties of merchant-
ability or fitness for a particular purpose. No warranty may be created or extended by sales representa-
tives, written sales materials, or promotional statements for this work. The fact that an organization,
website, or product is referred to in this work as a citation and/or potential source of further informa-
tion does not mean that the publisher and authors endorse the information or services the organiza-
tion, website, or product may provide or recommendations it may make. This work is sold with the
understanding that the publisher is not engaged in rendering professional services. The advice and
strategies contained herein may not be suitable for your situation. You should consult with a specialist
where appropriate. Neither the publisher nor authors shall be liable for any loss of profit or any other
commercial damages, including but not limited to special, incidental, consequential, or other damages.
Further, readers should be aware that websites listed in this work may have changed or disappeared
between when this work was written and when it is read.

Library of Congress Cataloging-in-Publication Data

ISBN 978-1-119-60543-0

Cover image: Pixabay.Com


Cover design by Russell Richardson

Set in size of 11pt and Minion Pro by Manila Typesetting Company, Makati, Philippines

Printed in the USA

10 9 8 7 6 5 4 3 2 1
This book is dedicated to my (Dr. R. N. Kumar) son,
the late Dr. Gopal Kumar who evinced keen interest in my
endeavor to write the book.
Contents

Preface xxi
Part A: Substrates, Adhesives, and Adhesion
1 Wood as a Unique Adherend 3
1.1 Introduction 3
1.2 Wood, An Adherend with Hierarchical Structure 3
1.3 Details of Structural Hierarchy in Wood 4
1.3.1 Physical Structure 4
1.3.1.1 Growth Rings and Ring-Porous
and Diffuse-Porous Wood 5
1.3.1.2 Wood Cells 6
1.3.1.3 Organization of Cell Walls in Wood 8
1.4 Chemical Composition 10
1.4.1 Cellulose 11
1.4.2 Hemicelluloses 12
1.4.3 Lignin 12
1.4.3.1 Lignin Isolation 14
1.4.3.2 Functional Groups in Lignin 15
1.4.3.3 Evidences for the Phenylpropane Units
as Building Blocks of Lignin 15
1.4.3.4 Dehydrogenation Polymer (DHP) 17
1.5 Influence of Hierarchical Structure of Wood
on Wood–Adhesive Interaction 18
1.5.1 Penetration 20
1.5.1.1 Penetration in Different Size Scales 20
1.5.2 Other Wood-Related and Process-Related Factors 22
1.6 Effect of Hierarchical Structure of Wood on Adhesive
Penetration 22
1.7 Wood Factors Affecting Penetration 24
1.8 Influence of Resin Type and Formulation on Penetration 25

vii
viii Contents

1.9 Effect of Processing Parameters on Penetration 26


References 27
2 Fundamentals of Adhesion 31
2.1 Introduction 31
2.2 Definitions 32
2.2.1 Adhesion 32
2.2.2 Cohesion 32
2.2.3 Adhesive 33
2.2.4 Adherend 33
2.2.5 Bonding 33
2.2.6 Adhesive, Assembly 33
2.3 Mechanism of Adhesion 33
2.3.1 Specific Adhesion 34
2.3.1.1 London Dispersion Force 35
2.3.1.2 Dipole–Dipole Interaction 36
2.3.1.3 Dipole–Induced-Dipole Interaction 36
2.3.1.4 Ion–Dipole Interaction 36
2.3.1.5 Hydrogen Bonds 38
2.3.1.6 Ionic Bonds 38
2.3.1.7 Chemical Bonds 38
2.4 Theories of Adhesion 38
2.4.1 Mechanical Theory 39
2.4.1.1 lllustration of Mechanical Adhesion
for Wood 40
2.5 Electronic Theory 41
2.6 Diffusion Theory 41
2.7 Adsorption/Covalent Bond Theory 42
2.8 Adhesion Interactions as a Function of Length Scale 43
2.9 Wetting of the Substrate by the Adhesive 43
2.10 Equilibrium Contact Angle 44
2.11 Thermodynamic Work of Adhesion 45
2.12 Spreading Coefficient 49
2.13 Zisman’s Rectilinear Relationship—Zisman’s Plots
and Critical Surface Tension of a Solid 50
2.14 Effect of Surface Roughness on Contact Angle 50
2.15 Weak Boundary Layer Theory 51
2.16 Measurement of the Wetting Parameters
for Wood Substrate 52
2.16.1 Some Results on Surface Energy of Wood 53
Contents ix

2.17 Covalent Bond Formation 56


References 56
3 Urea–Formaldehyde Resins 61
3.1 Introduction 61
3.2 Historical Review of UF Resins (Plastic Historical Society) 62
3.3 Reaction between Urea and Formaldehyde 62
3.4 Reaction Sequence 63
3.5 Manufacture of UF Resin 64
3.6 Chemistry of Reaction—Conventional Process
(Alkaline–Acid Process/Three-Step Process) 65
3.6.1 First Stage—Reaction under Alkaline Conditions 65
3.6.1.1 Reaction Mechanism 66
3.6.2 Second-Stage Condensation Reaction under Acid
Conditions: Chain Extension 66
3.6.2.1 Reaction Chemistry 67
3.6.2.2 Reaction Mechanism 67
3.6.3 Third Stage—Neutralization and Addition
of Second Urea 69
3.6.3.1 Reactions Involving Migration
of Hydroxymethyl Groups 69
3.7 Composition of the Commercial UF Resins 70
3.7.1 Monomeric Species 70
3.7.2 Oligomeric Species 70
3.7.3 General Structure of Commercial UF Resins 70
3.7.4 Urons 72
3.8 Reactions of UF during Storage 73
3.9 Reaction Parameters in the Production of Amino Resins
(General) 74
3.10 Four-Step Process for Low Formaldehyde Emission 74
3.11 Curing of UF Resins 75
3.11.1 Ammonium Salts 75
3.12 Cross-Linked Structure 76
3.13 Triazinone for Curing the UF Resin 77
3.14 Distinguishing Feature of UF from other Synthetic Resin
Adhesives such as MUF and PF 78
3.15 Other Curing Agents 78
3.16 Protic Ionic Liquids as a New Hardener-Modifier System 79
3.17 Improvement of Water Resistance and Adhesive
Performance of UF Resin 81
x Contents

3.18 Characterization of UF Resin 82


3.18.1 13C NMR Data 82
3.18.2 Free Formaldehyde Content in the Resin 83
3.18.3 Molecular Weight and Molecular Weight
Distribution 83
3.18.4 Size Exclusion Chromatography 84
3.18.5 MALDI-TOF MS Method 84
3.18.6 Cure Time 85
3.18.7 Differential Scanning Calorimetry 86
3.19 UF Resin Cure Kinetics 87
3.20 UF Resins with Low Formaldehyde Emission 87
3.21 Modification by Polyamines 88
3.22 Cyclic Urea Prepolymer 89
3.22.1 Preparation of Cyclic Urea Prepolymer 89
3.22.2 Cyclic Urea Prepolymer as a Modifying Resin
for other Adhesives 89
3.23 Improvement of UF and MUF Resins by Addition
of Hyperbranched Dendrimers 89
3.23.1 Urea and Melamine Resins without Formaldehyde 90
References 92
4 Melamine–Formaldehyde Resin 101
4.1 Introduction 101
4.2 Chemistry 101
4.2.1 Formation of Methylolmelamine 102
4.2.2 Condensation of Methylolmelamines 103
4.2.3 Cross-Linking 104
4.3 Melamine–Urea–Formaldehyde (MUF) Resin 105
4.3.1 Liquid MUF Resin Preparation 106
4.3.2 Phenol–MUF (PMUF) Resins 107
4.3.3 Melamine–Formaldehyde Resin Modification
by Acetoguanamine for Post-Formable
High-Pressure Laminate 108
4.3.4 MUF Adhesive Resins of Upgraded Performance 110
4.3.5 Cold-Setting MUF Adhesives 111
References 111
5 Phenol–Formaldehyde Resins 115
5.1 Introduction 115
5.2 Historical 115
5.3 Definitions and Types of Phenolic Resins 116
Contents xi

5.4 Basic Chemistry 116


5.4.1 Resols 116
5.4.2 Novolacs 116
5.4.3 Difference between the Acid and Base Catalysis 116
5.4.4 Reaction between Phenol and Formaldehyde
(Sodium Hydroxide Catalyzed) 117
5.4.4.1 Electron Delocalization in Phenol
and Phenoxide Anion 117
5.4.4.2 Hydroxymethylation of Phenol and Further
Condensation (under Alkaline Conditions) 118
5.4.5 Formation of Chelate Ring 123
5.4.6 Reaction between Phenol and Formaldehyde
(Ammonia and Amine Catalysis) 123
5.4.7 Manufacture of Phenolic Resins 124
5.4.7.1 Principles of Manufacture 124
5.5 Effect of Process Variables 129
5.5.1 Catalyst Types and pH of Resin 129
5.5.2 Effect of Viscosity 131
5.5.3 MW and Its Distribution of PF Resin 132
5.6 Commercial Phenolic Resin for Wood Products 132
5.6.1 Spray Drying of Phenolic Resin 133
5.6.1.1 The Spray Drying Process 133
5.6.2 Phenolic Dry Resin Film 133
5.6.2.1 Types and Grades of Dry Glue Film 134
5.6.2.2 Process of Making the Dry Adhesive Film 134
5.7 Curing of Phenolic Resin 134
5.7.1 PF Cure Acceleration 135
5.7.2 PF Cure Acceleration by Additives 136
5.7.3 Mechanism 137
References 141
6 Resorcinol–Formaldehyde Resins and Hydroxymethyl
Resorcinol (HMR and n-HMR) 147
6.1 Introduction 147
6.2 Reaction between Resorcinol and Formaldehyde 148
6.3 Comparison between Resorcinol and Phenol 149
6.4 Reactive Positions and Types of Linkages Comparison
between Resorcinol and Phenol 151
6.5 Hydroxymethyl Resorcinol 151
6.5.1 Introduction 151
6.5.2 Normal HMR 153
xii Contents

6.5.3 Formulation of HMR 154


6.5.3.1 Mixing Procedure 154
6.5.3.2 Limitations to the Use of HMR 155
6.6 Novolak-Based HMR 155
6.6.1 Preparation of n-HMR 156
6.7 Bonding Mechanism using HMR 156
6.7.1 Mechanism Based on the Material Properties of HMR 156
6.7.2 Mechanism Based on Surface Chemistry 157
6.8 Applications of HMR and n-HMR 157
6.8.1 Bonding to Preservative-Treated Wood 157
6.8.2 Epoxy–Wood Adhesion 158
6.8.3 Bonding of Fiber-Reinforced Polymer–Glulam Panels 158
6.8.4 Priming Agent for Bondability of Wax-Treated Wood 159
6.9 Special Adhesives of Reduced Resorcinol Content 159
6.9.1 Fast-Setting Adhesive for Fingerjointing and Glulam 159
6.9.2 Branched PRF Adhesives 161
6.9.3 Cold-Setting PF Adhesives Containing No Resorcinol 163
References 164
7 Polyurethane Adhesives 169
7.1 Introduction 169
7.2 Historical 170
7.3 Reactions of Isocyanates 171
7.4 Raw Materials 172
7.4.1 Isocyanates 172
7.4.1.1 Aliphatic Isocyanates 172
7.4.1.2 Aromatic Diisocyanates 173
7.5 Catalysts 176
7.6 Blocked Isocyanates 177
7.7 Advantages of pMDI 177
7.8 PU Adhesive–Wood Interaction 178
7.9 PU–UF Hybrid Adhesives 182
7.10 PU–PF Hybrid Adhesives 182
7.11 EMDI-Based Adhesives 183
7.11.1 Comparison between EMDI and pMDI 184
7.12 Emulsion Polymer Isocyanate (EPI) Adhesive 184
7.13 Non-Isocyanate Polyurethanes and Biobased PU Adhesives 185
References 192
8 Wood Surface Inactivation (Thermal) 201
8.1 Introduction 201
8.2 Causes and Sources of Inactivation 202
Contents xiii

8.3 Mechanisms of Inactivation 203


8.4 Factors Affecting Wood Surface Inactivation 203
8.4.1 Effect of Wood Species 203
8.4.2 Inactivation Due to High-Temperature Drying 204
8.4.2.1 Effect of Drying Technique 205
8.5 Physical Mechanisms of Inactivation 206
8.5.1 Effect of Extractives on Wettability and Adhesion 206
8.5.2 Molecular Reorientation at Surfaces 206
8.5.3 Micropore Closure 207
8.6 Chemical Mechanisms of Inactivation 207
8.6.1 Elimination of Surface Hydroxyl Bonding Sites 207
8.6.2 Oxidation and/or Pyrolysis of Surface Bonding Sites 208
8.6.3 Chemical Interference with Resin Cure or Bonding 208
References 208
9 Wood Surface Inactivation Due to Extractives 211
9.1 Introduction 211
9.2 Migration of Extractives to the Wood Surface 211
9.3 Influence of Extractives on Bonding Properties of Wood 212
9.4 Effect of pH of Wood on the Adhesion 217
9.5 Effect of Extractive Migrations during Kiln Seasoning
on Adhesion 218
9.6 Methods to Reduce the Influence of Extractives on Wood
Adhesion 218
9.6.1 Mechanical Method 218
9.6.2 Chemical Method 219
References 220
10 Surface Modification of Wood 223
10.1 Introduction 223
10.2 Surface Modification Methods 224
10.2.1 Plasma and Corona Treatments 224
10.2.2 Corona Treatment 227
10.2.3 Plasma Applications for Wood Surface Plasma
Treatments 228
10.3 Enzymatic Modification for Hydrophobicity 230
10.4 Modification of Wood Surface by Chemical Treatment—
Functionalization of Wood 231
10.5 Sol–Gel Method 233
References 234
xiv Contents

11 The Chemistry of Condensed Tannins 239


11.1 Introduction 239
11.2 Reactions of Condensed Flavonoid Tannins 244
11.2.1 Hydrolysis and Acid and Alkaline Condensation 244
11.2.2 Sulphitation 247
11.2.3 Catechinic Acid Rearrangement 248
11.2.4 Catalytic Tannin Autocondensation 248
11.2.5 Tannin Complexation of Metals 250
11.2.6 Tridimensional Structure 250
11.2.7 Reactivity and Orientation of Electrophilic
Substitutions of Flavonoids 251
11.2.8 Influence of Tannin Colloidal Behavior on Reactions 252
11.2.9 New and Unusual Tannin Reactions 253
11.2.10 Modern Instrumental Methods of Analysis 255
11.3 Conclusions 257
References 259
12 Thermosetting Adhesives Based on Bio-Resources
for Lignocellulosic Composites 267
12.1 Introduction 267
12.2 Tannin Adhesives 268
12.2.1 New Technologies for Industrial Tannin Adhesives 269
12.2.2 Tannin–Hexamethylenetetramine (Hexamine)
Adhesives 270
12.2.3 Hardening by Tannin Autocondensation 272
12.3 Lignin Adhesives 275
12.4 Protein Adhesives 277
12.5 Carbohydrate Adhesives 278
12.6 Unsaturated Oil Adhesives 279
12.7 Wood Welding without Adhesives 283
12.8 Conclusions 286
References 286
13 Environmental Aspects of Adhesives—Emission
of Formaldehyde 293
13.1 Introduction 293
13.2 Scientific Analysis of the Problem 294
13.3 Factors Affecting the Amount of Formaldehyde Emission 296
13.4 Exposure 297
13.5 Safe Level of Formaldehyde Exposure 297
Contents xv

13.6 Evolution of Formaldehyde Emission Standards 298


13.6.1 US HUD Manufactured Housing Standard 298
13.6.2 California Air Resources Board (CARB) Air Toxic
Control Measure for Composite Wood Products 298
13.7 CARB Green Adhesive Formaldehyde Emission Standards 299
13.8 Japanese JIS/JAS Formaldehyde Adhesive Emission Standards 299
13.9 European Formaldehyde Emission Standards 301
13.10 Standardization and Test Methods 301
13.10.1 Reference Methods 301
13.10.2 Certification Methods 302
13.10.3 Quality Control Methods 302
13.11 Different Standards and Test Methods 302
13.11.1 Reference Method 303
13.11.1.1 Chamber Methods 303
13.11.1.2 ASTM E 1333 303
13.11.1.3 ASTM D6007-02(2008) Standard Test
Method for Determining Formaldehyde
Concentration in Air from Wood
Products Using a Small-Scale Chamber 304
13.11.1.4 ISO 12460-1 and Part 2: 2007 305
13.11.1.5 Japanese Small Chamber Method JIS
A1901 305
13.11.2 Derived Methods 305
13.11.2.1 Gas Analysis according to EN 717-2 305
13.11.2.2 Flask Method 306
13.11.2.3 Desiccator Method 308
13.11.2.4 Criteria of Acceptance for Different
Grades are Given in the following Table 310
13.11.2.5 The Perforator Method (EN 120) 310
References 312
14 Rheology and Viscoelasticity of Adhesives 317
14.1 Rheology of Adhesives 317
14.2 Viscosity—Theory 317
14.3 Capillary Viscometry 318
14.4 Rotational Viscometers 320
14.4.1 Spring Type 321
14.4.2 Servo Systems 321
14.5 Cone-and-Plate Viscometer 321
14.6 Parallel Plate Viscometer 322
14.7 Concentric Cylinder Viscometer 322
xvi Contents

14.8 Ford Cup Viscosity 322


14.9 Gardner–Holt Tubes 323
14.10 Newtonian and Non-Newtonian Fluids 324
14.10.1 Types of Non-Newtonian Fluid Behavior 324
14.11 Viscoelasticity of Adhesives 325
14.11.1 Phenomenological Models for Viscoelastic Materials 326
14.11.1.1 Maxwell Element (Elastic Deformation +
Flow) 326
14.11.1.2 Voigt Element (Spring and Dashpot
in Parallel) 327
14.11.1.3 Maxwell–Voigt Mixed Model 327
14.12 Dynamic Mechanical Analysis 328
14.13 TTT and CHT Diagrams 332
14.14 Experimental Results 338
References 342
15 Hot Melt Adhesives 347
15.1 Introduction 347
15.2 Polymers Commonly Used for Hot Melt Adhesives 347
15.2.1 Ethylene Vinyl Acetate Copolymers 347
15.2.2 Styrenic Block Copolymers 350
15.3 Polyureathane Reactive Hot Melt Adhesives 351
15.4 Silane Reactive Hot Melt Adhesives 352
15.5 Polyamide Hot Melt Adhesives 352
15.6 Amorphous Polyolefin (APO/APAO) Hot Melt Adhesives 355
15.7 Tackifiers 356
15.7.1 Aromatic Hydrocarbon Resins 356
15.7.2 Aliphatic Hydrocarbon Resins 356
15.7.3 Mixed Aliphatic and Aromatic Resins 357
15.7.4 Terpene Resins 358
15.7.5 Terpene–Phenol Resins 359
15.7.6 Rosin and Rosin Derivatives 359
15.8 Antioxidants 361
15.8.1 Oxidation-Sensitive Components in Hot Melts 362
15.8.2 Antioxidants Used in Hot Melts 362
15.9 Plasticizers 363
15.10 Mineral Oil and Wax 364
References 364
Contents xvii

Part B: Polymer Matrix Materials For Biofiber Composites


16 Modification of Natural Fibers and Polymeric Matrices 369
16.1 Introduction 369
16.2 Strategies to Treat the Biofibers for Compatibility 370
16.2.1 Physical Methods 370
16.2.2 Steam Explosion Treatment 371
16.3 Chemical Methods 371
16.3.1 Mercerization 372
16.3.2 Acetylation of Natural Fibers 373
16.3.3 Silane Coupling Agents 373
16.3.4 Benzoylation Treatment 374
16.3.5 Acrylation of Natural Fibers 374
16.3.6 Treatment with Isocyanates 375
16.3.7 Peroxide Treatment 376
16.3.8 Permanganate Treatment 377
16.3.9 MAH Treatment 377
16.3.10 Treatment with Chlorotriazines 377
16.3.11 Additives 378
16.4 Functionalization of Matrices for Compatibility 378
16.5 MAH Grafted Polyolefins as Matrix Additives 382
16.6 Reactive Extrusion System 383
References 383
17 Polymer Matrix: Unsaturated Polyester 389
17.1 Introduction 389
17.2 Raw Materials 389
17.2.1 Diols 389
17.2.2 Cyclopentadiene-Based Resin 390
17.2.3 Isophthalic-Acid-Based Resin 391
17.2.4 Bisphenol A Fumarate Resins 391
17.2.5 Vinyl Ester 391
17.3 Polyesterification Reaction 392
17.4 Cross-Linking Reaction 393
17.4.1 Curing at Elevated Temperatures 393
17.4.2 Curing at Room Temperatures 394
17.5 Sheet Molding Compounds Based on UP Resins 395
17.6 UV Curable Compositions Based on UP/Vinyl Ester Resins 396
17.7 Biocomposites Based on UP Matrix 397
References 400
xviii Contents

18 Polymer Matrix: Epoxy Resins 403


18.1 Introduction 403
18.2 Resin Preparation 403
18.3 Characteristics of Epoxy Resins 404
18.3.1 Epoxy Equivalent 404
18.3.2 Enhancement of Properties 405
18.3.3 Types of Epoxy Resins 405
18.3.4 Bisphenol A Glycidyl Ethers 405
18.4 Preparation of DGEBA Epoxy Resin 406
18.4.1 Curing Agents 406
18.4.1.1 Tertiary Amines 408
18.4.1.2 Polyfunctional Amines 408
18.4.1.3 Calculations of the Proportion
of Amines for Curing Epoxy Resins 411
18.4.1.4 Special Amines 411
18.4.1.5 Acid Anhydrides 412
18.4.1.6 Anhydride Curing Mechanism 412
18.5 Other Types of Epoxy Resins 414
18.5.1 Epoxidized Novolac 414
18.5.2 Tetrabromo Bisphenol A Epoxy Resins 415
18.5.3 Epoxidized Vegetable Oils 416
18.5.4 Epoxidized Natural Rubber 417
18.6 Green or Sustainable Epoxy Matrix 417
18.7 Epoxy-Matrix-Based Biofiber Composites 419
References 420
19 Polymer Matrix: Polyethylene 425
19.1 Introduction 425
19.2 High-Pressure Process 426
19.3 Low-Pressure Processes—Catalysts for Polymerization 426
19.3.1 Ziegler–Natta Catalysts 427
19.4 Production of PE 428
19.4.1 Solution Process 428
19.4.2 Slurry Process 428
19.4.3 Gas Phase Fluidized Bed Reactor 429
19.5 Compatibilizers 429
19.6 Relevant Property of PE 430
19.6.1 Melt Flow Index 430
19.7 Treatment and Functionalizing of Biofibers and Matrix
Materials 430
Contents xix

19.8 Biocomposites Based on PE 430


19.8.1 Kenaf-Based biocomposites 430
19.8.2 Sisal-Fiber-Based Biocomposites 432
19.8.3 Flax-Fiber-Based Biocomposites 434
19.8.4 Hemp-Fiber-Based Biocomposites 435
19.8.5 Miscellaneous 436
References 437
20 Polymer Matrix: Polypropylene 441
20.1 Introduction 441
20.2 PP Manufacture 441
20.2.1 Catalysts 442
20.2.2 α- and β-Forms of PP 442
20.2.3 Polymerization Methods 443
20.2.3.1 Solvent Polymerization Process 443
20.2.3.2 Bulk Polymerization Process 444
20.2.3.3 Gas Phase Polymerization Process 444
20.3 Biofiber Composites Based on PP 445
20.3.1 Kenaf-Based Composites 445
20.3.2 Oil-Palm-Fiber-Based Composites 448
20.3.3 Flax-Fiber-Based Composites 454
20.3.4 Sisal-Based PP Composites 455
20.3.5 Hemp-Based PP Composites 456
References 460
21 Biodegradable Polymers as Matrix for Biocomposites 467
21.1 Introduction 467
21.2 Polyhydroxyalkanoates 467
21.2.1 Poly(3-hydroxybutyrate) PHB 469
21.2.2 Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) 470
21.3 Polylactic Acid 470
21.3.1 Synthesis of PLA 471
21.3.2 Direct Polymerization 472
21.3.2.1 Solution Polycondensation 472
21.3.2.2 Melt Polycondensation 472
21.3.2.3 Ring-Opening Polymerization 473
21.4 Polybutylene Adipate Terephthalate 474
21.5 All Green Composites 474
References 477
Index 483
Preface

The dramatic increase in the production of reconstituted wood products


over the past 100 years has been made possible through the systematic
development of new adhesives devised to meet the growing challenges of
the times. Originally, large-diameter trees were available, but as their avail-
ability decreased, followed by a corresponding increase in price, the wood
industry had to meet the challenge by attempting to use small-diameter
trees and comminuted wood particles, fibers and pulp mill waste. This
resulted in the development of reconstituted wood produced from the
comminuted particles bound together with adhesives. This development
had a significant impact on meeting challenges. The development of new
types of adhesives—both synthetic petroleum-based adhesives as well as
adhesives of natural origin—occurred during this period.
This trend of moving away from solid wood towards the utilization of
elements of regularly reduced dimensions was recognized by Marra [1],
who illustrated his concept with the “nonperiodic table of wood elements”
shown in Figure 1 [2].
This concept established the future trends in the wood-based industry,
namely, (1) use of smaller trees, (2) use of waste from other wood pro-
cessing, (3) removal of defects, (4) use of rare and hitherto unused wood
species, natural lignocellulosic fibers, (5) creation of more uniform com-
ponents, (6) development of composites stronger than the original solid
wood, (7) ability to make composites of different shapes and (8) glulams,
OSB, LVL, etc., (9) development of natural-fiber polymer-matrix compos-
ites, (9) development of more sophisticated engineered wood products and
structural elements, such as wooden I-Joist box beams, aided by the avail-
ability of new or improved wood adhesives, (10) development of sandwich
composites of wood and non-wood materials such as metal- and plastic-
faced wood panels, paper and metal honeycomb sandwiches, etc.
It is interesting to note that an answer is slowly emerging to the question
mark in Figure 1. Exciting new opportunities are emerging in the field of
biorefining to produce chemical feedstocks, syngas, and nanocrystalline

xxi
xxii Preface

Figure 1 Basic wood elements from largest to smallest (i.e., breakdown of solid wood into
finer elementary components [1].

cellulose. In the near future, nanocrystalline cellulose, produced as a high-


value by-product from the biorefining process, could likely compete with
carbon fiber for use in innovative high-strength biocomposites.
The above developments markedly increased the percentage of adhesives
used for the production of glued wood products. It should be mentioned
in this context that a high percentage, maybe 80% or more, of all wood
products produced today are glued, and that about 70% (by volume) of all
the adhesives produced in the world today are used for application to wood
[3]. These developments have led to an increase in the functional efficiency
of wood products as well as an efficient utilization of wood resources, thus
constituting an essential tool to directly or indirectly affect the sustainabil-
ity of forestry and wood-based industries.
Although a number of books have appeared on the subject of adhesives
in general and wood in particular, this book is unique because of the vast
academic teaching and research experience and hands-on industrial expe-
rience of the authors. Their skills have been brought to bear on identifying
very important and unique combinations of current topics constituting the
essential contents of the book. Furthermore, this book, besides the adhe-
sives for wood detailed in Part A, also deals with the polymeric matrix
Preface xxiii

materials for natural-fiber-based composites in Part B. The decision to


include polymer matrix materials was made in consideration of growing
global interest in wood-polymer composites based on natural fibers during
the past decade.
The first chapter of the book deals with the distinctiveness of wood as
an adherend in the midst of other substrates such as metals, polymers,
inorganic adherends like glass, etc. In contrast to other substrates, wood
presents adhesives with hierarchical structural elements of different sizes
which, along with its unique chemical and physical characteristics, greatly
influence the wood-adhesive interaction.
Knowledge of the fundamentals of adhesion is extremely important
for researchers as well as technologists in the industry, both for adhesive
formulations and troubleshooting during production. The importance of
establishing an intimate contact between the adhesive and wood has been
emphasized for an effective performance and durability of the bonded
wood products in actual service. Therefore, mechanical interlocking, cou-
lombic (ionic) interaction, hydrogen bonding, and apolar interactions
are discussed in Chapter 2. In addition, electronic or electrostatic theory,
adsorption (thermodynamic) or wetting theory, diffusion theory, chemi-
cal (covalent) bonding theory, theory of weak boundary layers and inter-
phases and interfacial forces based on specific donor-acceptor (acid-base)
interactions between adhesive and substrate molecules are also discussed.
In Chapters 3 to 7, the chemistry and technology of urea-formaldehyde,
melamine formaldehyde, phenol-formaldehyde, resorcinol-formaldehyde,
and polyurethanes are discussed in detail. Special mention is given to
non-isocyanate polyurethanes (NIPUs) and biobased polyurethane adhe-
sives in Chapter 7.
Surface inactivation peculiar to wood is dealt with in Chapters 8 and 9.
In order to resolve this problem, surface modification by suitable treatment
is dealt with in Chapter 10. Treatment of biofibers is considered in Chapter
16. All of these chapters are very important for technologists working in
the wood industry.
In order to reduce the use of petroleum-derived phenol for the manufac-
ture of phenolic resins, a lot of research has been carried out on the partial
or whole substitution of phenol by natural polyphenols, namely tannins.
Chapter 11 is an exhaustive account of the chemistry of condensed tan-
nins. A good understanding of the chemistry of condensed tannins is very
necessary for developing new adhesives based on natural polyphenols.
Chapter 12 discusses in great detail the technology of tannin adhesives,
particularly new technologies for industrial tannin adhesives, lignin adhe-
sives, protein adhesives, carbohydrate adhesives, unsaturated oil adhesives,
xxiv Preface

and cardanol-based adhesives. The chapter also deals with wood welding
without adhesives.
The environmental aspects of adhesives, namely formaldehyde emis-
sion, are discussed in Chapter 13. Formaldehyde is of particular concern
due to its classification as a “known human carcinogen” in the August 8,
2014 publication of the 12th Report on Carcinogens (RoC). Therefore,
formaldehyde emission standards are dealt with in detail in this chap-
ter. Next, the rheology and viscoelasticity of adhesives is the subject of
Chapter 14 and Chapter 15 discusses hot melt adhesives.
Chapters 17 to 21 are included in Part B (Polymer Matrix Materials for
Biofiber Composites) of the book. Both thermoplastic and thermosetting
matrix materials are discussed in detail.
The author (RNK) thanks Dr. V.V. Srinivasan, former Director of the
Institute of Wood Science, Bangalore, for suggesting the idea of this book.
The author also expresses his sincere gratitude to Prof. Pizzi for volun-
teering to co-author the book at a time when I had abandoned the idea
of writing it. His encouragement, chapter contributions, help in editing
the chapters and adding very important factual details, and particularly
his great patience in arranging the references, is gratefully acknowledged.
The author thanks Mr. P.K. Mayan, Managing Director, Western India
Plywoods Ltd, Baliapatam, Kerala, for his encouragement. I record my
thanks to my son, Dr. Suresh Nandakumar, for his helpful suggestions. The
authors thank Mr. Martin Scrivener for his unequivocal support.

References
1. Marra, G. Overview of wood as a material, J. Educ. Modules for Mater. Sci.
Eng. 1(4), 699–710, 1979.
2. Berglund L. and Rowell R.M. Wood composites, in Handbook of Wood
Chemistry and Wood Composites, Routledge/Taylor & Francis, 2005.
3. Pizzi, A. Special section: Wood adhesives. Foreword. Int. J. Adhes. Adhes. 18,
67, 1998.
Part A
SUBSTRATES, ADHESIVES,
AND ADHESION

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (1–30) © 2019
Scrivener Publishing LLC
1
Wood as a Unique Adherend

1.1 Introduction
In order to make durable wood adhesive bonds in composite wood products,
a clear understanding of the nature and uniqueness of wood as a substrate
and of the distinctiveness of the wood–adhesive interaction is essential.
In this context, it is necessary to mention that substantial differences exist
between bonding in the case of wood on the one hand and most other mate-
rials on the other. The most obvious characteristics of wood that distinguish
wood from other substrates are (a) its porosity, (b) presence of intercon-
nected cells into which adhesive can flow, and (c) the cell walls that have the
ability to allow low-molecular-weight chemicals and resins to pass through
and in some cases even to react with them. All the above features are due to
the special identity that wood possesses in contrast to other substrates.
It is known that wood exhibits multiscale hierarchical structures. As
reported by Gao [1], structural hierarchy is a rule of nature and can be
observed in many other natural and man-made materials. In recent years,
these materials have been called multiscale materials. Hierarchical solids
contain structural elements that themselves have further finer structures
[2]. In this respect, wood as an adherend is significantly different from
other adherends such as metals and plastics.

1.2 Wood, An Adherend with Hierarchical Structure


Wood is basically a fiber-reinforced polymer matrix composite with cellu-
lose as the fiber reinforcement and lignin along with polyoses functioning
as mixed polymer matrix. Wood has been found to have a hierarchical cel-
lular structure. The functional efficiency of wood, its exceptional mechan-
ical properties, and other unique characteristics are attributed to the
distinctive structure at all its levels of hierarchy. The complex hierarchical
structure of wood manifests in a wide range operating from the macro

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (3–30) © 2019
Scrivener Publishing LLC

3
4 Adhesives for Wood and Lignocellulosic Materials

scale in the case of tree to board to growth ring (macro scale), through
multilayered cell walls, composite cellulose/hemicellulose microfibrillar
structure (micro scale), and down to the structure of the three main poly-
mer components, cellulose, hemicellulose, and lignin (nanoscale).
Hierarchical solids contain basically structural elements that, in turn,
have further finer structures; i.e., intricate structural features occur at dif-
ferent size scales.
For instance, the hierarchical nature of wood showing the size scale of
each structural element within wood has been illustrated by Moon et al. [3]
with typical dimensions given in parentheses:
Tree height (in meters), tree cross section (in cm), growth rings (in mm),
cellular structure (in 500 μm), cell wall structure (in 25 μm), fibril-matrix
structure (in 300 nm), fibril structure (10 nm), and cellulose (1 nm).
This structural hierarchy of wood can play a significant role in influenc-
ing the phenomenon of wood adhesion [3].
In order to maximize the strength and durability of adhesive bonds, one
should understand the complex hierarchical structural elements of wood and
their interactions with the molecules of the adhesives at various size scales.
It should be appreciated that hierarchical structural elements of wood
when exposed to the molecules of adhesives during bonding interact at
different size scales in an extremely unique manner not encountered in
the case of other adherends. Although the concept of hierarchical model
of wood has not been explicitly mentioned, the importance of practical
length scale of wood composite elements in wood adhesive bonding has
been recognized and reported [4].

1.3 Details of Structural Hierarchy in Wood


The structural hierarchy confers on wood very unique properties.
Accordingly, wood is porous, permeable, hygroscopic, and orthotropic. It is
a biological composite material of extreme chemical diversity and physical
intricacy. Further, wood properties vary between species and even within
the same species. They can vary even within a tree. Variability within a
single species could be per se significant enough to throw challenge to an
adhesive for its consistent and satisfactory performance [5].

1.3.1 Physical Structure


There are two major types of wood: softwood (from needle-bearing
trees) and hardwood (from broad-leaved trees). The terms softwood and
Wood as a Unique Adherend 5

hardwood are misnomers and have little to do with hardness or density.


Many softwoods are actually much harder than many hardwoods [5].
In both softwoods and hardwoods, the most noticeable features are the
light outer ring and the dark inner core. Thus, the wood in the trunk of the
tree is typically divided into two zones, each of which serves an important
function distinct from the other. The actively conducting portion of the
stem in which parenchyma cells (explained later) are still alive and meta-
bolically active is referred to as sapwood [6].
A looser, more broadly applied definition is that sapwood is the band
of lighter-colored wood adjacent to the bark. Heartwood is the darker-
colored wood found to the interior of the sapwood. As the tree grows
larger in diameter, cells closer to the center of the tree, which are no longer
required for these activities, die and are converted to heartwood.
In the living tree, sapwood is responsible not only for conducting sap
but also for storage and synthesis of biochemicals. An important storage
function is the long-term storage of photosynthesis products. The primary
storage products of photosynthates are starch and lipids. Living cells of the
sapwood are also the agents of heartwood formation. Biochemicals must
be actively synthesized and translocated by living cells [6]. For this reason,
living cells at the border between heartwood and sapwood are responsible
for the formation and deposition of heartwood chemicals, one important
step leading to heartwood formation [7].
These chemicals are collectively named as extractives and are known for
protecting the wood. Extractives also play a significant role in the adhesive
bonding of wood. Extractives are formed by parenchyma cells at the heart-
wood–sapwood boundary and are then exuded through pits into adjacent cells
[7]. In this way, dead cells can become occluded or infiltrated with extractives
despite the fact that these cells lack the ability to synthesize or accumulate these
compounds on their own. When these chemicals oxidize, the heartwood dark-
ens, and the border between sapwood and heartwood becomes more distinct.
Heartwood formation differs markedly between species, even within trees of
the same species. These changes, especially the chemical changes, account for
much of the difficulty and unpredictability in the bonding of heartwood [5].

1.3.1.1 Growth Rings and Ring-Porous and Diffuse-Porous Wood


Wood cells are produced by the vascular cambium, the only living part of
the tree, at the boundary between bark and sapwood, one layer of cell divi-
sions at a time. These collections of cells produced together over a discrete
time interval are known as growth increments or growth rings. Cells formed
at the beginning of the growth increment are called earlywood cells, and
6 Adhesives for Wood and Lignocellulosic Materials

Earlywood

Latewood

Figure 1.1 Earlywood and latewood [5].

cells formed in the latter portion of the growth increment are called late-
wood cells (Figure 1.1). Springwood and summerwood were terms for-
merly used to refer to earlywood and latewood, respectively, but their use
is no longer recommended [8].
The growth rings are usually prominent because of cyclical variation in
color or porosity. These variations are in turn due to the formation of dif-
ferent types of cells and wood structures during different parts of the grow-
ing season. The lighter-colored (less dense) and more porous cell tissue of
earlywood is formed early in the growing season. The porous earlywood
cells are largely responsible for the movement of liquid and nutrients
within the tree. The darker (more dense) and less porous cell tissue of the
latewood, formed later in the growing season, is largely responsible for
mechanically supporting the tree [5] (Figure 1.1).
Large differences between the earlywood and latewood porosity and
density in some species like oak and southern pine often cause difficulty
in bonding [5].
Hardwoods may be divided into  ring-porous and diffuse-porous
woods. Diffuse-porous woods have vessels of roughly the same radial
diameter throughout the growing season. In the diffuse porous wood, the
pores are distributed evenly throughout the wood.

1.3.1.2 Wood Cells


A living plant cell consists of two primary domains: the protoplast and the
cell wall. The protoplast is the sum of the living contents that are bounded
by the cell membrane. The cell wall is a non-living, largely carbohydrate
matrix extruded by the protoplast to the exterior of the cell membrane. The
plant cell wall protects the protoplast from osmotic lysis and often provides
mechanical support to the plant at large [9, 10].
In wood, the ultimate function of the cell is borne solely by the cell wall.
This means that many mature wood cells not only do not require their
protoplasts, but indeed must completely remove their protoplasts prior
Wood as a Unique Adherend 7

to achieving functional maturity. For this reason, a common convention


in wood literature is to refer to a cell wall without a protoplast as a cell.
Although this is technically incorrect from a cell biological standpoint, this
convention is common in the literature [6].
In the case of a mature cell in wood in which there is no protoplast,
the open portion of the cell where the protoplast would have existed is
known as the lumen (plural: lumina). Thus, in the wood cells, there are two
domains: the cell wall and the lumen.
Wood cells are microscopic, long, thin, hollow tubes, like soda straws with
their ends pinched shut. Most longitudinal cells that are parallel to the longi-
tudinal axis or grain direction of the tree trunk are meant either for support
or for the movement of fluids in the living tree. Some special cells are orga-
nized into tissue called rays that lie perpendicular to the longitudinal axis of
the tree trunk and along its radii. Ray cells are responsible for the production
and storage of amorphous materials of complex chemical nature. The rays
are also the pathway for lateral movement of fluids in the tree [5].
There are two basic types of cells—prosenchyma and parenchyma.
Softwoods and hardwoods have different types of prosenchyma and paren-
chyma cells. Prosenchyma cells are generally the strong woody cells respon-
sible for mechanical support and the movement of fluids in the living tree.
Parenchyma cells are responsible for the production of chemicals and for
the movement and storage of food. The real differences between softwoods
and hardwoods are in the size, shape, and diversity of these two types of
cells [5].
The structure of softwoods is characterized by relatively few types of
prosenchyma and parenchyma cells compared to hardwoods as a result of
their lower position on the evolutionary scale. One type of prosenchyma
cell, the longitudinal tracheid, constitutes approximately 90–94% of the
volume of softwood wood. Tracheids perform both the support and fluid
movement for the tree. Earlywood tracheids are generally of large diam-
eter and thin walled. Earlywood cells are specifically adapted to moving
fluids through large openings (bordered pits) that connect adjoining cells.
Latewood tracheids, which are generally smaller in diameter, are thicker
walled, have smaller pits, and are specifically adapted for strength. The
remaining 10% of softwood consists of longitudinal parenchyma cells, ray
tracheids, and ray parenchyma cells. Generally, parenchyma cells play a
secondary strength role, but they are important for adhesive bonding as
paths for adhesive penetration. Moreover, the chemicals contained by the
cells affect adhesion and adhesive cure.
In comparison to softwoods, the structure of hardwoods is characterized
by a greater diversity of cell types and functions. One notable difference is
8 Adhesives for Wood and Lignocellulosic Materials

that one type of specialized prosenchyma cells is responsible for mechanical


support, and another type of specialized prosenchyma cells is responsible
for fluids movement. Support is provided by two types of small-diameter
thick-walled prosenchyma cells called libriform fibers and fiber tracheids.
Fluid movement is provided by medium- to large-diameter, thin-walled,
and open-ended cells called vessel elements. Normally, a number of vessel
elements link end to end along the grain to form long tube-like structures
known as vessels. The cavities are large enough to see with the naked eye.
Such large cavities obviously affect wood strength and adhesive flow when
pressure is applied during bonding. The longitudinally oriented fibers and
vessels together constitute the major volume of cells (roughly 70–90%)
in hardwoods. A number of other specialized longitudinal prosenchyma
and parenchyma cells and ray prosenchyma and parenchyma cells con-
stitute the remaining volume. As in the softwoods, some of these minor
hardwood cell types have important chemical roles and secondary, though
often minor, mechanical roles.

1.3.1.3 Organization of Cell Walls in Wood


The fiber ultrastructure is the hierarchic level ranging from the molecular
level up to the fiber cell wall layers. The structure of the actual wood cell
wall is very complex.
Under strong magnification, the visible height of various layers can be
recognized in the wood cell wall. A clear demarcation between the individ-
ual layers can be seen with an electron microscope. With the aid of an
electron microscope, the current knowledge of the structural composition
of the wood cell walls was obtained between 1950 and 1970 [11].
For a clear understanding of the cell wall structure of wood, one should
recognize the basic structural units, namely, fibrils, elementary fibrils, and
microfibrils. They are described below:
The cellulose molecules are linear and can exhibit a high degree of lateral
order, and they are therefore capable of forming strong intra- and intermo-
lecular hydrogen bonds and aggregated bundles of molecules. Thus, the
cellulose molecular chains can be organized into strands called fibrils. In
the literature, these bundles of cellulose molecules have been given many
different names, such as elementary fibrils, microfibrils, protofibrils, etc.
[12–14]. The term cellulose microfibrils will be used in this book. These
cellulose microfibrils have crystalline and amorphous regions.
Elementary fibrils are cellulosic strands of smallest possible diameter
(35 Å). An elementary fibril of this cross-sectional dimension could con-
tain about 40 cellulose chains. Aggregates of elementary fibers are classed
Wood as a Unique Adherend 9

as microfibrils and occur in nature in a broad spectrum of sizes depending


on the source of lignin. They are probably 100 to 300 Å wide and evidently
of indefinite length. Microfibrils aggregate into larger units called macro-
fibrils and they are joined into lamellae that are organized into cell wall
layers [11].
Each elementary fibril contains about 40 cellulose chains, 20–60 ele-
mentary fibrils are fasciculated to form a microfibril, and about 20 micro-
fibrils form a macrofibril.
The cell walls of wood are made up of oriented cellulose microfibrils
embedded in a matrix of lignin and hemicelluloses. They are highly struc-
tured layers that are arranged concentrically and are formed at different
periods during cell differentiation. Between the cells is the region called
middle lamella, which ensures adhesion between adjacent cells and is
made up of pectic substances and lignin [15].
The primary wall: Attached to the middle lamella is the primary cell
wall (0.1 mm), which has a randomly oriented, loose weaving of micro-
fibrils. This wall is very thin and does not show the lamellation observed
in the secondary wall. The primary wall is the first layer deposited during
the development of the cell and provides the framework for the subse-
quent formation of the secondary wall.
In the secondary wall, the microfibrils are closely packed. The amount
of lignin is low (10–20%), and cellulose content ranged from about 50%
to over 60%. The secondary wall is formed of three distinct layers, the
S1, S2, and S3 layers; each layer is much thicker than the primary wall
(Figure 1.2). This is the principal structural element of the wood cell with
the microfibrils aligned helically around the lumen. The degree of orienta-
tion of the microfibrils and the proportions of the various layers in the cell

S3

S2

S1

Figure 1.2 Schematic model of the cell wall layers [16].


10 Adhesives for Wood and Lignocellulosic Materials

wall determine the properties of the cell. This is the most important layer
in terms of mechanical properties.
The S1 layer is thin and consists of a  few lamellae. It has a crossed
microfibrillar texture, with its lamellae exhibiting an alternating left-hand
and right-hand helical arrangement. In each lamella, the helical angle is
about 50–90°, as measured from the longitudinal axis of the cell.
The S2 layer is thick, especially in latewood tracheids and thick-walled
fibers. It is composed of 30–150 lamellae. Adjoining lamellae are observed
to exhibit a similar (not crossed) microfibrillar orientation. The microfi-
brils show a high degree of parallelism in all lamellae, and they run approx-
imately parallel to cell axis, usually not exceeding an angle of about 30°.
Typically, the S2 layer, whose microfibrils are oriented nearly parallel to the
long axis of the cell, is responsible for resisting the principal stresses in the
living tree and for the high longitudinal strength and stiffness of lumber
cut from the tree. Thick S2 layers of latewood cells are considered as the
reason for their resistance to stresses parallel to the long axis of the cell.
The S3 layer is usually thinner than the S1 and it is lamellate (up to six
lamellae). The angle of microfibrils likewise varies from about 50° to 90°.
The S3 layer may sometimes be missing.
The warty layer may line the lumina and pit cavities of tracheids, fibers,
and vessels of many softwoods and hardwoods. This layer is attached to
the S3 layer, and may occur on top of spiral thickenings. The warty layer
forms during the final stages of cell wall development, and it is regarded
as a structure arising from the dying protoplasm (deposited by the living
protoplasm prior to its degeneration). The major chemical component of
warts is lignin; hemicelluloses are also present.
Unlike the lumen, which is a void space, the cell wall has a highly reg-
ular structure, differing in nature from one cell type to another and also
vary between species, and even within the same softwood and hardwood
species.
Various electron microscopic observations gave rise to the model of the
construction of wood cell walls as shown in Figure 1.2 [5].

1.4 Chemical Composition


Wood is made up of cell wall constituents and extraneous materials
(Figure 1.3). The cell wall constituents, namely, polysaccharides and
lignin, are collectively called wood substance, which typically accounts
for 95–98% of the weight of the wood, the remainder being extraneous
organic and inorganic materials. The ratio of polysaccharides to lignin in
Wood as a Unique Adherend 11

wood

Extraneous materials
Cell wall
components

Extractable Insoluables
materials Lignin
(Extraactives)
Polysaccharides

Cellulose Hemi-celluloses

Figure 1.3 Chemical composition of wood.

the wood substance is roughly 3:1. The most abundant polysaccharide,


cellulose, in the form of microfibrils, provides the structural framework
for all plant tissues. Pettersen [17] provides an extensive compilation of
the chemical composition of woods from the entire world.

1.4.1 Cellulose
Cellulose is the most abundant biopolymer on earth. It is synthesized in
plants (trees, grass, etc.), algae (Valonia, Cladophora, etc.), and even in
some animals (tunicates), and it can also be synthesized by some bacteria
(Acetobacter xylinum). Around 40% of the dry weight of wood consists of
cellulose. Cellulose is a linear polymer built up of D-anhydroglucose units
linked together by β-(1-4)-glycosidic bonds. The degree of polymeriza-
tion (DP) is normally 9000–10,000 glucose units, but DP values as high as
15,000 glucose units have been reported [18]. Most of the cellulose found
in wood fibers has approximately the same molecular size, i.e., a very low
polydispersity [18].
The cellulose molecule is linear and it is therefore capable of forming
strong intra- and intermolecular hydrogen bonds and aggregated bundles
12 Adhesives for Wood and Lignocellulosic Materials

of molecules. In the literature, these bundles of cellulose molecules, as has


been already mentioned, have been given many different names such as
elementary fibrils, microfibrils, protofibrils, etc. [12–14]. The term cellu-
lose microfibrils will be used in this book. These cellulose microfibrils have
crystalline and amorphous regions.

1.4.2 Hemicelluloses
Hemicelluloses are a group of heterogeneous polymers that play a support-
ing role in the fiber wall. Twenty to thirty percent of the dry weight of
wood consists of hemicelluloses. The hemicellulose polymers are built up
of several different monomers, such as mannose, arabinose, xylose, galac-
tose, and glucose. Some acidic sugars like galacturonic acid and glucuronic
acid are also constituents of hemicelluloses. One, two, or several types of
monomers usually build up the backbone of hemicellulose polymers. Most
of the hemicelluloses also have short branches containing types of sugars
other than those of the backbone. The degree of polymerization for the
hemicelluloses is between 100 and 200 [11].
Softwoods contain about 20–25% glucomannans. Acetyl groups
and galactose residues are attached to the polymer chain. The hydroxyl
groups at the C(2) and C(3) positions in the chain are partly substituted
by O-acetyl groups. Galactose units are also attached to the chain as α-(1-
6)-linkages. Hence, softwood mannans can be designated as O-acetyl-
galactoglucomannans [11]. Thus, the galactoglucomannans of softwood
have a backbone of (1-4) linked by β-D-glucose and β-D-mannose units in
the main chain with α-D-galactose linked to the chain through (1-6)-bonds.
An important structural feature is that the hydroxyl groups at C(2) and C(3)
positions in the chain units are partially substituted by O-acetyl groups.
It has a backbone of (1-4)-β-D-xylose, where most of the xylose residues
have an acetyl group at C(2) or C(3). About every 10th xylose unit also has
a 4-O-methyl-α-D-glucuronic acid residue linked by a (1-2)-bond (Figure
1.5). The backbone substitution and degree of branching can vary consid-
erably between hemicelluloses of the same category [19].

1.4.3 Lignin
Lignin is a heterogeneous three-dimensional polymer that constitutes
approximately 30% of the dry weight of wood. Lignin limits the penetra-
tion of water into the wood cells and makes wood very compact.
Lignin is the second most abundant and important organic substance
in the plant world. Lignin increases the mechanical strength and stiffness
Wood as a Unique Adherend 13

properties to such an extent that huge plants such as trees with heights of
even more than 100 m can remain upright.
Lignin is a complex polymer of phenylpropane units (monolignols)
joined through many different linkages. The three monolignols are
p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol (Figure 1.4).
The monolignols in various proportions are the building blocks for the
3-D structure of native lignin in higher plants [19]. Monolignols copoly-
merize mainly by radical polymerization.
Thus, the monolignols are the building blocks of the lignin macromolecule.
Lignin is therefore defined as an amorphous polyphenolic material arising
from an enzyme-mediated dehydrogenative polymerization (DHP) of three
phenylpropanoid monomers, coniferyl, synapyl, and p-coumaryl alcohols.
The following nomenclature of radicals (Figure 1.5) and the building
units from which lignin is derived should be kept in mind to understand
scientific publications on lignin:
Lignin’s functions in the tree are as follows (Wang Wood Chemistry
Class):

(a) Support for mechanical strength


(b) Antioxidant for protection
(c) Sealant and reinforcing agents, bonding cellulose and
hemicellulose together
(d) Cross-linker to cross-link he polymeric carbohydrates

The so-called lignin–carbohydrates complex defines the type of cova-


lent bonds existing between lignin and hemicelluloses. These are mainly
benzyl ester and benzyl ether linkages between the side chains of xylans
and phenyl glycosidic linkages with the main chain of glucomannans [11].

CH2OH CH2OH CH2OH

CH CH CH

CH CH CH

OCH3 CH3O OCH3


OH OH OH

p-Coumaryl alcohol Coniferyl alcohol Sinapyl alcohol

Figure 1.4 The three monolignols.


14 Adhesives for Wood and Lignocellulosic Materials

1 (γ) OH OH OH

3 (α) 2 (β)
1’
6’ 2’
5’ 3’
4’ OCH3 CH3O OCH3
OH OH OH
p-coumaryl coniferyl sinapyl

1’
6’ 2’
5’ 3’
4’ OCH3 CH3O OCH3
OH OH OH
p-hydroxyphenyl (H) guaiacyl (G) syringyl (S)

Figure 1.5 Radicals and units—nomenclature.

1.4.3.1 Lignin Isolation


Isolation of lignin in an unchanged form (native lignin) is rendered dif-
ficult due to its complex structure and its location within the cell wall.
Hence, determination of the exact chemical structure of lignin is therefore
difficult [11]. All methods of isolation have the disadvantage of either fun-
damentally changing the native structure of lignin or releasing only a part
of it in a relatively unchanged condition.
There are two methods by which lignin can be isolated from extractive
free wood [19]:

(1) As an insoluble residue after hydrolytic removal of the


polysaccharides.
(2) Alternatively, lignin can be hydrolyzed from wood or
converted into soluble derivative.

According to method (1) the polysaccharides can be removed in the


following procedures:

(1) Klason lignin: Klason lignin is obtained after removing the


polysaccharides from extractives-free wood by hydrolysis
with 72% sulfuric acid.
Wood as a Unique Adherend 15

(2) Cellulolytic enzymes may be used to dissolve polysaccha-


rides from finely divided wood meal leaving behind lignin
as residue. This lignin is called cellulolytic enzyme lignin
(CEL). This method is tedious but the CEL retains its orig-
inal structure essentially unchanged.

Bjőrkman lignin, also called milled wood lignin (MWL), has been widely
used for structural studies. Wood meal is ground in a ball mill either with-
out any solvent or with a solvent such as toluene, which is a non-swelling
solvent. The lignin can then be extracted by using a mixture of dioxane,
water, and HCl (dioxane lignin). Lower temperature extraction minimizes
structural changes of lignin.
A number of researchers have tried the dioxane method in lower tem-
perature (and consequently lower yields) to minimize structural changes
in extracted lignins [4].
For the extraction of lignins, a modified dioxane method and ionic
liquid and comparative molecular weight (MW) and structural studies
by chromatography and ¹³C NMR spectroscopy techniques were used
[20].
In recent years, ionic liquids have been used to dissolve carbohydrates,
and lignin residue is hoped to be relatively unchanged [21].
The second method involves the formation of soluble lignin derivatives,
namely, lignosulfonate.

1.4.3.2 Functional Groups in Lignin


Lignin contains characteristic methoxyl groups, phenolic hydroxyl groups,
and some terminal aldehyde groups in the side chains. Only relatively few
of the phenolic hydroxyls are free; most of them are occupied by linkages to
neighboring phenylpropane units. The syringyl units in hardwood lignin
are extensively etherified. Alcoholic hydroxyl groups and carbonyl groups
are introduced into the final lignin polymer during the dehydrogenative
polymerization process.

1.4.3.3 Evidences for the Phenylpropane Units as Building Blocks


of Lignin
The following methods based on classical organic chemistry, namely,
degradation and synthesis, led to the conclusion, already by 1940, that
lignin is built up by phenylpropane units [11].
16 Adhesives for Wood and Lignocellulosic Materials

(a) Permanganate oxidation (methylated spruce lignin) affords


veratric acid (3,4-dimethoxybenzoic acid) and minor
amounts of isohemipinic (4,5-dimethoxyisophthalic) acid
and dehydrodiveratric acid. The formation of isohemipinic
acid supports occurrence of condensed structures (e.g., β-5
or γ-5). See structures 1 to 3 in Figure 1.6.
(b) Nitrobenzene oxidation of softwoods in alkali results in the
formation of vanillin (4-hydroxy-3-methoxybenzaldehyde).
Oxidation of hardwoods and grasses results respectively in
syringaldehyde (3,5 dimethoxy-4-hydroxybenzaldehyde) and
p-hydroxybenzaldehyde. See structures 4 to 6 in Figure 1.6.

COOH COOH COOH COOH

OCH3 HOOC OCH3 CH3O OCH3


OCH3 OCH3 OCH3 OCH3
1 2 3
CH2OH

CH2
COOH COOH COOH CH2

OCH3 CH3O OCH3 OH


OH OH OH OH

4 5 6 7

CH3 CH3
CH3 CH3
C=O C=O
HCOC2H5 C=O
HCOC2H5 CH2 C=O
C=O

OCH3 OCH3 OCH3


OCH3
OH OH
OH OH
8 9 10 11

Figure 1.6 Various degradation products of lignin.


Wood as a Unique Adherend 17

(c) Hydrogenolysis yields propylcyclohexane derivatives. See


structure 7 in Figure 1.6.
(d) Ethanolysis yields so-called Hibbert ketones. See struc-
tures 8 to 11 of Figure 1.6.

1.4.3.4 Dehydrogenation Polymer (DHP)


The biosynthesis of lignin from the monomeric phenylpropane units can be
generally described as a dehydrogenative polymerization. The principal ideas
about such a pathway were elaborated by Freudenberg and co-workers [11].
They were the first to produce in vitro lignin called dehydrogenation polymer
(DHP) by treating coniferyl alcohol with a fungal laccase from the mushroom
Psalliota campestris or with a horseradish peroxidase by hydrogen peroxide.
The first step of the biochemical pathway for building up lignin macro-
molecules is the enzymatic dehydrogenation of p-hydroxycinnamyalcohols,
yielding mesomeric ring systems with a loosened proton. Figure 1.7 shows
the formation of phenoxy radicals from coniferyl alcohol by a one-electron
transfer:
The origin of the hydrogen peroxide was cleared up by discovering
cell-wall-bound enzyme systems able to deliver H2O2 [22, 23].
Only 4-phenoxyradical I to IV are actually involved in lignin biosynthesis.
Structure V is sterically hindered or thermodynamically not favored [24].
The polymerization of monomeric precursors by random coupling reac-
tions cannot be studied in vivo, but it is known from numerous in vitro
experiments to run without enzymatic control as a spontaneous process.
The first step in polymerization is the formation of dimeric structures.
Some prominent lignol dimers called dilignols are shown in Figure 1.8.
Further polymerization is called end-wise polymerization involving
coupling of monolignols with the phenolic end groups of di- or oligolignols
or a coupling of two end group free radicals, yielding a branched polymer
via tri-, tetra-, penta-, and oligolignols [11].

γ CH2OH CH2OH CH2OH CH2OH CH2OH CH2OH

β HC HC HC HC HC HC

α CH CH CH CH CH CH
–(e + H )
1
6 2
3
5
4 OCH3 OCH3 OCH3 H OCH3 OCH3 OCH3
O H O O O O O

Figure 1.7 Enzymatic dehydrogenation of coniferyl alcohol yielding phenoxy radicals.


18 Adhesives for Wood and Lignocellulosic Materials

C C C C
C O C C C C
C C O C C C

O O O O
β-O-4 α-O-4 5-5
C
C
C C
C C
C C O
C C
C
C O C

O
O O

β-5 4-O-5 β-1

C C
C C
C C

O O

β-β

Figure 1.8 Typical dilignol structures [25].

Summarizing the formation of lignin, as mentioned by Fengel and


Wegener [11], it is evident that these macromolecules are not formed by
a genetically prescribed regular mechanism, but by a random coupling
of lignols to form a nonlinear polymer. The final constitution of lignin is
therefore determined mostly by reactivity and the frequency of the build-
ing units involved in its polymerization.
Proportions of different types of linkages connecting the phenylpropane
units in lignin are given in Table 1.1.

1.5 Influence of Hierarchical Structure of Wood


on Wood–Adhesive Interaction
Marra [27] describes the process of adhesive bond formation in a wood sub-
strate by five steps: flow, transfer, penetration, wetting, and solidification.
Wood as a Unique Adherend 19

Table 1.1 Proportions of different types of linkages connecting the phenylpropane


units in lignin.
Percent of the total linkages
Linkage typeb Dimer structure Softwooda Hardwooda
β-O-4 Arylglycerol-β-aryl ether 50 60d
α-O-4 Noncyclic benzyl aryl ether 2–8c 7
β-5 Phenylcoumaran 9–12 6
5-5 Biphenyl 10–11 5
4-O-5 Diaryl ether 4 7
β-1 1,2-Diaryl propane 7 7
β-β Linked through side chains 2 3
a
Approximate values based on data of Adler [26] obtained for MWL from spruce (Picea
abies) and birch (Betula verrucosa).
b
For corresponding structures.
c
Values have been reported [25].
d
Of these structures, about 40% are of guaiacyl type and 60% are of syringyl type.

The flow involves the spreading of the liquid on the wood surface. This is
followed by transfer of adhesive to the adjacent wood surface. Capillary
forces within the cell lumens promote penetration, and bulk flow occurs
due to applied pressure. Wetting of the wood surface by the adhesive occurs
to an optimum extent, which promotes the molecular contact between the
adhesive and wood surface. Finally, solidification occurs as a result of the
formation of three-dimensional cross-linked structure when the glue line
is exposed to high temperature.
Penetration of adhesive into the porous network of wood cells is
believed to have a strong influence on bond strength [28–30].
Damaged wood cells may be reinforced by the adhesive, and stresses
may be more effectively distributed within a larger interphase region. The
optimum depth of penetration is required to ensure mechanical adhesion
[31], but excessive penetration causes insufficient adhesive remaining at
the interface [27], leading to a starved bondline. This constitutes a weak
boundary layer and a weak spot in the chain (see Figure 2.7) of Marra’s
chain-link analogy [27] (see Section 2.15). This analogy emphasizes the
fact that the overall strength of an adhesive bond in a composite is deter-
mined by the weakest portion of the chain. In other words, an adhesive
20 Adhesives for Wood and Lignocellulosic Materials

bond is just as good as the weakest link in the chain. In this respect, adhe-
sive penetration plays a vital role in this analogy.

1.5.1 Penetration
The hierarchical structure of wood profoundly influences the adhe-
sion phenomenon over a wide range of “length scales”. The adhesive
phenomenon occurs first by transport phenomenon (bulk flow, pen-
etration, and diffusion) followed by a number of possible processes
ranging from mechanical adhesion to the formation of chemical bonds
as shown in Table 1.3. Adhesive penetration in wood is commonly cat-
egorized into (a) gross penetration and (b) cell wall penetration. Gross
penetration is the flow of liquid adhesive into the porous structure of
wood in order to fill the lumens. For cell wall penetration, the adhesive
enters the woody cell wall. Adhesive gross penetration and cell wall
penetration are both critical to the performance and durability of adhe-
sive joints [32–35].

1.5.1.1 Penetration in Different Size Scales


Laborie proposed a categorization of adhesive penetration in wood in
terms of different size scales of adhesive penetration [32]. Accordingly, in
the following classification, four scales of penetration are envisioned:

(a) Macroscopic penetration (millimeters)


(b) Microscopic penetration (microns to tens of microns)
(c) Nanoscale penetration (nanometers to tens of nanometers)
(d) Angstrom scale penetration (up to tens of angstroms)

Wood as a porous, cellular material displays roughness on the micron


scale but can also exhibit roughness on the millimeter scale, depending on
how a particular wood element to be bonded is produced. For example,
production of rotary-peeled veneer can produce roughness on a millimeter
scale due to the creation of lathe checks. Pores or free volume also occur
within the amorphous regions of the cell wall material on the molecular
level. The length scales over which the wood–adhesive interactions occur
are given in Table 1.2 [36].
In order to understand the adhesion phenomenon operating at different
size scales in wood, the values of size scales shown in Table 1.2 should be
viewed in conjunction with the length scale (as shown in Table 1.3) [37]
over which different wood–adhesive interactions take place.
Wood as a Unique Adherend 21

Table 1.2 Comparison of wood–adhesive interactions relative to length scale.


Component μm nm
Adhesive force 0.0002–0.0003 0.2–0.3
Cell wall pore diameter 0.0017–0.002 1.7–2.0
PF resin molecular length 0.0015–0.005 1.5–5.0
Diameter of particles that can 0.2 200
pass through a pit
Tracheid lumen diameter 4–25
Glue line thickness 50–250

Table 1.3 Comparison of adhesion interactions relative to length scale [36].


Category of adhesion
mechanism Type of interaction Length scale
Mechanical Interlocking or entanglement 0.01–1000 μm

Diffusion Interlocking or entanglement 10 nm–2 nm

Electrostatic Charge 0.1–1.0 μm

Covalent bonding Charge 0.1–0.2 nm

Acid–base interaction Charge 0.1–0.4 nm

Lifshitz van der Waals Charge 0.5–1.0

In its most simple approach, adhesive penetration can be measured in


terms of depth of penetration. It typically spans from fractions of millime-
ters to several millimeters. One may therefore term this scale of penetra-
tion as macroscopic penetration.
A finer observation scale of adhesive domains can be understood by ref-
erence to wood anatomy. Hierarchically structured wood is an orthotropic
material, in which arrays of cells are aligned along the longitudinal and
radial directions (Figure 1.9).
In the case of wood cells, cell lumen dimensions thus fall on the micron
range. Perforation plates at the cell extremities and pits on the lateral
22 Adhesives for Wood and Lignocellulosic Materials

Figure 1.9 Arrays of cells aligned along the longitudinal and radial directions.

cell walls constitute additional cavities between adjacent cells. Such cav-
ities have diameters in the order of microns and may thus be termed
“micropores”. A microscopic porous network consequently imparts flow
pathways within a solid wood block. It follows that adhesive penetration
into cell lumens and “micropores” falls into the dimensional domain of
microscopic penetration. As expected, microscopic penetration is best
evaluated with microscopic techniques. Optical microscopy and scan-
ning and transmission electron microscopy (SEM/TEM) have been uti-
lized with success to probe adhesive microscopic penetration [31, 34,
35, 38].
Adhesive variables (MW distribution), substrate variables (wood sur-
face roughness and moisture content), and processing variables (adhesive
cure method) influence microscopic penetration of adhesives in wood [31,
38]. Wood surface energy and adhesive surface tension are also important
parameters of micron-scale penetration [39].

1.5.2 Other Wood-Related and Process-Related Factors


Other wood-related and process-related factors that have an influence on
adhesive penetration are direction of penetration with respect to the wood
structure, permeability, porosity, roughness, surface energy, temperature,
pressure, and time [27, 40–43].

1.6 Effect of Hierarchical Structure of Wood


on Adhesive Penetration
The flow of liquid resin into the porous structure of wood results in filling
the cell lumens. The gross penetration occurs when resin diffuses into the
cell wall or flows into micro fissures. Hydrodynamic flow is caused by an
Wood as a Unique Adherend 23

external compression force due to clamp pressure or pressure exerted by


the hydraulic press. Flow then proceeds into the interconnected network of
lumens and pits, with flow moving primarily in the path of least resistance
[44].
The flow path of the liquid adhesive into the wood is determined by
the resistance to the hydrodynamic flow. The least resistance is in the lon-
gitudinal direction through the lumens in the long and slender tracheids
of softwoods, or through the vessels of hardwoods. Since vessels are con-
nected end to end with perforation plates and no pit membrane, this cell
type dominates the penetration of adhesives in hardwoods. Using optical
microscopy, Kamke and Lee observed the presence of resin in pit chambers
of both hardwood and softwood species and in cell lumens as a result of
entry of the resin through a pit [44].
The rheology of the adhesives plays a significant role in the adhesive
penetration. Polymeric adhesives exhibit non-Newtonian behavior [45].
Also, the capillary pathways through cell lumens and pits are tortuous. As
a result, penetration of the adhesives through wood is further complicated.
Waterborne adhesives, such as the phenolics and amino resins, are het-
erogeneous with unique distribution of MWs. They are prone to separation
(1) when the water is absorbed by the cell wall and (2) when the high MW
polymer molecules are trapped by the pit membrane into different depths
depending on the MW of the polymer (comparable to gel permeation
chromatography). Because of the above wood-related factors, the influence
of fluid dynamics of the adhesives on its penetration into wood cannot
be generalized. Gross penetration can happen with most types of resin at
low viscosity, while cell wall penetration can occur only with resins having
small MW components. In order to determine what should be the critical
molecular size of the adhesives below which they can penetrate into the
cell wall, Tarkow et al. [46] studied the critical MW of polyethylene glycol
(PEG) needed to penetrate the cell wall of Sitka spruce. Their study showed
that the critical MW of PEG was 3000 at room temperature. MW fractions
less than 3000 are common in the case of phenolic, amino resins, and iso-
cyanate resins and therefore these resins can be expected to penetrate into
the cell wall. Further, prior to polymerization, adhesives penetrating the
cell wall swell and plasticize the wood. This is an additional factor in favor
of penetration of these resins into the cell wall. This has been reported for
pMDI and low MW PF adhesives [32, 47, 48].
While discussing the importance of gross and cell wall penetration
with respect to the adhesion mechanisms, Frihart proposed four scenar-
ios [49]. In the first case of micrometer penetration, the adhesive occu-
pies the free volume within the cell wall—thus inhibiting shrinking and
24 Adhesives for Wood and Lignocellulosic Materials

swelling. In the second case, there can be mechanical anchoring between


the cured adhesive and the substrate. The third case is the formation of
interpenetrating polymer network that is made up of the cross-linked
adhesive within the free volume of the cell wall. The last case claims
the formation of chemical cross-links with the cell wall polymeric
components.

1.7 Wood Factors Affecting Penetration


The actual penetration depth of solution or adhesive depends on the per-
meability of wood to liquids, the technological methods, and physicochem-
ical properties of the specific adhesive. In the case of liquid flow through
porous material under ideal conditions with no interactions occurring
between the liquid and the porous material, the permeability is defined by
Darcy’s law [50]:

A 1
Q K ΔP
L

where Q is the liquid volume flow [m3 s−1], K is the specific permeability
of wood [m2], A is the area perpendicular to the liquid flow [m2], L is the
sample length in the direction of flow [m], η is the dynamic viscosity of the
liquid [Pa s], and ΔP is the pressure gradient [Pa]. As described by Darcy’s
law, the pressure gradient ΔP is the cause for the liquid penetration into
wood.
The permeability and surface energy are the two wood-related factors
controlling adhesive penetration [44]. Permeability varies with species and
direction (e.g., tangential, radial, and longitudinal). However, longitudi-
nal permeability may be as much as 104 times greater than transverse per-
meability [51]. Wood species with low permeability, such as Douglas-fir
heartwood, severely restricts resin penetration in the radial and tangential
directions. High permeability of the wood surface may be problematic to
adhesive bonding if this leads to starvation at the bondline. Thus, bonding
endgrain is difficult [44]. There are earlywood and latewood differences,
as well as heartwood and sapwood differences. Pit aspiration sometimes
occurs in softwoods during drying [51], thus severely reducing permeabil-
ity. White [52] noted greater penetration of phenol-resorcinol into early-
wood than latewood cells of southern pine.
Wood as a Unique Adherend 25

1.8 Influence of Resin Type and Formulation


on Penetration
The penetration of adhesives in wood is influenced by the resin viscosity,
its MW and MW distribution, resin solid content, and the surface tension.
It has to be kept in mind that the curing process may lead to a change in the
above characteristics (e.g., viscosity, MW, etc.). Hence, penetration may be
influenced during the curing process. Additives that influence the curing
behavior may affect the penetration [44].
Hse reported a correlation between penetration and contact angle for
PF and southern pine wood [53]. The author employed 36 formulations
to determine the contact angle, and its influence on cure time, heat of
reaction, plywood shear strength, percent wood failure, bondline thick-
ness, and cure shrinkage. Penetration was not measured, but assumed to
be inversely proportional to bondline thickness (thickness of cured adhe-
sive between the veneers). Penetration increased with increasing caustic
content. There were no clear trends observed for penetration in relation to
adhesive solids content or formaldehyde–phenol mole ratio.
In the case of powdered adhesives, such as powdered PF used in OSB
manufacture, it has to melt before penetration. Johnson and Kamke [54]
noted that powdered PF resin remained on the surface of wood strands
during the blending process and was able to flow and penetrate only during
steam injection hot-pressing.
Frazier et al. [55] noted that low MW of pMDI resin would promote
penetration into wood cell walls. They further hypothesized that the
MDI forms an interpenetrating network of polyurea and biuret linkages
within the cell wall. Swelling of the cell wall by pMDI was also observed
[48].
The use of resin with low MW components has the potential for deeper
penetration than that with high MW. Stephen and Kutscha separated a
commercial PF resin into two MW fractions (approximately ±1000 MW)
[56]. They observed the penetration of the resin by placing a drop of resin
on the surface of aspen. The authors reported that no penetration through
the wood was observed for the high MW fraction of the resin, while a
penetration of one to two cells deep occurred for the low MW PF frac-
tion. The addition of NaOH improved the penetration of the high MW
fraction, which the authors assumed was due to swelling of the cell wall
by the NaOH. The improved penetration may also have been due to a
lower viscosity as a result of NaOH addition [57]. Gollob et al. reported
that PF bond performance was associated with penetration in Douglas-fir
26 Adhesives for Wood and Lignocellulosic Materials

plywood [58]. They noted that higher MW formulations tended to dry out
and had little penetration.
Zheng studied the penetration of the blends of MDI and PF into yellow-
poplar and southern pine [59]. The penetration of the adhesive blends
was characterized by a phase separation, with pMDI penetrating deeper.
PF tended to bulk the lumens and remain at the interface of the bondline.
In general, the blends resulted in less penetration than either of the neat
resins. The author attributed the reduction in penetration to increased
MW, and subsequent increased viscosity, due to the formation of ure-
thane bonds between the PF and the PMDI.
MW distribution of resin systems will impact their ability for cell wall
penetration. Laborie [32] reported evidence of cell wall penetration for two
PF formulations, one that had a number average Mn of 270 and a weight
average Mw of 330. The other PF had Mn and Mw values of 2840 and 14,200,
respectively. The more highly condensed PF resin had a broad MW distri-
bution, including a low Mw component that was similar to the low MW PF
resin. Using dynamic mechanical analysis, the author concluded that both
resin systems penetrated the cell wall.

1.9 Effect of Processing Parameters on Penetration


Processing factors that can influence the adhesive penetration are open
assembly time, pressing time, temperature, and consolidation pressure
involved in wood-based composite manufacture. There can be a complex
interaction between processing parameters, adhesive formulation, and
wood characteristics. Sernek et al. reported increasing penetration of
UF resin into beech as open assembly was increased [31]. Temperature
influences penetration by affecting resin viscosity and cell wall permea-
bility. At the same time, in the case of thermosetting adhesives, polym-
erization increases viscosity, countering the temperature effect on liquid
viscosity.
White studied the influence of consolidation pressure (from 3 to
1000  kPa) on penetration and consequently on the fracture toughness
of southern pine blocks bonded with resorcinol-formaldehyde [52].
Increasing consolidation pressure increased penetration into earlywood,
but had an erratic effect on latewood. Low permeability of the latewood
may result in the adhesive squeezing of the adhesive out of the bondline
during consolidation. Increasing consolidation pressure reduced fracture
toughness of the latewood specimens but had no significant influence on
the earlywood specimens.
Wood as a Unique Adherend 27

Johnson and Kamke [54] reported increased penetration of PF into


yellow-poplar strands as a result of steam injection pressing. The authors
suggested that condensate from the steam dilutes the resin and steam
pressure forces the resin deeper into the wood.
The use of radio-frequency heating of a veneer composite caused a
reduction in penetration of UF resin in comparison to matched samples
produced using conduction heat in a platen [31]. The authors noted that
the rate of polymerization was much faster using radio-frequency heating
and thus reduced the time for penetration.

References
1. Gao, H., Learning from Nature about Principles of Hierarchical Materials.
Nanoelectronic Conference (INEC) 3rd International Conference, Hongkong,
China, 4 March 2010, 2010.
2. Lakes, R., Materials with structural hierarchy. Nature, 361, 511–515, 1993.
3. Moon, R.J., Frihart, C.R., Wegner, T., Nanotechnology applications in the
forest products industry. For. Prod. J., 56, 5, 4–10, 2006.
4. Gardner, D.J., Blumentritt, M., Wang, L., Yldirim, N., Adhesion theories in
wood adhesive bonding, in: Progress in Adhesion and Adhesives, K.L. Mittal
(Ed.), pp. 125–168, Scrivener Publishing, 2015.
5. River, B.H., Vick, C.B., Gillespie, R.H., Wood as an adherend, in: Treatise
on Adhesion and Adhesives, vol. 7, J.D. Minford (Ed.), pp. 131–133, Marcel
Dekker, New York, 1991.
6. Wiedenhoeft, A.C., Structure and function of wood, chapter 3, in: Wood
Handbook, U.S. Department of Agriculture, Forest Service, Forest Products
Laboratory, Madison, WI, General Technical Report FPL-GTR-190, 3-1-3-
18, 2010.
7. Hillis, W.E., Formation of robinetin crystals in vessels of Intsia species. IAWA
J., 17, 405–419, 1996.
8. Wheeler, E.A., Baas, P., Gasson, P.E., IAWA list of microscopic features for hard-
wood identification, vol. 10, pp. 219–332, IAWA Committee, Rijksherbarium,
1989.
9. Esau, K., Anatomy of seed Plants, 2nd edn., John Wiley & Sons, New York,
1977.
10. Dickison, W., Integrative Plant Anatomy, Academic Press, Orlando, 2000.
11. Fengel, D. and Wegener, G., Wood: Chemistry, Ultrastructure, Reactions,
2nd edn., Walter de Gruyter, Berlin, 1989.
12. Barber, N.F. and Meylan, B.A., The anisotropic shrinkage of wood: A theoret-
ical model. Holzforschung, 18, 146–156, 1964.
13. Brandstrom, J., Bardage, S.L., Nilsson, D.G.T., The structural organisation of
the S1 cell wall layer of Norway spruce tracheids. IAWA J., 24, 27–40, 2003.
28 Adhesives for Wood and Lignocellulosic Materials

14. Heyn, A., The ultrastructure of wood pulp with special reference to the ele-
mentary fibril of cellulose. Tappi, 60, 11, 159–161, 1977.
15. Gierlinger, N. and Burgert, I., Secondary cell wall polymers studied by con-
focal Raman microscopy: Spatial distribution, orientation, and molecular
deformation. N. Z. J. For. Sci., 36, 1, 60–71, 2006.
16. Fahlén, J., The Cell Wall Ultrastructure of Wood Fibres—Effects of the
Chemical Pulp Fibre Line, KTH Fibre and Polymer Technology, Stockholm,
Sweden, 2005.
17. Pettersen, R.C., The chemical composition of wood, in: The Chemistry of
Solid Wood, R.M. Rowell (Ed.), Advances in Chemistry Series 207, American
Chemical Soc., Washington, DC, 1984.
18. Goring, D.A.I. and Timell, T.E., Molecular weight of native cellulose. Tappi,
45, 454–460, 1962.
19. Sjöström, E., Wood Chemistry: Fundamentals and Applications, Academic
Press, Orlando, USA, 1993.
20. Oghbaie, M., Mirshokraie, S.A., Massoudi, A.H., Partovi, T., Opimisation of
lignin extraction. Mod. Chem., 2, 5, 36–40, 2014.
21. Hossain, Md. M. and Aldous, L., Ionic liquids for lignin processing:
Dissolution, isolation and conversion. J. Chem., 65, 1465–1477, 2012.
22. Elstner, E.F. and Heupel, A., Formation of hydrogen peroxide by isolated cell
walls from horseradish (Armoracia lapathifolia Gilib.). Planta, 130, 175–180,
1976.
23. Halliwell, B., Lignin synthesis: The generation of hydrogen peroxide and
superoxide by horseradish peroxidase and its stimulation by manganese (II)
and phenols. Planta, 140, 81–88, 1978.
24. Glasser, W.G., Lignin, in: Pulp and Paper Chemistry and Chemical
Technology, 3rd edn., J.P. Casey (Ed.), pp. 39–111, Wiley Interscience,
New York, 1980.
25. Wang, D., Wang’s Wood Chemistry Class, Basic Lignin Chemistry, David
Wang’s Wood Chemistry Class, National Chung Hsing University, Forestry
Dept. http://web.nchu.edu.tw/pweb/users/taiwanfir/lesson/10476.pdf.
26. Adler, E., Lignin chemistry: Past, present and future. Wood Sci. Technol., 11,
169–218, 1977.
27. Marra, A., Technology of Wood Bonding Principles in Practice, Van Nostrand,
Reinhold, New York, 1992.
28. Brady, E. and Kamke, F.A., Effects of hotpressing parameters on resin pene-
tration. For. Prod. J., 38, 63–68, 1988.
29. Collett, B.M., A review of surface and interfacial adhesion in wood science
and related fields. Wood Sci. Technol., 6, 1–42, 1972.
30. Jakal, L., Effect of the penetration of adhesive on the strength of adhesion.
Faipar, 34, 2, 59–60, 1984.
31. Sernek, M., Resnik, J., Kamke, F.A., Penetration of liquid urea–formaldehyde
adhesive into beech wood. Wood Fibre Sci., 31, 1, 41–48, 1999.
Wood as a Unique Adherend 29

32. Laborie, M.-P., Investigation of the wood/phenol-formaldehyde adhesive


interphase morphology. PhD thesis, Virginia Polytechnic Institute and State
University, Blacksburg, Virginia, 2002.
33. White, M.S., Influence of resin penetration on the fracture toughness of
bonded wood, PhD thesis, Virginia Polytechnic Institute and State University,
Blacksburg, Virginia, 1975.
34. Nearn, W.T., Application of the ultrastructure concept in industrial wood
products research. Wood Sci., 6, 285–293, 1974.
35. Nearn, W.T., Wood–Adhesive Interface Relations. Conference on the Wood–
Paint Interface, Berkeley, October, 1964.
36. Gardner, D.J., Chapter 19, in: Adhesion Mechanisms of Durable Wood
Adhesive Bonds, D.D. Stokke and L.H. Groom (Eds.), Wiley, New York
BLIO009-Stokke, September 13, 2005, 2006.
37. Sellers, T., Jr., Adhesives in the wood industry, in: Handbook of Adhesive
Technology, 1st edn, A. Pizzi and K.L. Mittal (Eds.), pp. 599–614, Marcel
Dekker, New York, 1994.
38. Johnson, S.E. and Kamke, F.A., Quantitative analysis of gross adhesive pene-
tration in wood using fluorescence microscopy. J. Adhes., 40, 47–61, 1992.
39. Gray, V.R., The wetting, adhesion and penetration of surface coatings on
wood. JOCCA, 44, 756–786, 1961.
40. Hare, D.A. and Kutscha, N., Microscopy of eastern spruce plywood gluelines.
Wood Sci., 6, 294–304, 1974.
41. Kedziers, A., Effect of adhesive penetration into wood on the strength of
glued joints. Przemsyl-Drzewny, 37, 3, 18–20, 1986.
42. Smith, L.A., Resin penetration of wood cell walls—Implications for adhesion of
polymers to wood. PhD thesis, pp. 172–176, Syracuse University, University
Microfilms, Ann Arbor, Michigan, 1971.
43. Tarkow, H. and Southerland, C.E., Interaction of wood with polymeric mate-
rials: I. Nature of absorbing surface. For. Prod. J., 14, 184–186, 1964.
44. Kamke, F.A. and Lee, J.N., Adhesive penetration in wood—A review. Wood
Fiber Sci., 39, 205–220, 2007.
45. Levenspiel, O., Engineering Flow and Heat Exchange, Plenum Press, New
York, 1984.
46. Tarkow, H., Feist, W.C., Southerland, C.F., Penetration versus molecular size—
Interaction of wood with polymeric materials. For. Prod. J., 16, 10, 61–65, 1966.
47. Marcinko, J.J., Devathala, S., Rinaldi, P.L., Bao, S., Investigating the molecu-
lar and bulk dynamics of PMDI/wood and UF/wood composites. For. Prod.
J., 48, 6, 81–84, 1998.
48. Frazier, C.E., Isocyanate wood binders, in: Handbook of Adhesive Technology,
2nd edn, A. Pizzi and K.L. Mittal (Eds.), Marcel Dekker, New York, 2003.
49. Frihart, C.R., Adhesive interaction with wood, in: Fundamentals of Composite
Processing, pp. 29–53, Madison, WI, Nov. 5–6, 2003. Gen. Tech. Rep. FPL-
GTR-149. USDA, For. Serv., For. Prod. Lab, 2004.
30 Adhesives for Wood and Lignocellulosic Materials

50. Kučerová, I., Methods to measure the penetration of consolidant solutions


into dry wood. J. Cult. Heritage, 135, 5191–5195, 2012.
51. Siau, J.F., Wood: Influence of Moisture on Physical Properties, Virginia Tech,
Blacksburg, VA, Pub. No. 7282 Forest Products Society, Madison, WI, 1995.
52. White, M.S., Influence of resin penetration on the fracture toughness of
wood adhesive bonds. Wood Sci., 10, 1, 6–14, 1977.
53. Hse, C.-H., Properties of phenolic adhesives as related to bond quality in
southern pine plywood. For. Prod. J., 21, 1, 44–52, 1971.
54. Johnson, S.E. and Kamke, F.A., Characteristics of phenol-formaldehyde
adhesive bonds in steam injection pressed flakeboard. Wood Fiber Sci., 26,
259–269, 1994.
55. Frazier, C.E., Schmidt R, R., Ni, J., Towards a Molecular Understanding of
Wood–Isocyanate Adhesive Bondline. Proceedings 3rd Pacific Rim Bio-Based
Composite Symposium, Kyoto, Japan, Dec. 2–5, pp. 383–391, 1996.
56. Stephen, R.S. and Kutscha, N.P., Effect of resin molecular weight on bonding
flakeboard. Wood Fiber Sci., 19, 353–361, 1987.
57. Gollob, L., The correlation between preparation and properties in phenolic
resins, in: Wood Adhesives Chemistry and Technology, vol. 2, A. Pizzi (Ed.),
pp. 121–153, Marcel Dekker, New York, 1989.
58. Gollob, L., Krahmer, R.L., Wellons, J.D., Christiansen, A.W., Relationship
between chemical characteristics of phenol–formaldehyde resins and adhe-
sive performance. For. Prod. J., 35, 3, 42–48, 1985.
59. Zheng, J., Studies of PF resole/isocyanate hybrid adhesives. PhD thesis, 198 pp.,
Virginia Polytechnic Institute and State University, Blacksburg, VA, 2002.
2
Fundamentals of Adhesion

2.1 Introduction
Adhesion is a multidisciplinary science involving various subjects such
as rheology, materials science, organic chemistry, polymer science, and
mechanics. Study of fundamentals of adhesion is essential as it leads to bet-
ter understanding of the factors controlling the performance of the bonded
assemblies [1].
Wood is a complex substrate, and it is hard to understand why some
adhesives work better than other adhesives, especially under stringent
durability tests.
The recent trend in the wood industry is to use smaller-diameter logs
and employ other lignocellulosic raw materials to produce more versatile
and environmentally acceptable green engineered wood products. This in
turn increases the complexity in the choice of adhesives. In order to pro-
vide a scientific basis to make the correct choice of adhesives and their
formulations, a study of the fundamentals of adhesion is essential.
A clear understanding of wood adhesion mechanisms will enable pro-
duction of better adhesive and formulation systems suitable for a wider
array of wood composite materials. The study of the fundamentals of
wood adhesion is essentially distinctive and unique and involves multi-
disciplinary sciences with respect to both the adherend, the adhesives, and
their interactions. The uniqueness of wood as an adherend by virtue of
possessing a hierarchical structure has already been dealt with in detail in
Chapter 1. In this respect, wood differs significantly from other substrates
such as metals, plastics, elastomers, etc. Surface science, rheology, materi-
als science, surface chemistry and surface morphology, organic chemistry,
polymer science and polymer characterization, and solid mechanics and
interaction between polymers and wood—all contribute to the develop-
ment and understanding of the adhesion phenomenon.

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (31–59) © 2019
Scrivener Publishing LLC

31
32 Adhesives for Wood and Lignocellulosic Materials

Further, these studies will enable

• The identification of practical problems and root causes for


adhesion failure and provide practical preventive solution
from the knowledge of the nature of wood–adhesive inter-
action, i.e., a scientific approach to troubleshooting.
• Optimization of the performance of existing adhesives and
to develop new adhesives to meet stringent environmental
regulations.
• The increase in the durability of bonded wood products by
precisely understanding the role of internal and external
stresses to which wood bond lines are subjected.
• The development of new technologies based on the insights
gained from the knowledge of the basic principles, which
can be applied efficiently for bonding difficult to bond
wood species and preservative treated wood. This is so that
the carbon sequestration is possible for prolonged periods
of time to reduce the global warming potential of wood
products.

2.2 Definitions
We should first define the terms adhesive, adhesion cohesion, and other
related terms in order to understand their individual role in determining
the effectiveness of bonding.

2.2.1 Adhesion
Adhesion is defined as the state in which two surfaces are held together by
interfacial forces that may consist of valence forces or interlocking action,
or both. Adhesion is further classified as mechanical adhesion and specific
adhesion. Specific adhesion between two surfaces is caused by the valence
forces of the same type as those that give rise to cohesion, as opposed to
mechanical adhesion in which the adhesive holds the parts together by an
mechanical interlocking.

2.2.2 Cohesion
Cohesion is defined as the internal strength of an adhesive as a result of a
variety of interactions within the adhesive.
Fundamentals of Adhesion 33

2.2.3 Adhesive
ASTM defines an adhesive as a substance capable of holding materials
together by surface attachment.

2.2.4 Adherend
Adhered, also called a substrate, is defined as a body that is held to another
body by an adhesive used interchangeably. Various descriptive adjectives are
applied to the term adhesive to indicate certain characteristics as follows:
(1)  physical form, that is, liquid adhesive, tape adhesive, etc.; (2)  chemical
type, that is, silicate adhesive, resin adhesive, etc.; (3) materials bonded, that is,
paper adhesive, metal–plastic adhesive, can label adhesive, etc.; (4) condition
of use, that is, hot setting adhesive, room temperature setting adhesive, etc.

2.2.5 Bonding
Bonding is the joining of two substrates using an adhesive. According to
DIN EN 923, an adhesive is defined as a non-metallic binder that acts via
adhesion and cohesion. ASTM D907-06 defines an adhesive as “a substance
capable of holding materials together by surface attachment”. A material
attached using adhesive is called an adherend.

2.2.6 Adhesive, Assembly


Adhesive, assembly—an adhesive that can be used for bonding parts
together, such as in the manufacture of a boat, airplane, furniture, and the
like. Note: The term assembly adhesive is commonly used in the wood
industry to distinguish such adhesives (formerly called “joint glues”) from
those used in making plywood (sometimes called “veneer glues”).

2.3 Mechanism of Adhesion


The role of an adhesive for wood is to transfer and distribute loads between
components, thereby increasing the strength and stiffness of wood prod-
ucts [6].
This is achieved through the following three basic types of adhesion:

1. Specific Adhesion—Bonding between the adhesive and the


adherend is due to chemical reaction.
34 Adhesives for Wood and Lignocellulosic Materials

2. Mechanical Adhesion—occurs due to mechanical anchorage.


3. Effective Adhesion—combines specific and mechanical
adhesion for optimum joining strength.

One should distinguish between adhesion and cohesion.


Cohesion as defined earlier is the attraction of molecules and groups
within the adhesive (or other material) that holds the adhesive molecules
together. The combination of adhesion and cohesive strength determines
the bonding effectiveness. An adhesive bond fails if either the adhesive
separates from the substrate (interfacial adhesion failure) or the adhesive
breaks apart (cohesive failure). The adhesive and cohesive strengths of
some adhesives are high enough that the cohesive strength of the substrate
fails before the adhesive bond.

2.3.1 Specific Adhesion


Specific adhesion involves the bond created by chemical means, rather
than mechanical, as a result of the molecular attraction between the sur-
faces in contact. This can be ionic, covalent, or induced by any other inter-
molecular forces (Figure 2.1), as described below:

(a) Coulombic (ionic) or hydrogen bonding


Hydrogen bonds occur in molecules that have H–F, H–O, and
H–N bonds. Basically, this strong intermolecular force is due to
strong dipole–dipole forces.
Besides the above, there can exist non-covalent and non-
electrostatic interactions (apolar interactions) between neu-
tral atoms and molecules [2, 3]. However, they are not as
strong as Coulombic (ionic) or hydrogen-bond interactions.
They are ubiquitous and are always attractive between like
particles.
(b) Apolar interactions
There are three types of intermolecular forces that occur in
chemical compounds. These forces cause molecules or groups of
molecules to be attracted to one another, thus affecting many of
their properties. Collectively known as the van der Waals forces,
these electrodynamic intermolecular forces originated from
three distinct interactions. These are (a) Keesom (permanent–
permanent dipoles) interaction (b) Debye (permanent-induced
dipoles) force, and (c) London dispersion force (fluctuating
Fundamentals of Adhesion 35

+40

Repulsion
+30
+20 Van-der-Waals forces
+10 1 2 3 4 5 6
0
Distance r
–10
[0, 1 nm]
–20
Hydrogen bonds
Attraction

–30
–40
Chemical bonds
–50
–60

Figure 2.1 Potential energy diagram for different forces [4].

dipole-induced dipole interaction) [2]. While these three kinds


of interactions have distinct origins, they have in common the
fact that their interaction energies decay rapidly with the sixth
power of the interatomic or molecular distance. See Sections
2.3.1.1 to 2.3.1.3.
The London dispersion force is the weakest, followed in
increasing strength by dipole–dipole forces and then hydro-
gen bonding. Lewis acid–base interactions can also occur (dis-
cussed later) [3].
The mathematical relationships for the various potential
energies are given below:

2.3.1.1 London Dispersion Force

3 2I
V
4 r6

where α is the polarizability, r is the distance, and I is the first ionization


potential.
The negative sign indicates the attractive interaction.
36 Adhesives for Wood and Lignocellulosic Materials

2.3.1.2 Dipole–Dipole Interaction


Dipole–dipole interaction is between polar molecules. A polar molecule
has an electric dipole moment by virtue of the existence of partial charges
on its atoms. Opposite partial charges attract one another, and, if two polar
molecules are oriented so that the opposite partial charges on the mole-
cules are closer together, then there will be a net attraction between the two
molecules with a potential energy V given by

2 2
2 A B 1
V 2 6
3 (4 0 ) r k BT

μi are the dipole moments


ε is the permitivity of the medium
T is the temperature in Kelvin

2.3.1.3 Dipole–Induced-Dipole Interaction


In the dipole–induced-dipole interaction, the presence of the partial charges
of the polar molecule causes a polarization, or distortion, of the electron
distribution of the other molecule. As a result of this distortion, the second
molecule acquires regions of partial positive and negative charge, and thus
it becomes polar. The partial charges so formed behave just like those of a
permanently polar molecule and interact favorably with their counterparts
in the polar molecule that originally induced them. Hence, the two mole-
cules cohere with a potential energy V given by

2 2
V
r6

where μ is the dipole moment of the polar molecule, α is the polarizability


of non-polar molecule, and r is the distance between them.

2.3.1.4 Ion–Dipole Interaction


An ion-dipole force is an attractive force that results from the electrostatic
attraction between an ion and a neutral molecule that has a dipole. It is
most commonly found in solutions. It is especially important for solutions
Fundamentals of Adhesion 37

of ionic compounds in polar liquids. The potential energy of ion–dipole


interaction is given by

q
V
(4 0 )r 2

where q is the charge on the ion.


One should note that in all the above equations describing the inter-
molecular attractions, the denominator contains the factor r6. Thus, the
types of intermolecular interactions described above occur only at very
small distances, of the order of typical atomic bond lengths (the range of
non-bonding interactions is between 0.3 and 0.5 nm). For interactions to
occur, therefore, the two materials must be able to make intimate contact
with each other (i.e., they must be able to approach within a nanometer).
This is possible if the adhesive wets the substrate efficiently. The types of
interactions and the corresponding energies are given in Table 2.1.

Table 2.1 Bond types and typical bond energies [1].


Type of interaction Energy (kJ/mol) Basis of attraction
Bonding
Ionic 400–4000 Cation–anion
Covalent 150–1100 Nuclei–shared electron pair
Metallic 75–1000 Cations–delocalized electrons
Non-Bonding
Ion–dipole 40–600 Ion charge–dipole charge
Polar bond to hydrogen–dipole
Hydrogen bonding 10–40
charge
Dipole–dipole 5–25 Dipole charges
Ion charge–polarizable
Ion–induced dipole 3–15
electrons
Dipole charge–polarizable
Dipole–induced dipole 2–10
electrons
Interaction between polarizable
Dispersion forces 0.1–40
electrons
38 Adhesives for Wood and Lignocellulosic Materials

2.3.1.5 Hydrogen Bonds


This is an important intermolecular interaction specific to molecules con-
taining an oxygen, nitrogen, or fluorine atom that is attached to a hydro-
gen atom. This interaction is the hydrogen bond, an interaction of the
form A−H···B, where A and B are atoms of any of the three elements men-
tioned above and the hydrogen atom lies on a straight line between the
nuclei of A and B (Figure 2.1).

2.3.1.6 Ionic Bonds


Salts like NaCl.

2.3.1.7 Chemical Bonds


The acid–base character of the substrate may influence the reactivity between
adhesive and substrate. A covalent bond involves shared valence electrons
(Figure 2.1).

2.4 Theories of Adhesion


According to Schultz and Nardin (1994), the main adhesion theories are
as follows:

1. Mechanical interlocking
2. Electronic or electrostatic theory
3. Adsorption (thermodynamic) or wetting theory
4. Diffusion theory
5. Chemical (covalent) bonding theory
6. Theory of weak boundary layers and interphases

The adsorption hypothesis, which explains that adhesion is caused by


intermolecular forces such as van der Waals forces, hydrogen bonds, and
electrostatic interactions, is widely considered to be the most applicable
to wood–polymer adhesion [7]. However, in a porous material like wood,
penetration and mechanical interlocking must also play a significant role
in the bonding process.
Marra [5] described adhesive bond formation in wood-based panels as a
dynamic process consisting of flow, transference, penetration, wetting, and
solidification (cure).
Fundamentals of Adhesion 39

The mechanisms outlined above are not mutually exclusive since one or
more of the above mechanisms can occur simultaneously depending on
the specific conditions prevailing during bonding. The hierarchical cellular
characteristics of wood offer such varied conditions.
The mechanical interlocking theory has long been used to explain wood
bonding [6].
The electronic or electrostatic theory has been applied to wood in
finishing and coating operations, although this adhesion bonding
mechanism needs more fundamental research [21]. The adsorption or
wetting theory has been exhaustively studied on wood over the past 40
years [7, 8].
The diffusion theory is appropriate in wood bonding during the produc-
tion of compressed fibrous materials such as hardboard. The thermoplastic
matrix, namely, lignin, can soften beyond its glass-transition temperature
during the thermal conditions employed during hot pressing. Under these
conditions, lignin can diffuse throughout the fibrous mat and react with
the furfural liberated from hemicelluloses (pentosans) and solidify due to
chemical reaction and hence function as an adhesive.
Besides the diffusion and molecular interpenetration of lignin occurring
during wet process in the hardboard production as mentioned above, there
is also the phenomenon of diffusion of monomers/oligomers of synthetic
resin adhesives such as PF or UF into the wood cells followed by subsequent
polymerization. This is an important concept that speaks of monomers that
penetrate at a molecular level for thermosetting adhesives [9].
While discussing on the theories of adhesion in wood, one should keep
in mind the opposite process (debonding). Weak boundary layers have been
identified as the cause for the premature failure of the adhesive bond. In the
case of wood bonding, the theory of weak boundary layers has also been
proposed and studied. The weak boundary layers can be caused as a result of
the mechanical damages occurring during the machining of wood surfaces.
Further, the impact of surface aging the consequent inactivating of the wood
surfaces [10–12] can also be responsible for the weak boundary layer.

2.4.1 Mechanical Theory


McBain and Hopkins [13] first proposed the concept of “mechanical
adhesion” in their classical paper “On adhesives and adhesive action”.
According to McBain and Hopkins, there are two kinds of adhesion:
mechanical and specific adhesion. Specific adhesion involved interaction
between the adherend surface and the adhesive. This interaction might be
chemical interaction or adsorption.
40 Adhesives for Wood and Lignocellulosic Materials

Mechanical adhesion occurs “whenever a liquid adhesive penetrates


into the porous adherend and solidifies in situ in the pores”. Examples are
adhesion to wood, unglazed porcelain, pumice, and charcoal.
The surface of a substrate is never truly smooth but consists of a maze
of peaks and valleys. This type of topography allows adhesive to penetrate
and fill these valleys, displace the entrapped air, and secure mechanically in
position inside the substrate similar to the operation of the Velcro.
Porosity and roughness of the substrate increase the total area of contact
between the adhesive and the adherend. Hence, roughening the adherend
surface enhances the mechanical interlocking since total effective area
over which the forces of adhesion can develop increases. The mechanical
adhesion theory does not take into account the intrinsic incompatibility
between the adhesive and the substrate.

2.4.1.1 lllustration of Mechanical Adhesion for Wood


In wood adherends, there is a vast array of void spaces as shown in Figure
2.2. Spontaneous surface wetting and capillary effects allow the flow of
the adhesive resin into the cell lumen, vessels, or other interstices fol-
lowed by subsequent hardening of the resin and resulting in mechanical
interlocking. The resin acts to reinforce the surface/interface layers of
wood cells. An adhesive penetration of approximately 6–10 cell diame-
ters (fewer than 100 μm, maximum) is regarded as necessary for optimal
adhesive bonding. Filling the cell lumen with adhesive provides much
larger mechanical interlocks than are available with surface roughness

Rays
Earlywood
Tracheids

Latewood
Tracheids

Figure 2.2 Various wood elements.


Fundamentals of Adhesion 41

for other substrates. Absorption into the cell wall can provide microme-
chanical interlocks and interpenetrating networks [1].

2.5 Electronic Theory


This theory was mainly promoted by Deryaguin [14–16]. If the adhesive
and the substrate have different electronic band structures, there is likely
to be electron transfer on contact between the two surfaces. This results in
the promotion of a double layer of electrical charges at the adhesive sub-
strate interface. Electrostatic forces are formed at the adhesive–adherent
interface. This accounts for the resistance to separation. This theory gath-
ers support from the fact that electrical discharges have been noticed when
an adhesive is peeled from a substrate. Electrostatic adhesion is regarded
as a dominant factor in biological cell adhesion and particle adhesion. No
application of this theory to wood appears possible.
The electronic theory

• Depends on material properties that allow electron transfer


across the interface
• Requires intimate contact/smooth surfaces
• Interactions are very weak and rather insignificant
• Mechanism is not important for wood substrates

2.6 Diffusion Theory


The diffusion theory was proposed in the early 1960s by Voyutskii [17–19].
It states that the intrinsic adhesion of a resin to a polymeric substrate is due
to mutual diffusion of polymer molecules across their interface.
As a result of this interdiffusion of molecules of the adhesive and adher-
end, their interface disappears. Hence, the diffusion theory is applicable
only when both adhesive and adherend are compatible polymers that pos-
sess sufficient mobility and mutual solubility. Solvents or heat welding of
thermoplastic substances is caused by diffusion.
The prerequisite of the diffusion theory is that the polymers of the
adhesive and of the substrate should possess similar values of solubility
parameters.
Several problems are encountered when an attempt is made to apply
the diffusion theory to wood. Basically, wood is not homogeneous in com-
position. It is a cellular composite of three polymers, namely, cellulose,
42 Adhesives for Wood and Lignocellulosic Materials

hemicelluloses, and lignin. Furthermore, cellulose consists of both crys-


talline and amorphous regions. It is clear then from solubility parameter
concepts that some polymers, the amorphous ones such as hemicelluloses
and lignin and the amorphous portion of cellulose, could, under some
conditions, undergo mutual diffusion with the polymer chains of the syn-
thetic adhesives. The crystalline portion of cellulose is not likely to be
involved.
There is one specific instance in the case of wood adhesion [17] in which
interdiffusion appears to exist and is likely to play a significant role in wood
bonding. This is the production of fiberboard by the wet process in which
no adhesive is added. At high moisture content, high temperature and pres-
sure and long pressing times, the glass transition temperatures of lignin are
exceeded. Thus, the lignin in the fibers is mobilized and the interdiffusion
between lignin polymers on different fibers contributes to the bonding of
the fibers together.

2.7 Adsorption/Covalent Bond Theory


The adsorption theory of adhesion, and the most widely accepted one,
in wood science, which is sometimes also called “specific adhesion”
[20], states that an adhesive will adhere to a substrate because of inter-
molecular and interatomic forces between the atoms and molecules of
the two materials. The interatomic and intermolecular forces referred
to can be any type of either primary or secondary valency forces. van
der Waals forces, hydrogen bond, and electrostatic forces are as much
applicable as the primary valence forces such as ionic, covalent metal-
lic coordination bonds. In the case of wood adhesion, however, there
is an age-old mistaken notion that covalent linkages must be present
to ensure good joint strength. In fact, covalent bonding theory was
invoked to explain the durable wood bonding with thermosetting adhe-
sives. But as mentioned by Gardner [21], it is very likely that covalent
bonds between the wood and adhesive are not necessary for durable
wood adhesive bonds.
Calculations carried out by a number of authors based on the secondary
forces involved indicate that the wood bond strength in tension should be
over 100 MPa. This is considerably higher than the experimental values
obtained in the case of several wood adhesives. This discrepancy could be
due to the presence of voids, defects, and the geometrical features of the
test specimen. Pizzi concludes that these studies indicate that the second-
ary valency forces themselves are adequate to explain the practical results
Fundamentals of Adhesion 43

Table 2.2 Comparison of adhesion interactions relative to length scale.


Category of adhesion
Mechanism Type of interaction Length scale
Mechanical Interlocking or entanglement 0.01–1000 μm
Diffusion Interlocking or entanglement 10 nm–2 mm
Electrostatic Charge 0.1–1.0 μm
Covalent bonding Charge 0.1–0.2 nm
Acid-base interaction Charge 0.1–0.4 nm
Lifshitz van der Waals Charge 0.5–1.0 nm

and it is not necessary to invoke the involvement of covalent bonds [20].


An elaborate discussion on the relative importance of the primary and sec-
ondary valence forces has been furnished by Pizzi based on the adhesive
strengths obtained from wood joints and the common wood adhesives
such as phenolics, amino resins, and isocyanates [20].

2.8 Adhesion Interactions as a Function of Length


Scale
It is useful to know the scale of lengths over which the adhesion inter-
actions as described above do occur (Table 2.2) [21]. It is apparent
from Table 2.2 that the adhesive interactions relying on interlocking or
entanglement can occur over larger lengths than the adhesion involv-
ing charge interactions. Most of the charge interactions occur at the
molecular level or the nano-length scale. Electrostatic interactions are
the exception to this generalization. For the purpose of adhesive inter-
actions, they are considered to operate from a nano- to a micron-length
scale.

2.9 Wetting of the Substrate by the Adhesive


It is evident from the previous section that various interactions operate
effectively only when the molecules of the adhesives come as close as pos-
sible to those of the substrate in order that such a proximity will lead to
44 Adhesives for Wood and Lignocellulosic Materials

α
α

(a) (b)

Wetting Dewetting
Adhesion Forces > Cohesive Forces Adhesion Forces < Cohesive Forces
Spreading of the liquid on the The liquid pulls itself together into the
surface of the solid. shape of a droplet.
Contact angle θ: 0 < θ < π/2 Contact angle θ: π/2 < θ < 0

Figure 2.3 Wetting phenomenon.

θ = 0° θ < 90° θ > 90°

Spreading Wetting Dewetting

Figure 2.4 Wetting, spreading, and dewetting for different contact angles.

maximum mutual interaction. Such closeness is possible only when the


adhesives wet the substrate.
Wetting is the ability of liquids to form interfaces with solid surfaces.
To determine the degree of wetting, the contact angle (θ) that is formed
between the liquid and the solid surface is measured. The smaller the con-
tact angle and the smaller the surface tension, the greater the degree of
wetting (Figure 2.3).
For maximum adhesion, the adhesive must completely cover the sub-
strate, i.e., spreading is necessary. The contact angle is a good indicator of
adhesive behavior. This is illustrated in Figure 2.4.

2.10 Equilibrium Contact Angle


In 1805, Thomas Young provided the first good approach for describing
wettability, spreading, and their relationship to the contact angle.
A drop of adhesive on a surface will come to equilibrium under the
action of three forces as shown in Figure 2.5.
Fundamentals of Adhesion 45

γLV

vapor

liquid θ
γSL
γSV
solid

γLV cosθ = γSV – γSL (Young equation)

γLV = liquid-vapor interfacial tension or surface tension


γSV = solid-vapor interfacial tension, not true surface energy
γSL = solid-liquid interfacial tension
θ = contact angle (angle liquid makes with solid surface)

Figure 2.5 Equilibrium contact angle based on balance of forces.

Considering the component of γLV along the X-axis, we can write the
following force balance:

LV cos SL SV

SV SL
Or cos (2.1)
LV

Thus, when θ = 0, the liquid spreads spontaneously on the substrate; in


other words, when cos θ is high (i.e., as it approaches 1), there is sponta-
neous spreading.
From Equation 2.1, it is clear that wetting will be favored when the sur-
face tension of the liquid is low.
Since the tendency of the liquid to wet and spread spontaneously increases
as the contact angle decreases, the contact angle is a useful inverse measure
of wetting or the cosine of the contact angle is a direct measure of wetting.

2.11 Thermodynamic Work of Adhesion


Perhaps the most convenient way of interpreting the wettability of a low
energy solid is the formulation of the work of adhesion, WA, defined by
46 Adhesives for Wood and Lignocellulosic Materials

Dupré and Dupré [22] as the work required to separate a unit area of the
solid–liquid interface.
Consider the wetting of a solid substrate (S) by a liquid (adhesive) “L”. A
solid–liquid interface is formed as a result according to the following equation:

S + L = SL (2.2)

If γS, γL, and γSL are the surface free energies of solid substrate, liquid
(adhesive), and the interphase, then the free energy change of the process
(ΔGA) can be written as

ΔGA = γSL – (γS + γL) (2.3)

The work of adhesion WA = −ΔG can be written as

WA = −ΔGA = (gs + γL) − γSL (2.4)

This is the thermodynamic work of adhesion or the work needed to


separate unit area of the solid–liquid interface.
Assuming γLV = γL and γSV = γs from Equations 2.1 and 2.4, we get

WA = γL (1 + cos θ) (2.5)

This is known as Young–Dupre’s equation. Thus, if the contact angle,


θ, of a well-defined probe liquid against a solid is measured, the work of
adhesion can be determined.
Thus, the thermodynamic work of adhesion (W) is, by definition, the
free energy change per unit area required to separate to infinity two surfaces
initially in contact with a result of creating two new surfaces at the inter-
face between two materials, for example, an adhesive and an adherend.
It is related to the intermolecular forces that operate
at the interface between two materials, for example, an adhesive and an
adherend.
It is related to the intermolecular forces that operate
at the interface between two materials, for example, an adhesive and
an adherend.
It is related to the intermolecular forces that operate
at the interface between two materials, for example,
It is related to the intermolecular forces that operate
at the interface between two materials, for example,
Fundamentals of Adhesion 47

Fowkes [23] proposed that both reversible work of adhesion (W) and
the surface free energy (γ) had additive components and can be parti-
tioned into individual components. Accordingly, several equations were
proposed based on this important approach. This pioneering development
of Professor Frederick M. Fowkes regarding the acid–base theory in adhe-
sion have attracted the attention of several laboratories. A Festschrift in his
honor on the occasion of his 75th birthday was published in 1991.
The approach is described below:

(1) Partitioning of surface free energies into components


The principle of partitioning is based on the assumption that
the surface free energy is determined by various interfacial
interactions. These interactions in turn depend on the basic
properties of the interacting liquid and that of the solid–liquid
interface (SL) [23, 24].

d p h i ab o
s s s s s s s
(2.6)

d p h
where s , s , s , is , ab
s , o
s

are the dispersion, polar, hydrogen (related to hydrogen bonds),


induction, and acid–base components, respectively, while o
refers to all remaining interactions.
(2) Mode of combinations of the individual energy components
According to Fowkes, the dispersion component of the sur-
face free energy is connected with the London interactions.
The remaining van der Waals interactions, i.e., the Keesom and
Debye ones, have been considered by Fowkes as a part of the
induction interactions.
Fowkes investigated mainly two-phase systems containing a
substance (solid or liquid) in which the dispersion interactions
appear only. Considering just such systems, Fowkes determined
the SFE corresponding to the solid–liquid interface as follows:
For two-phase systems comprising of a solid and liquid, in
which only dispersion interactions occur, namely, between
d d
s, 1,
Fowkes employed geometric mean as the mode of com-
bination of these energy components to give the following
equation:

d d d d 0.5 (2.7)
SL s L 2( s L )
48 Adhesives for Wood and Lignocellulosic Materials

Fowkes [25] modified Equation 2.7 by changing from geometric mean to


arithmetic mean to arrive at the following equation:
d d d d (2.8)
SL s L ( s L )

Owens and Wendt [26] significantly changed the Fowkes idea while
assuming that the sum of all the components occurring on the right-hand
side of Equation 2.11, namely
p h i ab o
s , s , s , s , s ,
p
except that γd can be considered as associated with the polar interaction s ,
Consequently, the following equation was obtained:

d d p p
SL s L 2 s L 2 s L (2.9)

Wu [27, 28] accepted the idea by Owens and Wendt to divide the SFE into
two parts, but used the harmonic means of the interfacial interactions instead
of the geometric means in Equation 2.9 and derived the following equation:

d d p p
s L s L
SL s L 4 d d p p
(2.10)
s L s L

van Oss, Chaudhury, and Good proposed the latest concept of partition
in which surface energy is partitioned into two components [29, 30]:

(1) Long range interactions London, Keesom, and Debye


called the Lifshitz–van der Waals component (γLW)
(2) The short-range interactions (acid–base), called the acid–
base component (γAB) = 2(γ+ γ–)0.5 where γ+ and γ– mean
the acidic and basic constituents, respectively, which are
associated with the acid–base interactions.

Combining van Oss Chaudhury’s concept with Young’s equation


(Equation 2.5), we obtain the Young–Fowkes–van Oss–Good acid–base
equation referred to as the acid–base approach:

0.5 0.5 0.5


LH LH
(1 cos ) L 2 S L S L S L (2.11)
Fundamentals of Adhesion 49

Table 2.3 Physical properties and surface free energy components of test liquids
used at 20°C [31].
Surface free energy (mJm–2)
(LW–AB approach)
Density Viscosity LW + AB
Liquid (kg/m3) (mPas) 1v 1v 1v 1v γ1v
Water 1000 1.00 21.8 25.5 25.5 51.0 72.8
Formamide 799 1.02 39 2.28 39.6 19 58
Ethylene glycol 1109 19.9 29 1.92 47 19 48
Diethylene glycol 1130 26.8 44.7 – – – –
Diodomethane 3325 2.8 50.8 0 0 0 50.8
1-Bromo- 1483 – 44.4 0 0 0 44.4
naphthalene

This equation contains three unknowns and, therefore, we need contact


angle data of three liquids. Table 2.3 contains the physical properties and
surface free energy components of test liquids normally used. One of them
must be non-polar and the other two should be polar. The Lifshitz–van der
Waals component of the surface free energy is obtained from the contact
angle of the apolar test liquid, e.g., diiodomethane.
The acid–base interaction and its relevance to adhesives and adhesive
bonding have been reviewed in detail by Chehimi et al. [32].

2.12 Spreading Coefficient


For spreading, another parameter, the spreading coefficient γSV − γSL − γLV
appears to be important in classifying liquids that have a tendency to form
good films on a given substrate. In general, the larger and more positive
value of the spreading coefficient (S), the more energy is gained by interca-
lating a liquid film between a solid and air. Thus,

S > 0, spontaneous spreading


S < 0, not spontaneous spreading

Though the condition S > 0 is necessary for a liquid to spread sponta-


neously on a solid, it is insufficient to describe the final state of the film.
50 Adhesives for Wood and Lignocellulosic Materials

1.0
Dimethyl sulfoxide y = –0.0108x + 1.3778
R2 = 0.8871
0.9

0.8
Cos θ

Ethylene glycol
0.7
Glycerol Water

0.6

0.5
30 35 40 45 50 55 60 65 70 75 80

γc = 34.98 mJ/m2 Surface tension of liquid – mJ/m2

Figure 2.6 Zisman’s plot (https://www.researchgate.net/publication/279532584_Analysis_of_


the_Results_of_Surface_Free_Energy_Measurement_of_Ti6Al4V_by_Different_Methods/
figures?lo=1).

2.13 Zisman’s Rectilinear Relationship—Zisman’s


Plots and Critical Surface Tension of a Solid
Fox and Zisman [33] introduced the well-known concept of critical surface
tension, γc. In the method adopted by Zisman, the contact angles θ for a
series of organic homologous liquids were measured on a solid. A plot of
cos θ vs. surface tension of the liquids gave a straight line (Figure 2.6). The
point of intersection of the straight line and the line through cos θ = 1 is the
“critical surface tension” γc of the solid. All liquids whose surface tension is
less than the critical surface tension will wet the substrate.

2.14 Effect of Surface Roughness on Contact Angle


The above equation is true for smooth, contamination-free surface. However,
the real solid surface is not smooth and the roughness of the surface has a pro-
found effect on the wetting and adhesion. The innumerable small hills, valleys,
and crevices on the solid surface entrap and occlude air or vapor within them.
Even if θ = 0, it is not possible under real conditions to ensure that an intimate
contact between the adhesive and adherend is established. Surface roughness
plays therefore an important role in the wettability of a solid surface.
The impact of roughness on the contact angle is given by the Wenzel
equation (Equation 2.12)

Cos θW = r Cos θY (2.12)


Fundamentals of Adhesion 51

The Wenzel equation (Equation 2.12) relates the contact angle θw of


a liquid measured on a rough surface having a roughness ratio, r, with
the contact angle of the same liquid measured on a smooth surface, θY.
The roughness ratio is the ratio of the true surface area of a rough sur-
face to the surface area of the smooth surface. This ratio r will always be
larger than one. Wenzel’s relation also shows that surface roughness will
decrease the contact angle for a water droplet on a hydrophilic surface or
increase the contact angle for a water droplet on a hydrophobic surface.
The advancing and receding contact angles can throw light on the mag-
nitude of roughness of the wood surface. The difference between advancing
and receding contact angles is the contact angle hysteresis. The magnitude of
contact angle hysteresis is dependent on roughness, topography, morphol-
ogy, and chemical homogeneity of the solid surface [31]. Good [34] sug-
gested that the advancing contact angle represents hydrophobic areas on the
surface, while the receding contact angle characterizes hydrophilic areas.

2.15 Weak Boundary Layer Theory


The weak boundary layer theory explains the loss of adhesion as a failure in
an intermediate molecular layer between adhesive and adherent [35]. This
molecular layer consists of low-molecular-weight impurities of various ori-
gins including water. This theory has never been verified for wood, but it is
known that low-molecular-weight extractives can easily migrate to the sur-
face and might reduce adhesion due to weak boundary layer (see Chapter 9).
Bonding of wood is described as a chain of several links, comprising
wood, wood surface and its boundary layer, interphase of wood and adhesive
and interface between wood and adhesive, and the adhesive bond line itself.
One useful method for understanding the abovementioned links that
control the adhesive strength as well as weakness is the chain link analogy
proposed by Marra [5]. Different areas of the substrate and adhesive are
likened to a series of chain links, with the weakest link being the site of
fracture. This is depicted in Figure 2.7. These links are as follows:
Link 1, adhesive film; links 2 and 3, intra-adhesive boundary layers; links 4
and 5, adhesive–adherend interface (in this region, the weak boundary layer
exist); links 6 and 7, adherend subsurface; links 8 and 9, adherend proper.
Another cause for weak link is due to the stresses caused at the bond
line due to swelling and shrinkage due to moisture changes. If the adhesive
bond has to be durable, it has to adapt itself to the dimensional changes
and the consequent strain due to swelling and shrinkage accompanying the
changes in moisture conditions. Two distinct classes of wood adhesives have
52 Adhesives for Wood and Lignocellulosic Materials

Figure 2.7 Different links in adhesive bonding.

different ways to distribute this strain: (a) rigid in situ polymerized adhe-
sives relieves this stress in many cases by distributing the strain through the
wood interphase region. (b) The other class, the more flexible pre-polymerized
adhesives, can distribute the strain through the adhesive interphase. Failure to
adequately perform either of these strain distribution processes can lead to
high strains and subsequent failure zones. As wood dries, it shrinks back
to near its original dimensions. These failure zones can expand and become
more visible as delamination areas [36].
Mechanically weak wood surfaces can be another source of weak
boundary layer. The causes of this are many [5, 35]. One cause is physical
crushing of the surface, especially by abrasive planing or by too high of
a bonding pressure. This occurs when more pressure is applied than the
thin-walled earlywood cells can withstand. The strength of these cells is
reduced due to deformation and fracture of the cell walls. A second cause
is sanding dust or other dust accumulation on surfaces. A third cause is
tearing of the surface that occurs during planing and sanding. A fourth
cause, associated with high-density wood species, is cells becoming sepa-
rated from one another due to the force of planning [36].

2.16 Measurement of the Wetting Parameters


for Wood Substrate
Unlike other substrates, wood exhibits complex anatomical features that
make it heterogeneous and porous. It is basically hygroscopic. It contains
extractives.
Fundamentals of Adhesion 53

As a result of these unique physicochemical characteristics, wetting


measurements on wood is difficult. Direct measurement of contact angle
of adhesives on wood surface may not be reproducible and hence not sat-
isfactory. For example, a drop of water on a wood surface will, in most
cases, quickly change its size and shape over time, which will thus lead to a
change in the contact angle.
New methods for determining the wetting parameters of wood are nec-
essary. A detailed and critical review of the various methods to determine
the wetting parameters have been reported by Magnus Wålinder. The read-
ers may refer to the same [37]. A brief summary from the above review is
given below.
“One way to address some of the difficulties in wood wetting measure-
ments may be to apply the Wilhelmy (1863) method [38]. In contrast to
direct measurement of contact angles, as in the drop method, the Wilhelmy
method involves determining the force acting on a specimen when it is
immersed in and withdrawn from a liquid. An apparent contact angle can
then be estimated from an analysis of the recorded force”.
“Other promising techniques for estimating the surface energetics of
wood may be the Axisymmetric Drop Shape Analysis-contact diameter
(ADSA-CD) technique. Contact angle measurements determined as con-
stant wetting rate angle values (cwra) and also a capillary rise technique
(column wicking) applied to wood”.
“Inverse gas chromatography (IGC) is a useful technique for determin-
ing surface energetics of particle surfaces. By using appropriate gas probes,
IGC can provide information on the surface thermodynamic character-
istics of particles including surface free energy, acid-base interactions,
enthalpy, and entropy. IGC has been applied to many materials such as
polymers, wood pulp and wood particles”.
Some spectroscopic methods, namely, X-ray photoelectron spectros-
copy (XPS) and FT Raman, have also been recommended by the author.
Both IGC and wicking methods rely on wood powder, which will give
different results compared to measurements on solid wood.

2.16.1 Some Results on Surface Energy of Wood


As discussed in Chapter 1, wood is a hierarchical cellular material and is
therefore anisotropic in nature. Further wood surfaces are topographically
different in radial, tangential, and transverse sections [31]. The wood sur-
face consists of earlywood and latewood. The water contact angle of early-
wood is often different from that of latewood. At the microscale, the wood
surface consists primarily of lumen surfaces and cross-sectional walls.
54 Adhesives for Wood and Lignocellulosic Materials

Thus, wetting of wood by the adhesives is therefore complex. Freeman was


the first to report on the wettability of wood [39].
Gray carried out an extensive investigation on the wettability of 20
species of wood [7]. Gray was first to determine the surface free energy
of wood. Sessile drop method was used to determine the contact angle,
and the critical surface tension (γc) was obtained by the Zisman method.
Species used by Gray were Western Hemlock (Tsuga heterophylla),
Douglas-fir (Pseudotsuga taxifolia), Afrormosia (Afrormosia elata),
Parana pine (Araucaria angustifolia), European redwood (Pinus sylves-
tris), English oak (Quercus robur), and Beech (Fagus sylvatica). The val-
ues of critical surface tensions ranged from 34.5 to 81.0 mJ/m2. One of
the important findings was that freshly sanded surfaces were approxi-
mately 20 mJ/m2 higher in surface energy than un-sanded, aged surfaces.
It was also shown that surface contamination occurs rapidly on freshly
cleaned surfaces.
Herczeg reported on the wettability of Douglas-fir wood [40]. The crit-
ical surface tensions, γc, were found to be between 44 and 50 dynes/cm for
summerwood and springwood, respectively. The surface-free energy and
the maximum work of adhesion were also reported. It was also reported
that aging increased contact angle, showing that wood wettability was
reduced. Chen reported that removal of extractives from some tropical
woods improved wettability [41]. Hse measured the wettability of south-
ern pine veneers by measuring the contact angles formed with 36 phenol-
formaldehyde resins [42]. The contact angle of resins on earlywood was
less than that on latewood, apparently because earlywood surfaces were
rougher. Also, the contact angle was positively correlated with the glue bond
quality as tested by wet shear strength, percent of wood failure, and per-
cent of delamination. Nguyen and Johns found that the surface free energy
of Douglas-fir and redwood decreased with aging time [43]. Extractives
and aging had significant influence on the surface energy. The surface free
energy of Douglas-fir was 48.0 mJ/m2, and after extraction, it increased
to 58.9 mJ/m2. These results emphasize the influence of extractives on the
wood surface energy.
Kalnins et al. [44] employed a video-type technique to measure the
dynamic contact angle of distilled water as test liquid on wood with mea-
surements conducted at the elapsed time of 3 to 5 s.
Gardner et al. found dynamic contact angle measurements to be useful
for determining the effect of wood processing and environmental condi-
tioning on surface energetic [45].
Kajita and Skaar evaluated the wettabilities of the surfaces of some
American softwoods species (using cosine 0 as the index of wettability) [46].
Fundamentals of Adhesion 55

Also, the earlywood wet more easily than did the latewood (earlywood has
a greater roughness factor and a greater porosity). The wetting angles var-
ied from 68° (eastern red cedar) to 14° (Alaska cedar). The greater wettabil-
ity of sapwood compared with heartwood was attributed to the extractive
content of the heartwood.
Cosine of the advancing contact angle was employed as the measure of
wettability. The wettability, pH, and specific gravity were closely related to
glue-bond quality of resorcinol-phenolic and urea formaldehyde-bonded
adhesive joints [47].
Shi and Gardner developed a wetting model to describe to quantify the
adhesive penetration and spreading during the adhesive wetting process
[8]. Sapwood and heartwood of southern pine and Douglas-fir were stud-
ied. Two resin systems, polymeric diphenylmethane diisocyanate (PMDI)
and phenol formaldehyde (PF), were evaluated. It was learned from this
study that the wetting model could accurately describe the dynamic adhe-
sive wetting process on wood surfaces. Shen et al. presented a systematic
study of surface free energy and acid–base properties of pine (P. sylvestris
L.); for evaluation of the data, the Lifshitz–van der Waals/acid–base (LW–
AB) approach was applied [48].
Nussbaum observed a decrease of wettability as a function of time
on wood surface after sawing due to the migration of wood extractives
to the surface [49]. Gindl et al. compared the applicability of different
approaches to determine the surface free energy of wood and found the
LW–AB approach to be the most effective among the generally accepted
models [50].
de Meijer et al. [51] employed contact angle measurements to calculate
the Lifshitz−van der Waals, acid−base, and total surface free energies of
wood species spruce (Picea abies) and meranti (Shorea spp.). These species
were characterized by low surface energy with a dominant Lifshitz−van
der Waals component. The authors report that thermodynamic equilib-
rium conditions as assumed by Young’s equation are generally not fulfilled
with wood surfaces because of chemical heterogeneity, surface roughness,
and the adsorption of the test solvent.
An exhaustive review of wettability of wood has been published by
Piao et al. [31]. The review also includes calculation of surface tension of
wood, Zisman’s critical surface tension, Owens–Wendt’s geometric mean,
and Wu’s harmonic mean; Young–Fowkes–van Oss–Good acid–base
approaches and the inverse gas chromatography method have been dis-
cussed in detail. The review also deals with variables that affect the wet-
tability and surface energy of wood. Detailed overview of literature data
obtained on wood surfaces was presented by de Meijer et al. [51].
56 Adhesives for Wood and Lignocellulosic Materials

2.17 Covalent Bond Formation


Covalent bonds between an adhesive and wood are believed to improve
bond durability. Although covalent bonds—chemical bonds between the
adhesive and wood—seem plausible with some adhesives, they have never
been unambiguously detected in an adhesive bondline and no evidence
exists that they contribute to the strength of adhesive bonds To determine
whether an adhesive forms covalent bonds to wood, it must have the fol-
lowing characteristics: (1) be highly reactive with wood polymer hydrox-
yls, (2) be capable of permeating the cell wall, (3) exhibit strong wettability
to wood, and (4) be amenable to study using a monofunctional model
compound. Ideally, the reaction between the monofunctional model com-
pound and wood will produce distinct chemical shift differences between
unreacted and reacted wood polymers in solution-state nuclear magnetic
resonance (NMR) spectroscopy.
While there is no doubt that adhesive-to-wood covalent bonds can
form under specific experimental conditions, the conditions employed in
studies reporting such bond formation generally have not corresponded
to the conditions commonly used in the bonding of wood, in particular
hot-pressing of wood panels [52]. For instance, covalent bonds between
wood and a synthetic adhesive can form at temperatures higher than
120°C maintained for 2 h [52]. But extensive covalent bonding appears
unlikely in the core of a particleboard, which is able to reach only 115°
to 120°C for no more than 1 to 1.5 min when pressed at 200°C for 3 min.
Allan and Neogi found in the case of phenol-formaldehyde bonding of
wood at 120°C for 2 h that only one adhesive-to-substrate covalent bond
was formed for approximately every 1200 cross-links within the resin itself
[53]. This was also the case for isocyanate binders, for which the original
misconception of a predominance of covalent bonding between adhesive
and substrate that has been used to explain the high strength of the panels
obtained in that manner was disproved [54, 55].

References
1. Frihart, C.R., Wood adhesion and adhesives, in: Handbook of Wood Chemistry
and Wood Composites, 2nd edn., R. Rowell (Ed.), pp. 255–319, CRC Press,
Boca Raton, Florida, 2013.
2. Israelachvili, J.N., Intermolecular and Surface Forces, 2nd edn., Academic
Press, London, 1991.
Fundamentals of Adhesion 57

3. van Oss, C.J., Interfacial Forces in Aqueous Media, Marcel Dekker, New York,
1994.
4. Arif Butt, M., Arshad Chautai, A., Ahmad, J.A.R., Theory of adhesion and its
practical implications. J. Faculty Eng. Technol., 21–45, 2007.
5. Marra, G.G., Technology of Wood Bonding: Principles in Practice, Van
Nostrand, New York, 1992.
6. Browne, F.L. and Brouse, D., Nature of adhesion between glue and wood. Ind.
Eng. Chem., 21, 80–84, 1929.
7. Gray, V.R., The wettability of wood. For. Prod. J., 12, 452–461, 1962.
8. Shi, S.Q. and Gardner, D.J., Dynamic adhesive wettability of wood. Wood
Fiber Sci., 33, 56–68, 2001.
9. Marcinko, J.J., Phanopoulos, C., Beachey, P., Wood Adhesives 2000, Proceedings
No. 7252, pp. 111–121, Forest Products Society, Madison, Wisconsin, 2001.
10. Christiansen, A.W., How overdrying wood reduces its bonding to phenol–
formaldehyde adhesives: A critical review of the literature. Part I. Physical
responses. Wood Fiber Sci., 22, 4, 441–459, 1990.
11. Christiansen, A.W., How overdrying wood reduces its bonding to phenol–
formaldehyde adhesives: A critical review of the literature. Part II. Chemical
reactions. Wood Fiber Sci., 23, 1, 69–84, 1991.
12. Stehr, M., Adhesion to machined and laser ablated wood surfaces, Dissertation,
KTH Stockholm, Sweden, 1999.
13. McBain, J.W. and Hopkins, D.G., On adhesives and adhesive action. J. Phys.
Chem., 29, 2, 188–204, 1925.
14. Deryaguin, B.V., Problems of Adhesion. Vestnik Akademie, 8, 70, 1955.
15. Deryaguin, B.V., Krotova, N.A., Karassev, V.V., Kirillova, Y.M., Aleinikova,
I.N., Proceedings of the Second International Congress on Surface Activity—
III, Butterworths, London, 1957.
16. Deryaguin, B. and Smiliga, V.P., Adhesion Fundamentals and Practice,
MacLaren and Sons, London, 1969.
17. Voyutskii, S.S., Adhesives Age, 5, 4, 30, 1962; Voyutskii, S.S., Vakula, V.L.,
The role of diffusion phenomena in polymer-to-polymer adhesion. J. Appl.
Polym. Sci., 7, 2, 475–491, 1963.
18. Voyutskii, S.S., Autohesion and Adhesion of High Polymers, Wiley Interscience,
New York, 1963.
19. Voyutskii, S.S., Markin, S., Yu, I., Gorchakova, v.M., Gul, V.E., Adhesion of
polymers to metals. Adhesives Age, 8, 24, 1965.
20. Pizzi, A., A brief, non-mathematical review of adhesion theories as regards
their applicability to wood. Holzforsch Holzververt, 44, 1, 6–11, 1992.
21. Gardner, D.J., Adhesion mechanisms of durable wood adhesive bonds.
BLIO009-Stokke September 13, 21, 254–266, 2005.
22. Dupré, A. and Dupré, P., Théorie mécanique de la chaleur, Gauthier-Villars,
Paris, 1869.
23. Fowkes, F.M., Attractive forces at interfaces. Ind. Eng. Chem., 56, 12, 40–52, 1964.
58 Adhesives for Wood and Lignocellulosic Materials

24. Fowkes, F.M., Role of acid–base interfacial bonding in adhesion. J. Adhes.


Sci. Technol., 1, 7–27, 1987.
25. Fowkes, F.M., Calculation of work of adhesion by pair potential summation.
J. Colloid Interface Sci., 28, 493–505, 1968.
26. Owens, D.K. and Wendt, R.C., Estimation of the surface free energy of
polymers. J. Appl. Polym. Sci., 13, 1741–1747, 1969.
27. Wu, S., Calculation of interfacial tension in polymer system. J. Polym. Sci. C,
34, 19–30, 1971.
28. Wu, S., Polar and nonpolar interactions in adhesion. J. Adhes., 5, 39–55,
1973.
29. van Oss, C.J., Good, R.J., Chaudhury, M.K., The role of van der Waals forces
and hydrogen bonds in “hydrophobic interactions” between biopolymers
and low energy surfaces. J. Colloid Interface Sci., 111, 378–390, 1986.
30. van Oss, C.J., Chaudhury, M.K., Good, R.J., Interfacial Lifshitz–van der
Waals and polar interactions in macroscopic systems. Chem. Rev., 88, 927–
940, 1988.
31. Piao, C., Winandy, J.E., Shupe, T.F., From hydrophilicity to hydrophobicity:
A critical review: Part I. Wettability and surface behaviour. Wood Fiber Sci.,
42, 4, 490–510, 2010.
32. Chehimi, M.M., Azioune, A., Cabet-Deliry, E., Acid–base interactions:
Relevance to adhesion and adhesive bonding, in: Handbook of Adhesive
Technology, 2nd edn., A. Pizzi and K.L. Mittal (Eds.), Marcel Dekker,
New York, 2003.
33. Fox, H.W. and Zisman, W.A., The spreading of liquids on low energy sur-
faces. J. Colloid Sci., 5, 6, 499–595, 1950.
34. Good, R.J., Contact angles and the surface free energy of solids. Page 1, in:
Surface and Colloid Science, vol. VII, R.J. Good and R.R. Stromberg (Eds.),
Plenum Press, New York, 1979.
35. Stehr, M. and Johansson, I., Weak boundary layers on wood surfaces.
J. Adhes. Sci. Technol., 14, 1211–1224, 2000.
36. Frihart, C.R. and Hunt, C.G., Adhesives with wood materials: Bond forma-
tion and performance, in: Wood Handbook: Wood as an Engineering Material,
Centennial edn., General Technical Report FPL–GTR–190. pp. 10.1–10.24,
U.S. Dept. of Agriculture, Forest Service, Forest Products Laboratory, USDA,
2010.
37. Walinder, M., Wetting Phenomena on Wood: Factors Influencing Measurements
of Wood Wettability, Dissertation, KTH Royal Institute of Technology,
Stockholm, 2000.
38. Wilhelmy, L., Ueber die Abhängigkeit der Capillaritäts-Constanten des
Alkohols von Substanz und Gestalt des benetzten festen Körpers. Ann.
Physique Chimie, Band CXIX, 5, 6, 12, 1863.
39. Freeman, H., Properties of wood and adhesion. For. Prod. J., 9, 451–458,
1959.
40. Herczeg, A., Wettability of wood. For. Prod. J., 15, 499–505, 1965.
Fundamentals of Adhesion 59

41. Chen, C., Effect of extractive removal on adhesion and wettability of some
tropical woods. For. Prod. J., 20, 1, 36–40, 1970.
42. Hse, C.-Y., Wettability of southern pine veneer by phenol–formaldehyde
wood adhesives. For. Prod. J., 22, 1, 51–56, 1972.
43. Nguyen, T. and Johns, W.E., The effects of aging and extraction on the surface
free energy of Douglas fir and redwood. Wood Sci. Technol., 13, 1, 29–40,
1979.
44. Kalnins, M.A., Katzenberger, C., Schmieding, S.A., Brooks, J.K., Contact
angle measurement on wood using videotape technique. J. Colloid Interface
Sci., 125, 344–346, 1988.
45. Gardner, D.J., Generalla, N.C., Gunnels, D.W., Wolcott, M.P., Dynamic
wettability of wood. Langmuir, 7, 2498–2502, 1991.
46. Kajita, H. and Skaar, C., Wettability of the surfaces of some American soft-
woods species. Mokuzai Gakk., 38, 516–521, 1992.
47. Mantanis, G.I. and Young, R.A., Wetting of wood, Wood Sci. Technol., 31,
339–353, 1997.
48. Shen, Q., Nylund, J., Rosenholm, J.B., Estimation of the surface energy and
acid–base properties of wood by means of wetting method. Holzforschung,
52, 521–529, 1998.
49. Nussbaum, R.M., Natural surface inactivation of Scots pine and Norway
spruce evaluated by contact angle measurements. Holz Roh-Werkst., 57,
419–424, 1999.
50. Gindl, M., Sinn, G., Gindl, W., Reiterer, A., Tschegg, S., A comparison of
different methods to calculate the surface free energy of wood using contact
angle measurements. Colloids Surf. A Physicochem. Eng. Asp., 181, 279–287,
2001.
51. de Meijer, M., Haemers, S., Cobben, W., Militz, H., Surface energy determi-
nations of wood: Comparison of methods and wood species. Langmuir, 16,
9352–9359, 2000.
52. Hubbe, M.A., Pizzi, A., Zhang, H., Halis, R., Critical links governing per-
formance of self-binding and natural binders for hot-pressed reconstituted
lignocellulosic board products: A review. Bioresources, 13, 1, 1–67, 2018.
53. Allan, G.C. and Neogi, A.N., Fiber surface modification, Part VIII: The
mechanism of adhesion of phenol–formaldehyde resins to cellulosic and lig-
nocellulosic substrates. J. Adhes., 3, 1, 13–18, 1971.
54. Pizzi, A. and Owens, N.A., Interface covalent bonding vs. wood-induced
catalytic autocondensation of diisocyanate wood adhesives. Holzforschung,
49, 269–272, 1995.
55. Wandler, S.L. and Frazier, C.E., The effects of cure temperature and time
on the isocyanate–wood adhesive bondline by 15N CP/MAS NMR. Int. J.
Adhes. Adhes., 16, 3, 179–186, 1996.
3
Urea–Formaldehyde Resins

3.1 Introduction
Urea–formaldehyde (UF) adhesives, a member of the so-called aminoplas-
tics, are the product of the reaction between urea and formaldehyde. They
are thermosetting resins. Currently, in excess of an estimated 6 million
tons of UF resins are produced yearly worldwide, based on a typical solids
content of 66% by mass [1, 2].
UF resins account for about 80% of the amino resins produced world-
wide, with the remaining 20% being almost melamine–formaldehyde
resins except for minor amounts of resins that are produced from other
aldehydes or amino compounds (especially aniline), or both.
Although the raw materials are few and simple chemicals, their interac-
tion results in the production of a multitude of linear and branched oligo-
meric species of varied complexities depending on the reaction conditions.
UF resins are relatively inexpensive. They are colorless and hence will not
impart objectionable discoloration to the light-colored decorative wood
and veneers. UF resins can also be easily handled. UF resins will not cure
by itself, although it will increase in viscosity on aging precipitating out of
solution in extreme cases. They require a hardener to cross-link and trans-
form the liquid resin into a thermoset with a hardened three-dimensional
network structure.
Relatively low pressing temperatures and shorter pressing times are
needed for bonding wood composite panels such as plywood, particle-
board, and medium-density fiberboard (MDF).
UF resins do however suffer from the drawback of presenting a much
lower water resistance than phenolic and melamine formaldehyde resins.
Because of this defect, products made by using urea resin adhesives such
as plywoods cannot withstand either outdoor conditions or being used in
high-humidity environments. Furthermore, UF-bonded products emit
formaldehyde due to the low stability of their type of amino-methylene
bonds.

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (61–100) © 2019
Scrivener Publishing LLC

61
62 Adhesives for Wood and Lignocellulosic Materials

3.2 Historical Review of UF Resins (Plastic Historical


Society) [3]
Hans John disclosed for the first time uncatalyzed condensation prod-
ucts of UF in his patent in 1920. This was followed by a number of patents
between 1920 and 1924 by F. Pollak and his co-workers in Vienna who used
a variety of catalysts. One of the products of their patent was a glassy trans-
parent resin that was given the name Pollopas, an effort directed towards
the production of the so-called “organic glasses” [3].
Goldschmidt and Neuss in Germany also worked on UF, but the first com-
mercially successful thermosetting molding material was produced by the
British Cyanides Co.—based on a mixture of thiourea and UF in 1928 [3].
I. G. Farben developed UF especially as adhesives and stoving lacquers
with patents in 1925 and 1928. In 1933, the various firms making UF mate-
rials agreed to exchange patent rights in order to avoid possible disputes
and rapid development followed [3].
Although UF resins are products of only two main reactants, i.e., urea
and formaldehyde, they present a broad variety of possible consecutive and
parallel reactions and structures.
The reaction between urea and formaldehyde can be catalyzed by alkali
or acid and is quite complex. When acid catalyst is used, the polymer so
obtained tends to have longer linear chains with side branches and present-
ing good adhesive properties. When alkali-catalyzed, the polymer is less
suitable as an adhesive.
Although the primary reaction products of urea and formaldehyde were
investigated as early as 1884, intensive study of UF resins did not begin until
the early 1920s [3]. The first polymer of this type to achieve commercial
importance, around 1930, was a urea–thiourea–formaldehyde resin [4].
The reactions between urea and formaldehyde have since been exten-
sively studied and documented [5–12].

3.3 Reaction between Urea and Formaldehyde


The reaction between urea and formaldehyde has proved to be compli-
cated. Excellent and extensive reviews dealing with it already exist. The
combination of these two chemical compound results in both linear and
branched polymers, as well as three-dimensional network, in the cured
resin. This is due to a functionality of 4 in urea (due to the presence of four
replaceable hydrogen atoms) (in reality, only trifunctional as the reactivity
decreases in proportion 9:3:1 as each site is substituted) and a functionality
Urea–Formaldehyde Resins 63

of 2 in formaldehyde [13]. The most important factors determining the


properties of the reaction products are as follows [13]:

(1) The relative molar proportion of urea and formaldehyde


(2) The reaction temperature
(3) The various pH values at which condensation takes place

The above factors influence the rate of increase of the molecular weight
of the resin with time. Therefore, the characteristics of the reaction prod-
ucts, especially the solubility, viscosity, water tolerance, and rate of curing
of the adhesives differ considerably based on whether the product is at
lower or higher stages of condensation, i.e., thus depend to a large extent
on molecular weights [4].

3.4 Reaction Sequence


The reaction between urea and formaldehyde can be divided into two
stages [13, 14]:

Stage 1. Hydroxymethylation
Akaline condensation to form mono-, di-, and trimethylol
ureas. Tetramethylol urea has never been isolated.
Stage 2. Polycondensation (chain extension). The second stage
is the acid condensation stage of methylol ureas, first to sol-
uble and then insoluble cross-linked resins.

Further details on this are published by Christjanson et al. [15].


The first reaction between formaldehyde and urea under alkaline con-
ditions is the formation of monomeric methylol ureas (hydroxymethylated
ureas), a complex mixture consisting of mono-, di-, and tri-substituted ureas.
The kinetics of the formation and condensation of mono- and dimethylol
ureas have been studied extensively. The formation of monomethylol urea
in weak acid alkaline aqueous solution is characterized by an initial fast
phase followed by a slow bimolecular reaction [11, 16–18]. The reactions
are reversible, with an equilibrium being established between reactants and
products. The rate of reaction varies according to the pH, with minimum
rate of reaction in the pH range 5 to 8 for a urea:formaldehyde at a molar
ratio of 1:1 [13] and at the pH of about 6.5 for a 1:2 molar ratio [9]. The
overall reaction is characterized by a rapid initial reaction followed by a slow
condensation reaction resulting in the formation of the prepolymer [13].
64 Adhesives for Wood and Lignocellulosic Materials

The most common method of manufacture of UF resin adhesives


employs an addition of a second amount of urea during the reaction [13].
This concept is employed to reduce the toxicity of UF resins by adjusting
them to a low final molar ratio of F/U of 1.05. It means that this second urea
is added after the alkalinization of the polycondensate to stop the conden-
sation reaction to proceed further. Because of the high content of unreacted
urea and reactive hydroxymethyl groups in the product, numerous reac-
tions take place during the storage of the resins [19]. It means that resins
having been aged for different times can have different technical properties.
Kim and co-workers [20] reported from 13C-NMR studies that migra-
tion of significant amount of hydroxymethyl groups occurs in the resin
system when the second urea is added as described in the usual resin
preparation procedure.

3.5 Manufacture of UF Resin


In contrast to the manufacture of phenolic resins for wood panel products,
which essentially involves condensation under single pH (alkaline) conditions
to form resoles, UF resins are condensed by changing the pH of the reaction
medium in a consecutive manner according to well-established protocol.
Two main strategies for synthesizing UF resins are as follows:

A. The first process consists of three steps:


(1) An alkaline methylolation
(2) Acid condensation (chain extension)
(3) Neutralization and then the addition of a final amount of urea

B. The second process consists of four steps:


(1) initial condensation under a strongly acid environment
(2) alkaline condensation
(3) acid condensation, and
(4) neutralization.

These formulations have long been used commercially with great suc-
cess for UF resins in an F/U ratio between 1.45 and 1.65 [21].
A number of “preparation diagrams” depicting details of operating con-
ditions such as pH, temperature, formaldehyde/urea ratio, and reaction
time for the manufacture of UF resins have been published [13, 22]. These
reaction protocols are very useful and can be used as guidelines in the
industry for the production of UF resins. Similar diagrams have been made
Urea–Formaldehyde Resins 65

available for the production of melamine–formaldehyde and melamine–


urea–formaldehyde (MUF) resins [13].

3.6 Chemistry of Reaction—Conventional Process


(Alkaline–Acid Process/Three-Step Process)
As described in the previous section, the conventional process for man-
ufacturing UF resin consists of three stages: (a) reaction in the alkaline
medium, (b) reaction under acidic conditions, and (c) neutralization and
addition of second urea. The chemical reactions occurring under the above
conditions are outlined below.

3.6.1 First Stage—Reaction under Alkaline Conditions


In the first stage, the amino groups of urea react with formaldehyde to form
hydroxymethyl ureas (Figure 3.1). This reaction is a consecutive process lead-
ing to the formation of mono-, di-, and trimethylolureas. Tetramethylolurea
is apparently not produced, at least not in a detectable quantity [14].
Methylolation reactions are reversible consecutive reactions as shown
below:

NH2 NHCH2OH NHCH2OH


CH2O CH2O
C O C O C O

NH2 NH2 NHCH2OH


UREA MONOMETHYLOL DIMETHYLOL UREA
UREA

NHCH2OH NHCH2OH
CH2O
C O C O

NHCH2OH N-(CH2OH)2

DIMETHYLOL UREA TRIMETHYLOL UREA

N-(CH2OH)2

C O

N-(CH2OH)2
Tetramethylol urea

Figure 3.1 Reaction between urea and formaldehyde: mononuclear methylol ureas.
66 Adhesives for Wood and Lignocellulosic Materials

Tetramethylolurea is apparently not produced, at least not in a detectable


quantity. Each methylolation step is characterized by its own rate constants
ki for both the forward and backward reactions. The reverse reaction can be
considered as hydrolysis. The reversibility of the reactions is an important
feature of the system and is of great practical importance since it is respon-
sible for the hydrolytic instability of the cured UF resins and the consequent
release of formaldehyde when the UF-bonded wood products are exposed to
high humidity or water. Side products are acetals, hemiacetals, and etherified
products with residual methanol, which is always present in small amounts
from the production of formaldehyde [23]. Monomethylol, dimethylol, and
trimethylol ureas are formed in the ratio 9:3:1, respectively [24, 25].
The initial addition of formaldehyde to urea is reversible with the for-
ward reaction about 100 times faster than the reverse reaction [14, 26, 27]
and is subject to general acid–base catalysis. Secondary products contain-
ing methylene–ether bonds are also reported to be formed, presumably the
methylene–ethers of hydroxymethylureas.
The forward bimolecular reaction has an energy of activation of
13 kcal mol−1. The reverse unimolecular reaction has an energy of acti-
vation of 19 kcal mol−1.

3.6.1.1 Reaction Mechanism [14]


The mechanism of the base catalyzed reaction is given below:
See in the reactions from Equations 3.1 to 3.4 that the OH– (BASE)
enters the system in Equation 3.1 and leaves in Equation 3.4 after com-
pletion of its role without undergoing any change. Thus, the formation of
monomethylol urea is a base-catalyzed reaction. Similarly does occur the
formation of di- and trimethylol ureas. Methylene ethers bridges (-CH2-
O-CH2-) are also formed in this reaction by elimination of water between
the methylol groups of two methylol ureas. These are also a major cause of
formaldehyde emission in service as they tend to rearrange to methylene
bridges with emission of formaldehyde [13, 24].

3.6.2 Second-Stage Condensation Reaction under Acid


Conditions: Chain Extension
The second stage of UF resin synthesis involves chain extension reaction in
which methylolureas condense to form low-molecular-weight polymers. The
rate at which these condensation reactions occur is dependent on the pH
(Figure 3.2) and, for all practical purposes, occurs only at pHs lower than 8 and
accelerates progressively as the pH progresses to more acidic pH values [24].
Urea–Formaldehyde Resins 67

NH2CONH2 + OH– NH2–CO–NH–1 + H2O (3.1)


H

C=O H2C+ O– (3.2)

NH2–CO–NH–1 + H2C+ O– NH2–CO–NH–CH2–O– (3.3)

NH2–CO–NH–CH2–O– + H2O NH2–CO–NH–CH2–OH + OH– (3.4)

Figure 3.2 Reaction mechanism.

3.6.2.1 Reaction Chemistry


Chain extension occurs as a result of the following alternate polycon-
densation steps:

3.6.2.2 Reaction Mechanism


The effect of pH on the rate of addition of formaldehyde to urea is shown
in Figure 3.3. The rate of successive addition of formaldehyde to urea to
form mono-, di-, and trimethylol derivatives has been estimated to be in
the ratio of 9:3:1, respectively [13, 24, 26].
The chain extension reactions taking place under acidic conditions
result in an increase in the molecular weight of the UF as a result of for-
mation of either methylene bridges or methylene ether bridges between
individual units described below (see Figures 3.4 and 3.5):

1. The reaction of methylol groups of one molecule (UNIT 1)


with the amido H of another molecule (UNIT 2) to form a
methylene link between them (Equation 3.1).
2. Reaction between two methylol groups (of UNIT 1 and
UNIT 2) to form a methylene ether link (Equation 3.2).
3. Methylene ethers can split to release formaldehyde and
the formation of methylene link between the two units
(Equation 3.3).
4. Two methylol groups, one from each molecule (i.e., between
UNIT1 and UNIT 2), can react to produce methylene links,
formaldehyde, and water (Equation 3.4).
68 Adhesives for Wood and Lignocellulosic Materials

–7

–8

–9
Addition Reaction
–10
In k

–11

–12
Condensation Reaction
–13

–14
1 2 3 4 5 6 7 8 9 10 11 12
pH

Figure 3.3 Effect of pH on the rate of addition and condensation reactions [24].

1. Reaction between the methylol group of UNIT 1 with the amidohydrgen of UNIT 2 to produce
methylene bridge Equation 3.1
O O O O

NH2-C-NHCH2OH + NH2-C-NHCH2OH NH2-C-NH-CH2-NH-C-NHCH2OH


UNIT 1 UNIT 2

2. Reaction of methylol group of UNIT 1 with the methylol group of UNIT 2 to produce
an ether linkage Equation 3.2
O O O O

NH2-C-NHCH2OH + HO-CH2NH-C-NHCH2OH NH2-C-NHCH2-O-CH2NH-C-NHCH2OH


UNIT 1 UNIT 2

3. Ether linkage decomposes to yield methylene bridge and formaldehyde Equation 3.3
O O O O

NH2-C-NHCH2-O-CH2NH-C-NHCH2OH NH2-C-NH-CH2-NH-C-NHCH2OH + CH2O

4. Reaction between methylol group of UNIT 1 with methylol groups of UNIT 2 to produce
methylene bridge and formaldehyde Equation 3.4
O O O O

NH2-C-NHCH2OH + HO-CH2NH-C-NHCH2OH NH2-C-NH-CH2-NH-C-NHCH2OH


UNIT 1 UNIT 2 + CH2O

Figure 3.4 Reaction chemistry.


Urea–Formaldehyde Resins 69

NHCH2OH NHCH2OH2+ NHCH2+

C=O + H+ C=O C=O + H 2O

NHCH2OH NHCH2OH NHCH2OH

+
NHCH2+ NH2 NH CH2 NH2

C=O + C=O C=O C=O

NHCH2OH NHCH2OH NHCH2OH NHCH2OH

+
NH CH2 NH3 NH CH2 NH

C=O C=O C=O C=O + H+

NHCH2OH NHCH2OH NHCH2OH NHCH2OH

Figure 3.5 Reaction mechanism for chain extension.

According to Equations 3.1 to 3.4, diurea compounds containing meth-


ylene and dimethylene ether bridges are formed. Kumlin and Simonson
[28] carried out investigations on the formation of the diurea as well as
more highly condensed compounds formed during the preparation of UF
resin under different process conditions. They employed liquid chroma-
tography for the quantitative determination of each individual diurea com-
pound. Identification of diurea compounds with methylene ether bridges
was made by 1H NMR spectroscopy.
The ratio of methylene ether bridges to methylene bridges is important
with respect to formaldehyde emission. This ratio depends on the condi-
tions employed for the synthesis of UF. Formaldehyde emissions and the
durability of the cured resin can be controlled by controlling this ratio.

3.6.3 Third Stage—Neutralization and Addition of Second Urea


The third stage consists of neutralization, cooling, and addition of the
second urea. The addition of the second urea is necessary to have a good
storage life.

3.6.3.1 Reactions Involving Migration of Hydroxymethyl Groups


Kim [29–31] and Kim et al. [20, 32], in a series of publications, reported
from the 13C-NMR studies that migration of a significant amount of
70 Adhesives for Wood and Lignocellulosic Materials

hydroxymethyl groups occurs in the resin system when a second urea is


added. The hydroxymethyl groups of the UF resin migrate to the “second
urea” to form monomeric hydroxymethyl ureas. In general, the migra-
tion of the hydroxymethyl groups decreases the viscosity of the resin. The
migration is relatively slow at room temperature and becomes relatively
fast above 50°C.

3.7 Composition of the Commercial UF Resins


1
H- and 13C-NMR spectroscopy have proved to be valuable tools to deter-
mine the identity of the various components present in commercial UF
resin. Several researchers have worked on the NMR of the UF prepoly-
mers, and based on their studies on model compounds, they have been able
to assign values of chemical shifts corresponding to individual chemical
functional groups and the interlinking units present in the UF oligomers.
The commercial resins essentially consist of unreacted urea; formaldehyde;
mono-, di-, and trifunctional monomers; and oligomers of varying chain
lengths. Kim [29–31] and Kim et al. [20, 32] have identified the compo-
nents of commercial UF resins. These components, along with the assigned
13
C NMR chemical shifts, are given below in Figures 3.6 and 3.7.

3.7.1 Monomeric Species


The monomeric species formed by the initial reaction of formaldehyde on
urea are shown in Figure 3.6.

3.7.2 Oligomeric Species


A number of different oligomeric species are formed during the polycon-
densation phase under acid conditions of the reaction of urea with formal-
dehyde. These are shown in Figure 3.7.

3.7.3 General Structure of Commercial UF Resins


General simplified structure of the UF resin for bonding wood panels can
be represented as follows [14]:

HOCH2 NH CO NH CH2 NH CO NH CH2OH (3.5)


n
Urea–Formaldehyde Resins 71

HOCH2-O-CH2OH
(a) & (b)
(a) Oxymethylene (91 ppm) HO-CH2-OH
(b) Oxyhydroxymethylene (87 ppm) (c) Dihydroxymethane (83 ppm)
O
H A
C
N N
H H
(d) Monosubstituted urea
162 (ppm)
O
O
H A C A
C H(A)
N N N N
A A A A
(e) Di- and trisubsituted urea (160.7 ppm) (f ) Tetrasubstituted urea (158 ppm)
O

C H
N N
CH2-OH
(g) Type I Hydroxymethyl (65.2 ppm)
O O
C CH2OH A
N C
N N N
CH2-OH CH2-OH
(h) Type IIi(x) and Type IIi(y) Hydroxymethyl (72 ppm)
O
H H
C
N N
H H
(i) Urea (164 ppm)

Figure 3.6 Monomeric species.

O O
Y'' C
C Y'
N N N
N
CH2
(j) type I methylene (47.1 ppm) Y' = Y'' = H
(k) type II methylene (53.9 ppm) Y' = H; Y'' = A
(l) type III methylene (60.1 ppm) Y' = Y'' = A
O O
Y''
C Y' C
N N N N
CH2-O-CH2
(m) type I methyleneether 69.5 ppm Y' = Y'' = H
(n) type II methyleneether 75.1 ppm Y' = H; Y'' = A
(o) type III methyleneether 79.1 ppm Y' = Y'' = A A = CH2 N ; CH2OCH2 N ; CH2OH

Figure 3.7 Oligomeric species.


72 Adhesives for Wood and Lignocellulosic Materials

O O

C C Y
Y
N X N N
N
Y
Y
Z Z
x

X = CH2, CH2-O-CH2
Y = H, CH2OH, or CH2O-CH2OH
Z = Y, X-NZ(CONZ)y Y
x = degree of polymerization of main polymer chain
y = degree of polymerization of branch polymer chain (3.6)
Figure 3.8 General chemical structure of commercial UF resin.

The above structure however is an oversimplified version of a com-


mercial UF resin. More precisely, the chemical structure of UF resins can
be specified as poly(methylene methylene ether hydroxymethylureas).
Commercial UF resins can be considered as a mixture of different linear
and branched macromolecules characterized by the average molar mass
(or molar mass between branches) and the content of different functional
groups (Figure 3.8). This definition of UF resins leads to a more refined
and statistical structural representation (Equation 3.6) as reported by Kim
[29–31] and Kim et al. [20, 32] based on 13C-NMR studies.
The commercial resin represented by the prepolymer structures of
Equations 3.5 and 3.6 usually has a free formaldehyde content in the range
0.5–1.8%.

3.7.4 Urons
The reaction between urea and formaldehyde also produces cyclic deriv-
atives: uron, monomethyloluron, and dimethyloluron [26].

C
HN NH

H 2C CH2
O

The ability of uron to reduce the emission of formaldehyde has stimu-


lated interest in the compound recently. Gu et al. have reported the syn-
thesis of urons [33]. A high acidic environment was employed by them
Urea–Formaldehyde Resins 73

for the synthesis. Soulard et al. proposed the mechanism of formation of


urons produced in the acid step method [34]. They studied on the stability,
the effect of pH on the proportion of uron to urea, and the influence of
the same on the resin gel time. Gao and Li adopted the acid-step method
(see Section 3.5 B) and found that uron resins showed a distinct effect on
decreasing free formaldehyde content and formaldehyde emission levels
of UF resin and its bonded plywood; the former can be reduced by 76%
and the latter, 84% [35]. Similar results on the decrease of formaldehyde
emission were reported by Soulard et al. although with a slowing down of
the resin curing reaction [34].
The long duration of the synthesis of uron resins [33, 34] could be short-
ened to 8 h [35]. A mixture of 20 parts of uron resin (synthesized in 8 h)
with 100 parts of UF resins (F/U molar ratio of 1.3) has been reported to
have a potential advantage on industrial applications based on the encour-
aging results on free formaldehyde content, formaldehyde emission levels,
and bond strength. The improved effect of uron resin on the performance
of UF resin can be attributed to the opening of uron ring structure followed
by the reaction with free formaldehyde and its oligomeric glycol forms.

3.8 Reactions of UF during Storage


Understanding of the various reactions occurring during the storage of the
UF resin is very important to control the storage life of the resin [32].

1. A long duration of storage at room temperature or heating at


elevated temperature may convert an alkaline UF resin into a
solid mass. However, a solid mass of this type does not show
the kind of cure required to function as wood adhesive. Such
a transformation of UF into a solid mass is just due to physi-
cal agglomeration rather than chemical cross-linking, which
is essential for the suitability as a wood adhesive.
2. Free formaldehyde present in the UF resin can undergo a
number of reactions during storage: (1) polymerization and
precipitation of the polymer, (2) Cannizzaro reaction, and (3)
oxidation to formic acid. Any of these reactions are detrimen-
tal to product quality. Formaldehyde itself exists in aqueous
solution as methylene glycol, which can react with urea lead-
ing to a product with ether linkages. The ether linkages are
prone to emission of free formaldehyde during storage and
later during the process of curing. Hydrolysis, isomerization,
74 Adhesives for Wood and Lignocellulosic Materials

and decomposition of urea may take place simultaneously


during improper storage conditions, such as high humidity,
high temperatures, and industrial atmosphere [32].

3.9 Reaction Parameters in the Production of Amino


Resins (General)
The reaction parameters that control the performance characteristics of
the amino resin are as follows:
The pH variation sequence
The temperature variation sequence
The types and amount of alkaline and acidic catalysts
The sequence of addition of the different raw materials
The duration of the different reaction steps in the cooking
procedures.

3.10 Four-Step Process for Low Formaldehyde Emission


The mole ratio of formaldehyde to urea has been continuously decreasing
from its high initial value over the years. Reduction of the ratio has the
effect of weakening the internal bond strength of particleboards, although
the residual formaldehyde is reduced. Some resin manufacturers adopt
procedures of programmed addition of formaldehyde and urea in two or
more stages to achieve the desired final F/U ratio.
It is known that the hydrolytic stabilities of the various inter unit linkages
and functional groups in the cured UF resin flows in the following order:

Methylene bridge > dimethylene ether > methylol group

Thus, an ether bridge is not hydrolytically stable compared with a meth-


ylene bridge and tends to evolve formaldehyde during service conditions.
Hence, it is imperative to follow procedures that minimize the formation
of such ether groups in UF resins.
A hydrolytically stable UF resin with a formaldehyde-to-urea mole ratio
of 1:1 was disclosed by Williams in U.S. Patent No. 4,410,685 [36]. The spe-
cial feature of the resin is that it contains no free formaldehyde. A new con-
densation procedure was adopted, which maximized the methylene groups
in the cured resin compared to the methylene ether groups. This ensured low
Urea–Formaldehyde Resins 75

emission of formaldehyde from the cured resin. This is a four-stage conden-


sation process, with the first stage involving a very high acid environment.

3.11 Curing of UF Resins


The curing of the UF resin is effected under acidic conditions. Curing reac-
tion can be considered as an extension of the acidic condensation process
employed in the manufacture of UF resin in the reaction kettle. If the acid
condensation in the kettle is not stopped by neutralization when the desired
viscosity is reached, the resin will gel in the kettle. This should be avoided
during the manufacture of UF resin. In contrast, this reaction under acidic
conditions leading to gel formation at the glue line is the desired reaction
during the curing of the UF resin.

3.11.1 Ammonium Salts


As mentioned above, the curing of the UF resin occurs under acidic condi-
tions provided by hardeners. Direct addition of acids such as maleic acid,
formic acid, phosphoric acid, etc. or of acidic substances, which dissociate
in water (e.g., aluminum sulfate), can cause the curing of the UF resins at
room temperature. The acidic conditions can also be provided by latent
hardeners such as ammonium sulfate reacts with the free formaldehyde in
the resin to generate sulfuric acid, which decreases the pH; this low pH and
hence the acidic conditions enable the restart of the condensation reaction
and finally the gelling and hardening of the resin [37].

2 (NN4) SO4 + 6 CH2O (CH2)6N4 + 2 H2SO4 + 6 H2O

Conversely, ammonium chloride reacts with free formaldehyde to form


hydrochloric acid and hexamethylenetetramine in accordance with the fol-
lowing equation:
4 NH4Cl + 6 CH2O (CH2)6N4 + 4 HCl + 6 H2O

The residual hydrochloric acid in the bondline contributes to the slow


hydrolysis of the aminomethylene bond in the hardened resin. This causes
formaldehyde emission from the wood panel products during actual ser-
vice conditions.
It is generally supposed that, during the early stages of the curing
process, imido groups in the chain react with free formaldehyde present
76 Adhesives for Wood and Lignocellulosic Materials

in solution to give pendant methylol groups Thereafter, the principal


reactions are as follows:

1. Reaction between the pendant or terminal methylol groups


with imido hydrogen to form methylene bridges between the
chains

N N

CH2OH CH2

CO NH CH2 CO N CH2

2. Self-condensation between methylol groups to form methy-


lene bridges with the elimination of water and formaldehyde

N
N
CH2OH
CH2 + CH2O + H 2O
CH2OH
N
N

The role of ammonium sulfate or ammonium chloride on the UF resin


cure is first the release of hydrochloric acid (as stated in the above chemical
equation). The hydrochloric acid thus formed catalyses of the cross-linking
reaction in the UF resin system. The gel time of the UF resin decreases with
the increase of catalyst concentration, the increase of resin solid content,
and the decrease of the pH [38].
The traditional viewpoint, as explained above, is that the curing of UF
resin is a result of the formation of a cross-linked polymeric network struc-
ture [14]. However, there are clear and abundant evidences to show that a
colloidal phase also occurs during resin cure and resin aging [39–43]. The
same is also true in the case of melamine–formaldehyde resins [44, 45].

3.12 Cross-Linked Structure


The final hardened network formed by a urea-formaldehyde resin is a tri-
dimensional network in which the ureas are linked to each others through
Urea–Formaldehyde Resins 77

CH2 CH2

N CO N CH2 N CO N CH2

CH2 CH2

N CO N CH2 N CO N CH2

CH2 CH2

N CO N CH2 N CO N CH2

CH2 CH2

Figure 3.9 Cross-linked structure of cured UF resin.

methylene bridges, although some methylene ether bridges may also occur
(Figure 3.9).

3.13 Triazinone for Curing the UF Resin


In the literature, various other resin preparation procedures are also
described, e.g., yielding of uron structures [46, 47] or triazinone rings in
the resins [48, 49]. The triazinone rings are formed by the reaction of urea
and an excess of formaldehyde under basic conditions in the presence of
ammonia, a primary or a secondary amine, respectively. These resins are
used, among other applications, to enhance the wet strength of paper.
For curing liquid UF resins containing less than 10% by weight of
melamine, a catalyst consisting essentially of a solution of a triazinone [26]
derivative of the formula

C
R N N R

H2O CH2
N

wherein R, being the same or different, is H or CH2OH is used in a pro-


portion of 0.01 to 0.2 mol per 100 g of dry resin. The catalyst is obtained
78 Adhesives for Wood and Lignocellulosic Materials

by reaction of urea and of formaldehyde or of a UF precondensate with an


ammonium salt.

3.14 Distinguishing Feature of UF from other


Synthetic Resin Adhesives such as MUF and PF
The distinguishing feature of UF resins is their high reactivity resulting in
short press times. High productivity of medium thickness particleboards
is thus achievable with the modern and long continuous press lines with
highly reactive UF resins containing optimum amount of hardener and
at high temperature. Moisture content difference maintained between the
glued particles of core and surface imparts the so-called “steam effect”
without an additional provision of steam injection system used in North
American plants [2, 37]. Dunky further cautions how this moisture gradi-
ent should be regulated in practice in order to avoid problems leading to
steam blisters.

3.15 Other Curing Agents


A non-conventional curing system consisting of a simple mixture of UF
resins and polyfunctional hydrazides under neutral conditions has been
developed by Tomita et al. [50]. Low-molecular-weight dihydrazides such
as malonyl, adipoyl, sebacoyl, and isophthaloyl dihydrazides, which were
obtained easily from the corresponding diester of dicarboxylic acids, were
used as the cross-linking agents of UF resins [50].
Examples are as follows:

Malonyl dihydrazide H2N-NH-CO-CH2-CO-NH-NH2


Adipoyl dihydrazide H2N-NH-CO-(CH2)4-CO-NH-NH2
Sebacoyl dihydrazide H2N-NH-CO-(CH2)8-CO-NH-NH2
Isophthaloyl H2N-NH-CO-(C6H5)-CO-NH-NH2

Utilization of hydrazide polymer as a hardener for UF resins was


reported [51]. Poly(methacryloyl hydrazide) (PMAH), which was synthe-
sized easily from poly(methyl methacrylate) (PMMA), was investigated
on its practical applicability as wood adhesive. From a practical point of
view, the new curing system has been reported to have the great advan-
tage of reducing formaldehyde emissions from the final products. There is,
Urea–Formaldehyde Resins 79

however, a difficulty to be solved in this system. The extremely short pot


life will make its practical use difficult.
Other curing agents proposed by Cui and Du are based on PVAc-type
emulsions, including PVAc, the copolymer of PVAc and N-hydroxymethyl
acrylamide (PVAc–NMA), and the ternary copolymer of PVAc, NMA,
and urea (PVAc–NMA–urea) [52]. Water, aluminum chloride, ammonium
dihydrogen phosphate, polypropylene glycol, silicone oil, and urea were
the other components. On heating, aluminum chloride and ammonium
dihydrogen phosphate undergo decomposition and hydrolysis in solution
and produce free acid to cure UF resin. Curing rate is enhanced and the
time of curing is shortened. Further ammonium dihydrogen phosphate
and urea serve as formaldehyde scavengers. Since the methylol groups
contained in the PVAc–NMA and PVAc–NMA–urea system can undergo
cross-linking reaction, the adhesive bonding strength is improved. More
importantly, the results from the industrial production experiments were
shown to be very good.

3.16 Protic Ionic Liquids as a New Hardener-Modifier


System
As has been discussed earlier, ammonium salts of strong inorganic acids,
primarily nitrates, sulfates, and chlorides (the use of the latter discontin-
ued for environmental reasons), inorganic acids (phosphoric acid) and
organics (formic acid, oxalic acid), as well as acid anhydrides [13] are used
for the curing of amino resin adhesives. In order to increase hydrolytic
resistance of adhesive joints, melamine salts are used [52–54].
Research on hardeners alternative to the above has been conducted
independently [53, 55–59].
A new generation of “multitasking” hardener based on Ionic liquids
has been reported by Jóźwiak et al. [60]. Ionic liquids are a highly solvat-
ing, non-coordinating medium in which a variety of organic and inor-
ganic solutes are able to dissolve. They are outstanding good solvents for
a variety of compounds. Ionic liquids are nonvolatile and nonflammable,
have a high thermal stability, and are relatively inexpensive to manufac-
ture. They usually exist as liquids well below room temperature up to a
temperature as high as 200°C [61]. Further, their physical and chemical
properties can be fine-tuned by the adequate selection of the cation and
anion constituents [62].
Dialkylmethylammonium dodecylbenzenesulfonate based protic ionic
liquids (PILs) have been reported as a hardener-modifier system for MUF
80 Adhesives for Wood and Lignocellulosic Materials

adhesive resin [60]. A general procedure for synthesis of PILs has been
described by the authors (Figure 3.10). PILs with the following alkyl
substituents—4 (butyl—C4H9), 6 (hexyl—C6H13), 8 (octyl—C8H17), 10
(decyl—C10H21)—were developed and synthesized according to the fol-
lowing scheme:
The plywood fulfilled the requirements of the EN-314-02 standard in
terms of strength and water resistance of glue lines, regardless of PILs
used.
Younesi-Kordkheili and Pizzi [63, 64] used ionic liquids as harden-
ers to a new generation of urea adhesives [urea–glyoxal (UG) resins] for
wood particleboards formulated without added formaldehyde. These
UG resins [65] and UG–formaldehyde copolymers [66] with ionic liquid
hardeners exhibited a performance that is equivalent to UF resins. There
is further an additional advantage of having no formaldehyde emission,
thereby yielding more acceptable and environmentally friendly adhe-
sives. The ionic liquid employed by them was N-methyl-2-pyrrolidone
hydrogen sulfate prepared in the laboratory according to the proce-
dure of Wang et al. [67]. The energy of activation of the curing of urea–
aldehyde resins was reported to have decreased significantly by the use of
ionic liquid hardeners. Further, the use of aldehydes other than formalde-
hyde for the preparation of urea resin is of noteworthy significance. The
same principle was put to good use to develop melamine–glyoxal (MG)
[68] and MG–glutaraldehyde adhesives [69] for paper impregnation and
plywood to improve markedly the reaction rate by use of an ionic liquid
hardener.
Younesi et al. investigated the effect of polymeric 4,4’-diphenyl meth-
ane diisocyanate (pMDI) on the physical and mechanical properties of
plywood panels bonded with an ionic-liquid-treated lignin–UF resin

C12H25
C12H25
R R
+ +
N NH +
R R
HO3S –O S
3

R = CnH2n+1 where = 4, 6, 8, 10

Figure 3.10 Synthesis of PILs.


Urea–Formaldehyde Resins 81

[70]. Soda lignin modified by 1-ethyl-3-methylimidazolium acetate


([Emim][OAc]) ionic liquid was added to a UF resin during resin synthe-
sis to prepare a lignin–urea–formaldehyde (LUF) resin. pMDI at various
contents (2%, 4%, and 6% on resin solids) was then added to prepare a
LUF resin.

CH3
O
N+
–O CH3
N

CH3

1-Ethyl-3-methylimidazolium acetate

3.17 Improvement of Water Resistance and Adhesive


Performance of UF Resin [71]
Low cost, ease of application, low temperature cure, water solubility,
and excellent mechanical properties when cured are undisputed advan-
tages of UF resins. However, the UF resin system is beset with major
and somewhat serious disadvantage, namely, very poor water resistance
[72–74]. Moist environments, especially when heated, lead to hydrolysis
of methylene and methylene–ether bridges according to the mechanism
proposed by Fleischer and Marutzky [74] as follows:

-(CO)NHCH2NH(CO)- + H2O -(CO)-NH2 + -(CO)-NH-CH2OH

-(CO)-NH-CH2O-CH2NH(CO)- + H2O 2 -(CO)-NH-CH2OH

The ease of hydrolysis of UF resins restricts their usefulness only to inte-


rior applications. Efforts have therefore been made to improve their water
resistance:

(1) Addition of melamine acetate to the UF adhesive [75, 76].


(2) Addition of hydrolyzed waste nylon as a hardener for UF
resin [55].
(3) Incorporation of a hydrophobic chain into the network,
namely, urea-capped aliphatic amines and/or amine hydro-
chlorides as curing agents. This resulted in lower resin sus-
ceptibility to hydrolytic attack [77–80].
82 Adhesives for Wood and Lignocellulosic Materials

3.18 Characterization of UF Resin


The desirable characteristics of the UF resin adhesives are as follows:

(1) Optimum content of oligomeric and monomeric species


and functional groups.
(2) High stability, optimum curing rate, and minimum form-
aldehyde emission.
(3) Acceptable level of mechanical properties should be
ensured.

Formerly, the UF resins were characterized by physical properties such


as viscosity, density, solid content, gel time, and storage life. These char-
acteristics do not completely reflect the structural features of the resin at
the molecular level. The final performance of the resin will be influenced
essentially by the molecular structure, molecular weight, and molecular
weight distribution. Hence, the important properties that will uniquely
characterize the UF resins are as follows:

1. Chemical structural features of UF resins as determined by


1
H-NMR and 13C-NMR
2. Thermochemical characteristics of the resin during the cure
as determined by differential scanning calorimetry (DSC)
3. Molecular weight and molecular weight distribution as
determined by gel permeation chromatography (GPC)
or high-performance liquid chromatography and matrix-
assisted laser desorption/ionization time-of-flight mass
spectrometry (MALDI-TOF MS)
4. Free formaldehyde content of the resin
5. Viscosity
6. Solid content
7. Gel time
8. Storage life

13
3.18.1 C NMR Data
13
C-NMR is a useful technique employed to characterize the UF resin at
the molecular level. Chemical shifts have been assigned by a number of
investigators for identifying different functional groups and interlinking
units in the UF resin [20, 29–32, 81–84].
Urea–Formaldehyde Resins 83

Accordingly, peak assignments and relative intensities are collected and


tabulated as the basic characteristics of the UF resin. Potentially signifi-
cant peak intensity ratios, i.e., the ratios of integrated intensities of peaks
characteristic of chemical groups that are known to contribute to the cured
strength and formaldehyde emission of the UF resin as proposed by Ferg
et al., are calculated and reported as the resin characteristics [85]. These
data can be employed to calculate from various regression equations the
following properties of particleboards bonded with the resin [85]:

1. Degree of resin cross-linking


2. Internal bond strength
3. Formaldehyde emission

3.18.2 Free Formaldehyde Content in the Resin


A number of factors such as formaldehyde-to-urea molar ratio, hardener
system, the type of wood, etc. affect the formaldehyde emission from wood
panel products. But the most important factor is the free formaldehyde con-
tent of the resin before the curing process. For instance, E1 emission class
can be achieved only if the free formaldehyde content of the resin is lower
than 0.2% by weight. UF resins containing higher than 0.5% free formalde-
hyde by weight will not conform to the requirement for the emission class
E2. Since the free formaldehyde content of the UF resin is the determining
factor in the ultimate emission of formaldehyde in the cured product, a low
formaldehyde content must be ensured during resin synthesis.

3.18.3 Molecular Weight and Molecular Weight Distribution


Although UF resins consist of only two main components, i.e., urea and
formaldehyde, they exhibit a broad variety of possible reactions and struc-
tures as explained in Section 3.7.3. This variety leads to a wide range of
molecular species and molecular weight distribution in the final UF res-
ins. The molecular weight may range from very low to very high values
depending on the process conditions employed in the synthesis [15, 23,
86–88].
From the viewpoint of end-use applications of UF resins, the molar
weight distribution is one of the important chemical characteristics, hav-
ing an influence on several important properties of the resin, such as vis-
cosity, flowability, penetration into the wood surface [89–91], distribution
on the wood furnish (particles or fibers), water tolerance, etc.
84 Adhesives for Wood and Lignocellulosic Materials

The molecular weight and the molecular weight distribution can be


determined by

1. Size exclusion chromatography (SEC)


2. MALDI-TOF MS method

3.18.4 Size Exclusion Chromatography


SEC or GPC is a standard technique for determining average molec-
ular weight and molecular weight distribution of a polymer. Ludlam
and King have developed a rapid and reproducible method for inves-
tigating the molecular weight distribution of UF resin by SEC [92]. In
order to enable polymer of high viscosity and high molecular weight
to dissolve easily in the solvent (dimethyl formamide), a concentrated
lithium chloride solution was used. Dimethyl formamide containing
lithium chloride breaks up hydrogen bonding and ensures realistic val-
ues of molecular weight averages. Braun and Bayersdorf monitored the
growth of UF reaction products using GPC [93]. They employed polyvi-
nyl acetate gel and DMF/DMSO solvent mixture to investigate the influ-
ence of pH and F/U molar ratio on molecular weight distribution The
influence of pH, temperature, and F/U molar ratio was also examined
by Kumlin and Simonson [94]. They employed GPC, which permitted a
separation of the various low-molecular-weight components present in
a UF resin. The results further showed that the dimethyl ether linkages
are favored at alkaline pH and increased with increasing pH from 8.0 to
9.4, whereas the formation of methylene bridges was strongly favored
when the reaction mixture was made acidic and increased with decreas-
ing pH from 5.1 to 3.5.

3.18.5 MALDI-TOF MS Method


MALDI-TOF MS has been found to be a powerful method for the charac-
terization of both synthetic and natural polymers [58, 95–97]. This tech-
nique is usually combined with a TOF mass analyzer. Only a very small
amount of analyte is required.
The MALDI-TOF MS method can also throw light on the nature of
the structural units and the interlinking units in polymeric resins, sepa-
rating clearly different types of oligomers without degrading them down
to ion pieces as standard MS spectrometry. It renders then easy the iden-
tification of the types of reactions that have occurred and the products
formed. For instance, Du et al. employed (MALDI-TOF) MS to study
Urea–Formaldehyde Resins 85

phenol–urea–formaldehyde (PUF) resins [95]. The results indicate in all


the PUF resins tried the proportions of:

(1) Methylene links between phenol and urea units in the


co-condensate.
(2) Methylene bridges that connect between phenol to phenol
units and urea to urea units.

Their relative proportions were determined and related to the synthesis


procedures by Gavrilovic-Grmusa et al. [98]. Despres et al. utilized C13
NMR and particularly MALDI-TOF to follow first the reaction of urea and
formaldehyde and then to follow the reaction of the UF polymer so formed
with melamine to follow the preparation of MUF adhesives [99]. They
identified all the different species formed along the preparation diagram of
first UFs and then MUFs [99]. The analysis allowed us to identify and fol-
low the appearance, increase, decrease, and disappearance of a multitude
of chemical species during the preparation of both the initial UF phase of
the reaction and the subsequent reaction of melamine with the UF resin
that formed. The analysis indicated [99] that (1) the increase and decrease
in the species that formed proceeded through a cycle of the formation and
degradation of species occurring continuously through what appeared
to be a series of complex equilibria, (2) even at the end of the reaction
a predominant proportion of methylene ether bridges was still present,
(3) some small proportion of methylene bridges already had formed in the
UF reaction phase of the resin even under rather alkaline conditions, and
(4) the addition of melamine to the UF prepolymer induced some notice-
able rearrangement of methylene ether bridges to methylene bridges. The
main results were confirmed by FTIR analysis [100].

3.18.6 Cure Time


Cure time of UF resins is strongly affected by the following variables and
wood-related factors:

(1) Wood extractives [101, 102].


(2) Wood pH values and buffering capacities [103–107]. Park
et al. determined that the fiber acidity strongly affected the
internal bond strength of medium-density fiberboard pan-
els bonded with a UF resin [108]. Xing et al. also reported
that the pH value and buffering capacities influence the gel
time of UF resin [106].
86 Adhesives for Wood and Lignocellulosic Materials

(3) The wood particle size. Medved and Resnik [109] sug-
gested that reducing the wood particle size could reduce
the gel time of UF resins.

3.18.7 Differential Scanning Calorimetry


DSC measures the temperatures and heat flows associated with transi-
tions in materials as a function of temperature or time in a controlled
atmosphere. This technique provides quantitative and qualitative infor-
mation about physical and chemical changes that involve endothermic
or exothermic processes, or changes in heat capacity. DSC is ideally
suited for the study of curing reactions of thermostetting polymers that
are normally exothermic in nature. The effect of operating variables on
the rate of cure of the reins and the accompanying exothermic heat are
of great practical importance. Stepwise polymerization reactions under
normal atmospheric conditions are accompanied by the evolution of vol-
atile products such as water, alcohol, etc. The consequent endothermic
peaks can be large enough to obliterate exotherm caused by cross-linking
reactions during curing. Such condensation reactions therefore require
encapsulation or pressurization if the curing exotherm is to be observed.
Sealed capsules must be used in which pressure can build up during the
course of a reaction [110].
Valuable information has been collected on the DSC measurements
on UF resin curing. Szestay et al. have investigated the curing process of
a variety of UF resin adhesives suitable for particleboards and plywoods
using DSC in anisothermal mode [111]. They found that the maximum
curing reaction rate was detected in the range of 80–85°C. Elimination of
formaldehyde was observed at 105–150°C. At temperatures higher than
150°C, thermal decomposition of methylene linkage takes place, which
affects the mechanical properties of the adhesion.
Myers and Koutsky have examined the resin cure behavior of UF resins
using DSC methods [112]. Several UF resins containing different catalyst
systems and potential acid-neutralizing additives were examined by DSC.
They have found that the more acidic catalyst–additive system cures the
resin more completely and at lower temperature. However, the more acidic
system also exhibits greater post-cure formaldehyde liberation.
Šebenik et al. have also determined the kinetics of reaction between urea
and formaldehyde in acid and neutral medium by using the DSC method
[113]. From their studies, the activation energy, heat of polymerization,
reaction order, and temperature maximum have been determined.
Urea–Formaldehyde Resins 87

Overall, these studies demonstrate that a great variety and diversity of


UF structure leading to resins with different performance can be produced
by manipulating the synthetic condition. In turn, the control of the syn-
thetic conditions during industrial synthesis is critical in defining the qual-
ity and reproducibility of the product.

3.19 UF Resin Cure Kinetics


When UF resin is mixed with the hardener, the cure reaction rates can be
measured using a differential scanning calorimeter [13].
DSC investigation involves calculation of energy of activation by the
Kissinger equation. This equation relates the ln of the heating rate (β)
divided by the square of maximum of temperature of the exotherm (Tmax)
as a function of the inverse of maximum temperature reached by the exo-
therm being examined.

Ea
In /(Tmax )2 In ( AR/E )+ C
RT max

From the Kissinger equation [114], after a series of scans at various


heating rates, a linear relationship is observed between ln β/(Tmax)2 and
1/Tmax. From this, the energy of activation can be calculated.

3.20 UF Resins with Low Formaldehyde Emission


The importance of reduced formaldehyde emission from wood panel
products has been discussed in another chapter. The reduction in formal-
dehyde emission levels from products bonded with UF adhesive resins has
been achieved by employing one or more several technological methods.
In general, these methods include the following:

1. Changing the formulation of the UF adhesive resin (e.g.,


lowering the F/U ratio).
2. Adding formaldehyde-scavenging materials directly to the
UF adhesive resin.
3. Separately adding formaldehyde-scavenging materials to the
wood finish.
88 Adhesives for Wood and Lignocellulosic Materials

4. Treating panels after their manufacture either with a formalde-


hyde scavenger or by the application of coatings or laminates.
5. Changing to an entirely different adhesive resin system.

Recent research has suggested possible new methods to lower formalde-


hyde emission levels. This research involves two strategies: the modification
of the chemistry of UF resins and the replacement of the formaldehyde com-
ponent in UF resins with a less volatile aldehyde or its chemical equivalent.

3.21 Modification by Polyamines


Polyamines can function as internal plasticizers for UF resins. As a result
of internal plasticization, they can reduce the internal stresses developed
during the resin curing. This in turn enhances the stability to withstand
cyclic stresses and hence the durability [77–80].
Polyamine modification was carried out by Conner in the following
ways [115]:

1. Polyamines can be incorporated directly during the synthe-


sis of UF resin.
2. Polyamines can be converted into urea-capped derivatives
and incorporated in the resin synthesis.
3. Ammonium chloride can be replaced by polyamine hydro-
chloride salts as curing catalyst.
4. A combination of methods 2 and 3 can be employed.

The polyamines in question are as follows (Figure 3.11):

H2N-(CH2)6-NH2
CH2-[OCH2CHCH3]x-NH2
Hexamethylenediamine
CH3CH2 C CH2-[OCH2CHCH3]y-NH2

CH2-[OCH2CHCH3]z-NH2
(x + y + z = 5.3)
H2N-(CH2)6-NH-(CH2)6-NH2
bis-hexamethylenetriamine Poly(propyleneoxide)triamine

(H2N-CH2-CH2)3N H2N-CH2[OCH2CHCH3]33-NH2

Triethylaminetriamine Poly(propyleneglycol)diamine

Figure 3.11 Structure of polyamines.


Urea–Formaldehyde Resins 89

3.22 Cyclic Urea Prepolymer


Cyclic urea prepolymer has been found to be useful as a modifier in phenol–
formaldehyde resins and melamine–formaldehyde resins. It is produced by
a reaction of urea, formaldehyde, and ammonia or a primary amine.

3.22.1 Preparation of Cyclic Urea Prepolymer [116]


US Patent 6,379,814 discloses the method to produce cyclic urea prepoly-
mer from urea, formaldehyde, and ammonia in the ratio 1.0:2.0:0.5 at a
temperature of 90°C for about 1 h. After the completion of the reaction, the
product is cooled to room temperature. 13C-NMR confirmed the following:

(i) About 42.1% of urea was contained in the triazone ring.


(ii) About 28.5% of the urea was di/tri-substituted.
(iii) About 24.5% of the urea was mono-substituted.
(iv) About 4.9% of the urea was free.

3.22.2 Cyclic Urea Prepolymer as a Modifying Resin


for other Adhesives
Cyclic urea prepolymer may be used as a modifier for thermosetting
phenol/HCHO-and melamine/HCHO-based resins for a variety of end
uses.
The modification can be by either blending or co-condensation. A num-
ber of patents have been filed on the use of cyclic urea prepolymers in mod-
ifying UF, phenol–formaldehyde, and melamine–formaldehyde resins.
For instance, Dopico et al. in US Patent 6,399,719 deals with cyclic UF
prepolymer-modified MUF resins [117]. Phenol–urea–melamine resins
can be modified according to Dupre et al. in US Patent 6,379,814 [116],
and phenol–urea resin modification by cyclic urea prepolymer is covered
by Schmidt et al. in the patent application US 2006/0100412A1 [107].

3.23 Improvement of UF and MUF Resins by


Addition of Hyperbranched Dendrimers
An interesting development for application to UF and MUF resins is the
use of small proportions of hyperbranched dendrimers as an upgrading
additive. Hyperbranched poly(amidoamine)s exhibiting various levels of
90 Adhesives for Wood and Lignocellulosic Materials

hydrophobicity have been used as modifiers for UF [118–120] and MUF


[121–123] adhesives. Co-condensability of these additives with the UF and
MUF resins was shown to occur. The use of these additives as MUF mod-
ifiers resulted in manifold advantages. Their addition either immediately
before the resin final use or at the last stage of resin preparation yielded
considerable upgrading of the dry internal bond (IB) strength of the parti-
cleboards bonded with the modified MUF resin.

3.23.1 Urea and Melamine Resins without Formaldehyde


Urea is a natural raw material. It is also obtained industrially in enormous
quantities by the catalytic reaction of oxygen and nitrogen of the air on
glowing coals or other glowing carbon materials, even charcoal or wood.
The material to substitute is formaldehyde, although even this can be of
natural origin. Formaldehyde is now classified as toxic and oncogenic. To
substitute it, the purely commercial need to maintain the resulting resin
white must be respected, a fact that greatly complicates formaldehyde sub-
stitution. It pays, seeing the volumes involved, to start concentrating on
developing urea-based adhesives using aldehydes that are not toxic or vol-
atile, but still maintaining the clear or white appearance of UF resins as
such adhesives can be classified as natural too. While many approaches
can be taken to develop urea-based adhesives in this manner, recently a
first important success in the bonding of plywood with this approach was
achieved [65]. Resins based on UG for textiles are well known [124], but
these are low-condensation resins not adaptable for wood. Hybrid resins
UF–another aldehyde have been the initial target of several researchers,
with good results for plywood [65, 125, 126]. The problem of these is that
formaldehyde, although in much lower proportion, is still there. An old
technology based on urea–furfural resins [127] works but is not a good
substitute for UF not only for the furfural lower reactivity but mainly
for the dark color of the resin imparted by the condensation of furfural.
Equally, the use of urea–furfuryl alcohol or urea–hydroxymethylfurfural
while yielding better cross-linking also presents some of the drawbacks
outlined for urea–furfural resins. Thus, the first truly urea adhesive for ply-
wood without any formaldehyde has only recently been developed, open-
ing a new chapter on natural environmental friendly adhesives [65].
The nonvolatile and nontoxic aldehyde glyoxal (G) was used to substi-
tute formaldehyde to react with urea (U) to synthesize a UG resin under
weak acid conditions (pH = 4–5). The strength of the bonded plywoods
was tested, and the curing process of the UG resin was studied. The results
showed that the bonded plywood could be used as interior decoration and
Urea–Formaldehyde Resins 91

furniture material without formaldehyde emission in dry conditions [65].


The problem of these resins is that while they are suitable for plywood
application, they are unsuitable for the faster-pressing particleboard appli-
cation due to cross-linking difficulty. Their cured behavior is that of mainly
a physically entangled network formed by very long linear chains, rather
than of a chemically cross-linked network. However, this considerable
drawback was recently solved by using an ionic liquid as the hardener of
the UG adhesive resin [63, 64]. Thus, a UG resin was hardened with 1%
to 3% N-methyl-2-pyrrolidone hydrogen sulfate-produced particleboards
with excellent IB strength results, marked lowering of the resin energy of
activation, and resin gelling and hardening times comparable to UF resins
at equivalent press temperatures, well in line with today’s industrial prac-
tice. This constitutes a breakthrough as it renders feasible urea resins that
can be classed as renewable resources using a nontoxic and nonvolatile
aldehyde and with no formaldehyde emission simply because no formalde-
hyde is used.
Melamine is synthesized starting from urea according to the reaction

(NH2)2CO → C3H6N6 + 6 NH3 + 3 CO2

Thus, it can also be considered a biosourced chemical compound. MG


resins were also synthesized with different M/G molar ratios, and their
properties were tested [68]. The results showed that the synthesized MG
resins remain stable for at least 10 days after preparation at ambient tem-
perature. Conjugated structures and large amounts of -OH, -NH-, and
C-N groups with different substitution levels exist in the MG resins pre-
pared. Again, long linear chains are preferentially formed. For these resins,
however, chemical cross-linking was achieved by using chromium nitrate
as a cross-linking catalyst. Gel times of around 3 min were obtained, but
the temperature needed to obtain these was 150°C rather than the tradi-
tional 100°C. This means that the lower reactivity of the aldehyde involved
causes a much higher energy of activation barrier to advance to curing for
these adhesive resins [68]. This reactivity problem was solved by prepar-
ing double aldehyde resins, namely, MG–glutaraldehyde resins where the
energy of activation of hardening was again markedly decreased by using
ionic liquids as hardeners. These resins performed well as wood adhesives
without any formaldehyde [69].
Aminoresin precursors prepared by the addition of a new, colorless,
nonvolatile, and nontoxic aldehyde, dimethoxyethanal (DME), a prod-
uct derived from glyoxal, to melamine or urea gave resins for boards that
were able to harden [97]. However they were underperforming because of
92 Adhesives for Wood and Lignocellulosic Materials

the lower reactivity of DME in relation to formaldehyde. Melamine and


urea react with one and two (melamine up to three) molecules of DME
to form M-DME and U-DME, but the subsequent cross-linking reaction
to form bridges does not occur unless the reaction is catalyzed during
resin preparation by the addition of glyoxylic acid. Such bridges between
two melamine molecules form only up to the formation of dimers and no
more. The use of glyoxylic acid during the reaction has allowed the forma-
tion of different oligomers by both aldol condensation and condensation
of melamine and glyoxylic acid with two molecules of melamine to form
dimers. These were observed by 13C-NMR and MALDI-TOF spectrometry.
However, the addition of 20% isocyanate (pMDI) was necessary to satisfy
the relevant mechanical strength standards for panels prepared with these
resins. pMDI contributed to cross-linking of M-DME and U-DME by its
reaction to form urethane bridges according to reactions already described
[128–130]. The adhesive resins so formed had excellent performance and
were colorless, and they produced boards that well satisfied the require-
ments of the relevant standards for interior panels. Formaldehyde emis-
sion was down to what would be expected by just heating the wood chips
in absence of adhesives. The panels emission was sufficiently low to even
satisfy the most severe relevant F**** level of the JIS A 5908 [131] and JIS
A 1460 [132] Japanese standards. These adhesives are colorless as MUF
and UF resins.

References
1. Moubarik, A., Pizzi, A., Allal, A., Charrier, F., Khouk, A., Charrier, B.,
Cornstarch–mimosa tannin–urea formaldehyde resins as adhesives in the
particleboard production. Starch—Starke, 62, 131–138, 2010.
2. Dunky, M., Urea–formaldehyde (UF) adhesive resins of wood. Int. J. Adhes.
Adhes., 18, 95–107, 1998.
3. Plastics Historical Society. (http://plastiquarian.com/?page_id=14236).
4. Widmer, G., Amino resins, in: Encyclopedia of Polymer Science and
Technology. Plastics, Resins, Rubbers, Fibers, N.M. Bikales (Ed.), pp. 1–92,
Interscience Publishers, New York, 1965.
5. de Jonge, J.L. and de Jonge, J.J., The reaction of urea with formaldehyde. Rec.
Trav. Chim., 71, 643–660, 1952.
6. de Jonge, J.L. and de Jonge, J.J., The formation and decomposition of dimeth-
ylolurea. Rec. Trav. Chim., 71, 661–667, 1952.
7. de Jonge, J.L. and de Jonge, J.J., The reaction between urea and formaldehyde
in concentrated solutions. Rec. Trav. Chim., 71, 890–898, 1952.
Urea–Formaldehyde Resins 93

8. de Jonge, J.L. and de Jonge, J.J., The reaction of methylene diurea with form-
aldehyde. Rec. Trav. Chim., 72, 213–217, 1953.
9. Bettelheim, L. and Cedwall, J., The slow condensation to polymers of methy-
lolureas. Sven. Kemisk Tidskr., 60, 208–212, 1948.
10. Smythe, L.E., A kinetic study of the urea–formaldehyde reaction. J. Phys.
Coll. Chem., 51, 369–378, 1947.
11. Smythe, L.E., Urea–formaldehyde kinetic studies. I. Variation in urea solu-
tions. J. Amer. Chem. Soc., 73, 2735–2738, 1951.
12. Vale, C.P. and Taylor, W.G.K., Aminoplastics, London Iliffe Books. Ltd,
Published for the Plastics Institute, 1964.
13. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York,
1994.
14. Saunders, K.J., Organic Polymer Chemistry, Chapman and Hall, London, 1973.
15. Christjanson, P., Pehkb, T., Siimera, K., Structure formation in urea–
formaldehyde resin synthesis. Proc. Est. Acad. Sci. Chem., 55, 4, 212–225,
2006.
16. Growe, G.A. and Lynch, C.L., Urea–formaldehyde kinetic studies. J. Amer.
Chem. Soc., 70, 3795–3797, 1948.
17. Growe, G.A. and Lynch, C.L., Polarographic urea–formaldehyde kinetic
studies. J. Amer. Chem. Soc., 71, 3731–3733, 1949.
18. Growe, G.A. and Lynch, C.L., Polarographic acetamide–formaldehyde and
benzamide–formaldehyde kinetic studies. J. Amer. Chem. Soc., 72, 3622–
3623, 1950.
19. Christjanson, P., Siimer, K., Pehk, T., Lasn, I., Structural changes in urea–
formaldehyde resins during storage. Holz Roh Werkst., 60, 379–384, 2002.
20. Kim, M.G., Wan, H., No, B.Y., Nieh, W.L., Examination of selected synthe-
sis and room temperature storage parameters for wood adhesive-type urea–
formaldehyde resins by 13C NMR spectroscopy, IV. J. Appl. Polym. Sci., 82,
1155–1169, 2001.
21. Rayner, C.A.A., Synthetic organic adhesives, in: Adhesion and Adhesives, 2nd
edn., vol. 207, R. Houwink and G. Salomon (Eds.), pp. 187, vol. 1—Adhesives,
Elsevier, New York, 1965.
22. Pizzi, A., Lipschitz, L., Valenzuela, J., Theory and practice of the preparation
of low formaldehyde emission UF adhesives for particleboard. Holzforschung,
48, 254–261, 1994.
23. Dunky, M., Urea–Formaldehyde (UF) Glue Resins: An Adhesive Ever
Young. The 5th Pacific Rim Bio-Based Composites Symposium, Proceedings,
Canberra, Australia, pp. 205–213, 2000.
24. Pizzi, A., Aminoresin wood adhesives, in: Wood Adhesives Chemistry and
Technology, vol. 1, A. Pizzi (Ed.), Marcel Dekker, New York, 1983.
25. Ormondroyd, G.A., Adhesives for wood composites, in: Wood Composites,
G.A. Ormondroyd and R.J. Ball (Eds.), Woodhead Publishing, Cambridge,
UK, 2015.
94 Adhesives for Wood and Lignocellulosic Materials

26. Meyer, B., Urea–Formaldehyde Resins, Addison-Wesley, Boston, Massachusetts,


1979.
27. Kumlin, K. and Simonson, R., Urea–formaldehyde resins: Formation of con-
densation products during resin preparation. Angew. Makromol. Chem., 93,
27–42, 1981.
28. Kumlin, K. and Simonson, R., Urea–formaldehyde resins. 1. Separation of
low molecular weight components in urea–formaldehyde resins by means of
liquid chromatography. Angew. Makromol. Chem., 68, 175–184, 2003.
29. Kim, M.G., Examination of selected synthesis parameters for typical wood
adhesive-type urea–formaldehyde resins by 13C NMR spectroscopy I.
J. Polym. Sci. A, 37, 995–1007, 1999.
30. Kim, M.G., Examination of selected synthesis parameters for typical wood
adhesive-type urea–formaldehyde resins by 13C-NMR spectroscopy II.
J. Appl. Polym. Sci., 75, 1243–1254, 2000.
31. Kim, M.G., Examination of selected synthesis parameters for wood adhesive-
type urea–formaldehyde resins by 13C NMR spectroscopy III. J. Appl. Polym.
Sci., 80, 2800–2814, 2001.
32. Kim, M.G., Young, B., Lee, S.M., Nieh, W.L., Examination of selected synthe-
sis and room-temperature storage parameters for wood adhesive-type urea–
formaldehyde resins by 13C-NMR spectroscopy V. J. Appl. Polym. Sci., 89,
1896–1917, 2003.
33. Gu, J., Higuchi, M., Morita, M., Hse, C.-Y., Synthetic conditions and chemical
structures of urea–formaldehyde resins I. Properties of the resins synthesized
by three different procedures. Mokuzai Gakkaishi, 41, 1115–1121, 1995.
34. Soulard, C., Kamoun, C., Pizzi, A., Uron and uron–urea–formaldehyde
resins. J. Appl. Polym. Sci., 72, 277–289, 1999.
35. Gao, W. and Li, J., Influence of uron resins on the performance of UF resins
as adhesives for plywood. Maderas Cie. Tecnol., 14, 3–12, 2012.
36. J.H. Williams, Hydrolytically stable urea–formaldehyde resins and process
for manufacturing them. US Patent 4,410,685, 1983.
37. Dunky, M., Pizzi, A., Van Leemput, M., COST Action E13-WG 1 Report on
the State of the Art 3, 1st Edition, European Commission, Bruxelles, Belgium,
2002.
39. Dunker, A.K., Johns, W.E., Rammon, R.M., Farmer, B., Johns, S.J., Slightly
bizarre protein chemistry: Urea–formaldehyde resin from a biochemical
perspective. J. Adhes., 19, 153–176, 1986.
40. Jahromi, S., The storage stability of melamine formaldehyde resin solutions:
III. Storage at elevated temperatures. Polymer, 40, 5103–5109, 1999.
41. Jahromi, S., Litvinov, V., Geladé, E., Physical gelation of melamine formal-
dehyde resin solutions. II. A combined light-scattering and low-resolution
relaxation proton NMR study. J. Polym. Sci., Phys. Ed., 37, 3307–3318, 1999.
42. Mijatovic, J., Binder, W.H., Kubel, F., Kantner, W., Studies on the stability of
MF resin solutions: Investigations on network formation, Macromol. Symp.,
181, 373–382, 2002.
Urea–Formaldehyde Resins 95

43. Despres, A. and Pizzi, A., Colloidal aggregation of aminoplastic polyconden-


sation resins: UF vs. MF and MUF resins, J. Appl. Polym. Sci., 100, 1406–
1412, 2006.
44. Zanetti, M. and Pizzi, A., Colloidal aggregation of MUF polycondensation
resins: Formulation influence and storage stability, J. Appl. Polym. Sci., 91,
2690–2699, 2004.
45. Pizzi, A., George, B., Zanetti, M., Meausoone, P.-J., Rheometry of ageing of
colloidal melamine–urea–formaldehyde polycondensates, J. Appl. Polym.
Sci., 96, 655–659, 2005.
46. K. Erhardt, H. Petersen, W. Reuther, O. Wittmann, F. Brunnmüller, O.
Grabowsky, J. Mayer, J. Lenz, Verfahren zur Hesrstellung aminoplastischer
Holzleime, German Patent DE 2 207 921, assigned to BASF, 1972.
47. D. Merkel, H. Petersen, O. Wittmann, J. Mayer, H. Schatz, Bindemittel fur die
Holzverleimung, German Patent DE 25 50 739, assigned to BASF, 1975.
48. R.W. Auten, and V.C. Meunier, Urea resins containing dimethylamino sub-
stituents, US Patent 2,605,253, assigned to Rhom and Haas, 1950.
49. J. Davidson and E.J. Romatowski, Production of ethylenepolyamine-modified
urea–formaldehyde resins, US Patent 2,683,134, assigned to Allied Chemical
and Dye Corp, 1951.
50. Tomita, B., Osawa, H., Hse, C.-Y., Myers, G., New curing system of urea–
formaldehyde resins with polyhydrazides I—Curing with dihydrazide com-
pounds. Mokuzai Gakkaishi, 35, 455–459; 736–741, 1989.
51. Cui, H. and Du, G., Development of a novel polyvinyl acetate type emulsion
curing agent for urea formaldehyde. Wood Sci. Technol., 47, 105–119, 2013.
52. Weinstabl, A., Binder, W.H., Gruber, H., Kantner, W., Melamine salts as
hardeners for urea formaldehyde resins. J. Appl. Polym. Sci., 81, 1654–1551,
2001.
53. Zanetti, M. and Pizzi, A., Low addition of melamine salts for improved
melamine–urea–formaldehyde adhesive water resistance. J. Appl. Polym. Sci.,
88, 287–292, 2004.
54. Zanetti, M. and Pizzi, A., Upgrading of MUF polycondensation resins by
buffering additives. II. Hexamine sulfate mechanisms and alternate buffers.
J. Appl. Polym. Sci., 90, 215–226, 2003.
55. Wang, S. and Pizzi, A., Waste nylon fibre hardeners for improved adhesives
water resistance. Holz Roh Werkst., 55, 91–95, 1997.
56. Proszyk, S., Krystofiak, T., Jóźwiak, M., Lis, B., Investigations on the strength
and durability of glue lines from MUF adhesives at various loading, Proc. of
IVth Inter. Symp. Composite Wood Materials, TU Zvolen, pp. 219–224,
2002.
57. Pizzi, A., Beaujean, M., Zhao, C., Properzi, M., Huang, Z., Acetal-induced
strength increases and lower resin content of MUF and other polycondensa-
tion adhesives, J. Appl. Polym. Sci., 84, 2561–2571, 2002.
58. Zanetti, M., Pizzi, A., Beaujean, M., Pasch, H., Rode, K., Dalet, P.J.,
Acetals-induced strength increase of melamine–urea–formaldehyde (MUF)
96 Adhesives for Wood and Lignocellulosic Materials

polycondensation adhesives. II. Solubility and colloidal state disruption.


J. Appl. Polym. Sci., 86, 1855–1862, 2002.
59. Kamoun, C., Pizzi, A., Zanetti, M., Upgrading melamine-urea–formaldehyde
polycondensation resins with buffering additives. I. The effect of hexamine
sulfate and its limits. J. Appl. Polym. Sci., 90, 203–214, 2003.
60. Jóźwiak, M., Pernak, J., Kot, M., Protic ionic liquids as a new hardener–
modifier system for melamine–urea–formaldehyde adhesive resins.
Drewno. Pr. Nauk. Donies. Komunik., 54, 186, 5–14, 2011.
61. Vickery, J., What are ionic liquids?. Green Chemistry University of Bristol,
2004, (http://www.chm.bris.ac.uk/webprojects2004/vickery/ionic_liquids.
htm).
62. Perez de los Ríos, A., Irabien A, A., Hollmann, F., Fernandez, F.J.H., Ionic
liquids: Green solvents for chemical processing. J. Chem., 1–2, 2013. Article
ID 402172.
63. Younesi-Kordkheili, H. and Pizzi, A., Acid ionic liquids as a new hardener in
urea–glyoxal adhesive resins. Polymers, 8, 3, 57–71, 2016.
64. Younesi-Kordkheili, H. and Pizzi, A., Ionic liquids as enhancers of urea–
glyoxal panel adhesives as substitutes of urea–formaldehyde resins. Eur. J.
Wood Prod., 75, 481–483, 2017.
65. Deng, S., Du, G., Li, X., Pizzi, A., Performance and reaction mechanism of
zero formaldehyde-emission urea–glyoxal (UG) resin. J. Taiwan. Inst. Chem.
Eng., 45, 2029–2038, 2014.
66. Deng, S., Pizzi, A., Du G, G., Zhang, J., Zhang, J., Synthesis and chemical
structure of a novel glyoxal–urea–formaldehyde (GUF) co-condensed res-
ins with different MMU/G molar ratios by 13CNMR and MALDI-TOF-MS.
J. Appl. Polym. Sci., 131, 21, 2014.
67. Wang, X., Han, M., Wan, H., Yang, C., Guan, G., Study on extraction of thio-
phene from model gasoline with Brønsted acidic ionic liquids. Fron. Chem.
Sci. Eng., 5, 107–112, 2011.
68. Deng, S., Pizzi, A., Du, G., Lagel, M.C., Delmotte, L., Abdalla, S., Synthesis,
structure characterization and application of melamine–glyoxal adhesive
resins. Eur. J. Wood Prod., 76, 283–296, 2018.
69. Xi, X., Pizzi, A., Amirou, S., Melamine–glyoxal–glutaraldehyde wood panel
adhesives without formaldehyde. Polymers, 10, 22, 2018.
70. Younesi-Kordkheili, H., Pizzi, A., Mohammadghasemipour, A., Improving
properties of ionic liquid-treated lignin–urea–formaldehyde resins by small
addition of isocyanate for wood adhesive. J. Adhes., 94, 406–419, 2018.
71. Maminski, M.L. and Pawlicki, J., Improved water resistance and adhesive per-
formance of a commercial UF resin blended with glutaraldehyde. J. Adhes.,
82, 629–641, 2006.
72. Freeman, H.G. and Kreibich, R.E., Estimating durability of wood adhesives
in vitro. For. Prod. J., 7, 39–43, 1968.
73. Troughton, G.E., Kinetic evidence for covalent bonding between wood and
formaldehyde glues, Forest Product Lab, Vancouver, 1967, Report VP-X-.
Urea–Formaldehyde Resins 97

74. Fleischer, O. and Marutzky, R., Hydrolyse von Harnstoff-Formaldehyd.


Harzen: Auflösung des Spangefüges in Holzwerkstoffen durch hydroly-
tischen Abbau der Leimfuge. Holz Roh Werkst., 58, 295–300, 2000.
75. Prestifilippo, M., Pizzi, A., Norback, H., Lavisci, P., Low addition of melamine
salts for improved UF adhesives water resistance. Holz Roh Werkst., 54, 393–
398, 1996.
76. Cremonini, C. and Pizzi, A., Improved waterproofing of UF plywood adhe-
sives by melamine salts as glue-mix hardeners. Holzforschung, 49, 11–15,
1997.
77. Ebewele, R.O., Myers, G.E., River, B.H., Koutsky, J., Polyamine-modified
urea–formaldehyde resins. I. Synthesis, structure, and properties. J. Appl.
Polym. Sci., 42, 2997–3012, 1991.
78. Ebewele, R.O., Myers, G.E., River, B.H., Koutsky, J., Polyamine-modified
urea—formaldehyde resins. 2. Resistance to stress-induced by moisture
cycling of solid wood joints and particleboard. J. Appl. Polym. Sci., 42, 1483–
1490, 1991.
79. Ebewele, R.O., Myers, G.E., River, B.H., Polyamine-modified urea–
formaldehyde-bonded wood joints. III. Fracture toughness and cyclic
stress and hydrolysis resistance. J. Appl. Polym. Sci., 49, 229–245, 1993.
80. Ebewele, R.O., River, B.H., Myers, G.E., Behavior of amine-modified urea–
formaldehyde-bonded wood joints at low formaldehyde/urea molar ratios.
J. Appl. Polym. Sci., 52, 689–700, 1994.
81. Ebdon, J.R. and Heaton, P., Characterization of urea–formaldehyde adducts
and resins by 13C-NMR spectroscopy. Polymer, 18, 971–974, 1977.
82. Tohmura, S., Hse, C.-Y., Higuchi, M., Formaldehyde emission and high sta-
bility of cured urea–formaldehyde resins. J. Wood Sci., 46, 303–309, 2000.
83. Kim, S. and Kim, H.J., Evaluation of formaldehyde emission of pine and
wattle tannin-based adhesives by gas chromatography. Holz Roh Werkst., 62,
101–106, 2004.
84. Panamgama, L.A. and Pizzi, A., A 13C NMR analysis method for MUF and
MF resins strength and formaldehyde emission. J. Appl. Polym. Sci., 59,
2055–2068, 1996.
85. Ferg, E.E., Pizzi, A., Levendis, D., A 13C NMR analysis method for urea–
formaldehyde resin strength and formaldehyde emission. J. Appl. Polym. Sci.,
50, 907–915, 1993.
86. Dunky, M. and Lederer, K., Untersuchungen der Molgewichtsverteilung von
Harnstoff-Formaldehyd-Leimharzen. Angew. Makromol. Chem., 102, 199–
213, 1982.
87. Billiani, J., Lederer, K., Dunky, M., Untersuchung der molmassenverteilung
von harnstoff-formaldehyd-leimharzen durch GPC gekoppelt mit lichtstreu-
ung. Angew. Makromol. Chem., 180, 199–208, 1990.
88. Huber, C. and Lederer, K., Flow-rate dependent degradation of high-
molecular-weight polyisobutylene in GPC. J. Polym. Sci., Polym. Lett., 18,
535–540, 1980.
98 Adhesives for Wood and Lignocellulosic Materials

89. Scheikl, M. and Dunky, M., Computerized static and dynamic contact-angle
measuring methods in connection with the wettability of wood. Holz Roh
Werkst., 54, 113–117, 1996.
90. Scheikl, M. and Dunky, M., Urea formaldehyde resins as liquid phase in the
wetting of wood. Holzforsch. Holzverwert., 48, 55–57, 1996.
91. Scheikl, M. and Dunky, M., Measurement of contact angles on wood.
Holzforschung, 52, 89–94, 1998.
92. Ludlam, P.R. and King, J.G., Size exclusion chromatography of urea form-
aldehyde resins in dimethylformamide containing lithium chloride. J. Appl.
Polym. Sci., 29, 3863–3872, 1984.
93. Braun, D. and Bayersdorf, F., Gelchromatographische Untersuchung von
Harnstoff-Formaldehyd-Harzen. Angew. Makromol. Chem., 85, 1–13, 1980.
94. Kumlin, K. and Simonson, R., Urea–formaldehyde resins. Part 2. The for-
mation of N,N -dimethylolurea and trimethylolurea in urea–formaldehyde
mixtures. Macromol. Mat. Eng., 72, 1, 67–74, 1978.
95. Du, G., Lei, H., Pizzi, A., Pasch, H., Synthesis–structure–performance rela-
tionship of co-condensed phenol–urea–formaldehyde resins by MALDI-
TOF and 13C NMR. J. Appl. Polym. Sci., 110, 1182–1194, 2008.
96. Pizzi, A., Pasch, H., Simon, C., Rode, K., Structure of resorcinol, phenol
and furan resins by MALDI-TOF mass spectrometry and 13C NMR. J. Appl.
Polym. Sci., 92, 2665–2674, 2004.
97. Despres, A., Pizzi, A., Vu, C., Pasch, H., Formaldehyde-free aminoresin wood
adhesives based on dimethoxyethanal. J. Appl. Polym. Sci., 110, 3908–3916, 2008.
98. Gavrilovic-Grmusa, I., Olivera, N., Điporovic-Momcilovic, M., Popovic, M.,
Molar-mass distribution of urea–formaldehyde resins of different degrees of
polymerisation by MALDI-TOF mass spectrometry. J. Serb. Chem. Soc., 75,
689–701, 2010.
99. Despres, A., Pizzi, A., Pasch, H., Kandelbauer, A., Comparative 13C NMR and
MALDI-TOF of species variation and structure maintenance during MUF
resins preparation. J. Appl. Polym. Sci., 106, 1106–1128, 2007.
100. Kandelbauer, A., Despres, A., Pizzi, A., Taudes, I., Testing by FT-IR species
variation during MUF resins preparation. J. Appl. Polym. Sci., 106, 2192–
2197, 2007.
101. Albritton, R.O. and Short, P.H., Effects of extractives from pressure-refined
hardwood fiber on the gel time of urea–formaldehyde resin. For. Prod. J., 29,
40–41, 1979.
102. Slay, J.R., Short, P.H., Wright, D.C., Catalytic effects of extractives from
pressure-refined fiber on the gel time of urea–formaldehyde resin. For.
Prod. J., 30, 22–23, 1980.
103. Johns, W.E. and Niazi, K.A., Effect of pH and buffering capacity of wood on the
gelation time of urea–formaldehyde resin. Wood Fiber Sci., 12, 255–263, 1980.
104. Xing, C., Zhang, S.Y., Deng, J., Riedl, B., Cloutier, A., Medium-density fiber-
board performance as affected by wood fiber acidity, bulk density, and size
distribution. Wood Sci. Technol., 40, 637–646, 2006.
Urea–Formaldehyde Resins 99

105. Guo, A.L., Zhang, H.S., Feng, L.Q., Gao, X.X., Zhang, G.L., pH value and
buffering capacity of 6 shrub species and relevant effect on curing time of UF
resin, Chin. Wood Ind., 12, 18–20, 1998.
106. Xing, C., Zhang, S.Y., Deng, J., Effect of wood acidity and catalyst on UF resin
gel time. Holzforschung, 58, 408–412, 2004.
107. K. Schmidt, D. Grunwald, J. Miertzsch, Phenol urea/melamine formalde-
hyde copolymers, method for the production thereof and use of the same.
US Patent Application 2006/0100412A1, 2003.
108. Park, B.-D., Kim, Y.S., Riedl, B., Effect of wood-fiber characteristics on
medium density fiberboard (MDF) performance. J. Korean Wood Sci.
Technol., 29, 27–35, 2001.
109. Medved, S. and Resnik, J., Influence of the acidity and size of beech particles
on the hardening of the urea–formaldehyde adhesive. Acta Chim. Slov., 51,
353–360, 2004.
110. Kay, R. and Westwood, A.R., DSC investigations on condensation polymers—I.
Analysis of the curing process. Eur. Polym. J., 11, 25–30, 1974.
111. Szestay, M., László-Hedvig, Z., Takács, C.E., Gasc-Baitz, E., Nagy, P., Tudos,
F., pH control of the condensation reaction and its effect on the properties of
formaldehyde/urea resin. Angew. Makromol. Chem., 215, 79–91, 1994.
112. Myers, G.E. and Koutsky, J.A., Formaldehyde liberation and curve behavior
of urea formaldehyde resins. Holzforschung, 44, 117–126, 1990.
113. Šebenik, A., Osredkar, U., Žigon, M., Vizovisek, I., Study of the reaction
between urea and formaldehyde by DSC and 13C NMR spectroscopy. Angew.
Makromol. Chem., 102, 81–85, 1982.
114. Kissinger, H.E., Reaction kinetics in differential thermal analysis. Anal.
Chem., 29, 1702–1706, 1957.
115. Conner, A.H., Polymeric Materials Encyclopedia, J.C. Salamone (Ed.),
pp. 8496–8501, CRC Press, Boca Raton, 1996.
116. F.C. Dupre, M.E. Foucht, W.P. Freese, K.D. Gabrielson, B.D. Gapud, W.H.
Ingram, T.M. McVay, R.A. Rediger, K.A. Shoemake, K.K. Tutin, J.T. Wright,
Cyclic urea–formaldehyde prepolymer for use in phenol–formaldehyde
and melamine–formaldehyde resin-based binders. US Patent 6,379,814,
2002.
117. P.G. Dopico, B.M. Peek, B.D. Gapud, K.A. Shoemake, J.C. Phillips, Cyclic
urea–formaldehyde prepolymer-modified melamine-containing resins, and
use in binders for cellulosic-based composites. US Patent 6,399,719, 2002.
118. Essawy, H.A., Moustafa, A.A.B., Elsayed, N.H., Improving the performance
of urea–formaldehyde wood adhesive system using dendritic poly(amido-
amine)s and their corresponding half generations. J. Appl. Polym. Sci., 114,
1348–1355, 2009.
119. Essawy, H.A., Moustafa, A.A.B., Elsayed, N.H., Enhancing the properties of
urea–formaldehyde wood adhesive systems using different generations of
core shell modifiers based on hydroxyl-terminated dendritic poly(amidoam-
ine)s. J. Appl. Polym. Sci., 115, 370–375, 2010.
100 Adhesives for Wood and Lignocellulosic Materials

120. Essawy, H.A. and Mohamed, H., Poly(amidoamine) dendritic structures,


bearing different end groups, as adhesion promoters for urea–formaldehyde
wood adhesive system. J. Appl. Polym. Sci., 119, 760–767, 2011.
121. Zhou, X., Essawy, H.A., Pizzi, A., Li, X., Pasch, H., Pretorius, N.O., Du, G.,
Poly(amidoamine)s dendrimers of different generations as components of
melamine urea formaldehyde (MUF) adhesives used for particleboards pro-
duction: What are the positive implications? J. Polym. Res., 20, 10, 267–280,
2013.
122. Zhou, X., Essawy, H.A., Pizzi, A., Li, X., Rode, K., Radke, W., Du, G.,
Upgrading of MUF adhesives for particleboard production using oligomers
of hyperbranched poly(amine-ester). J. Adhes. Sci. Techn., 27, 1058–1068,
2013.
123. Amirou, S., Zhang, J., Essawy, H.A., Pizzi, A., Zerizer, A., Li, X., Delmotte, L.,
Utilization of hydrophilic/hydrophobic hyperbranched poly(amidoamine)s
as additives for melamine–urea–formaldehyde adhesives. Polym. Comp., 36,
2255–2264, 2015.
124. Petersen, H., Process for the production of formaldehyde-free finishing
agents for cellulosic textiles and the use of such agents. Textilveredlung, 2,
51–62, 1968.
125. Zhang, Y.F., Zeng, X.R., Ren, B.Y., Synthesis and structural characterization
of urea-isobutyraldehyde-formaldehyde resins. J. Coatings Technol. Res., 6,
337–344, 2009.
126. Zhang, J., Chen, H., Pizzi, A., Li, Y., Gao, Q., Li, J., Characterisation and
application of urea–formaldehyde–furfural co-condensed resins as wood
adhesives. Bioresources, 9, 6267–6276, 2014.
127. E.E. Novotny and W.W. Johnson, Furfural–urea resin and process of making
the same. US Patent 1,827,824, 1931.
128. Pizzi, A. and Walton, T., Non-emulsifiable, water-based diisocyanate adhe-
sives for exterior plywood, Part 1: Novel reaction mechanisms and their
chemical evidence. Holzforschung, 46, 541–547, 1992.
129. Pizzi, A., Valenzuela, J., Westermeyer, C., Non-emulsifiables, water-based,
diisocyanate adhesives for exterior plywood, Part 2: Industrial application.
Holzforschung, 47, 69–72, 1993.
130. Despres, A., Pizzi, A., Delmotte, L., 13C NMR investigation of the reaction in
water of UF resins with blocked emulsified isocyanates. J. Appl. Polym. Sci.,
99, 589–596, 2016.
131. Japanese Standard JIS A 5908: Particleboards. Testing method, Japanese
Standards Association, 2003.
132. Japanese Standard JIS A 1460: Building boards determination of formalde-
hyde emission, Japanese Standards Association, 2001.
4
Melamine–Formaldehyde Resin

4.1 Introduction
Melamine-based resins belong to the class of aminoplastic resins and are
made by the reaction of formaldehyde either with mainly melamine, or
employing urea, phenol, or other components as co-monomers. Melamine
adhesives are predominantly used as adhesives for wood panel products
such as particleboards, medium-density fiberboard (MDF), oriented
strand board (OSB), plywood, blockboards, and others. Melamine-based
adhesives are also used for the production of paper laminates. In rare cases,
the resins and panels are also used in the furniture industry.
For most applications as wood adhesives, melamine resins are in liquid
form. For special applications, powdered (spray dried) types are used. The
resins consist of linear or branched oligomeric and polymeric molecules
in an aqueous solution, and sometimes partly as a dispersion of molecules
in an aqueous phase. The resins show a thermosetting behavior, leading to
three-dimensional hardened networks.

4.2 Chemistry
Melamine is a 1,3,5-amino substituted triazine derivative, industrially pre-
pared by cyclic condensation of three urea molecules and possessing the
following structure:

NH2

C
N N
H2N C C NH2
N
Melamine

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (101–113) © 2019
Scrivener Publishing LLC

101
102 Adhesives for Wood and Lignocellulosic Materials

The reaction between melamine and formaldehyde is similar to that of


urea.
As can be noticed in the above structure, the melamine molecule con-
tains three primary amine groups, each of which has the potential of react-
ing with two moles of formaldehyde forming up to a hexa-substituted
product if the molar ratio F/M is high enough. Because of the significantly
higher reactivity than urea of the amino groups of melamine towards sub-
stitution with formaldehyde, the melamine resins show the ability to form
polymer structures with a much higher cross-link density compared to UF
resins [1].

4.2.1 Formation of Methylolmelamine


Under slightly alkaline conditions, melamine reacts with formaldehyde to
form methylol derivatives with up to six methylol groups per molecule.
Initial reaction between the monomers (usually at temperatures in excess
of 80°C, in aqueous solution, and a slightly alkaline environment) gives
methylolmelamines. With M:F ratios of 1:2 to 1:3, dimethylolmelamine is
formed. Decreasing the M:F ratio to 1:8 will give hexamethylolmelamines,
in which all of the NH2 groups are converted to methylol. These are often
used as cross-linkers in reactions with other resins, like acrylates or polyes-
ters. The methylol derivatives so formed can be presented by the following
structures:

NH2 NH2

C 2 CH2O C
N N N N
H2N C C NH2 HOCH2-HN CC NH-CH2OH
N N
Melamine Dimethylol melamine

Similarly, depending on the ratio of formaldehyde to melamine,


trimethylolmelamine and up to hexamethylolmelamine are formed:

HO-CH2 CH2OH
N
NH-CH2OH
C
C HO-CH2 N N CH2OH
N N N
N C C
HOCH2-HN C C NH-CH2OH HO-CH2
N CH2OH
N
Trimethylol melamine Hexamethylol melamine
Melamine–Formaldehyde Resin 103

NH2 NH2

C C
N N N N
H 2N C C OH HO C C OH
N N
Ammeline Ammelide

OH

C
N N
HO C C OH
N
Cyanuric acid

Figure 4.1 Ammeline, ammelide, and cyanuric acid.

Pure MF resins are usually prepared by the reaction of melamine with


formaldehyde in an aqueous solution, yielding a precondensate consisting
of a mixture of different monomeric as well as short linear and branched
oligomeric species. These reactions are controlled by the following param-
eters: temperature, duration of condensation, pH, as well as the order and
time course of heating and reagent addition. Usually, all types of methylo-
lated melamine species together with oligomeric parts (usually more than
six melamine residues are linked from the beginning of the reaction) are
present in the reaction mixture [2].
At the condensation stage, attention must be paid to the formation
of hydrolysis products of the melamine before preparation starts. The
hydrolysis products of melamine are obtained when the amino groups of
melamine are gradually replaced by hydroxyl groups. Complete hydrolysis
produces cyanuric acid (Figure 4.1) [1].
Ammeline and ammelide can be regarded as partial amides of cyanuric
acid. They are acid and have no use in resin production. They are very
undesirable by-products of the manufacture of melamine because of their
catalytic effect in the subsequent MF resin production, due to their acidic
nature. If present, both must be removed from crude melamine by an alkali
wash and/or crystallization of the crude melamine [1].

4.2.2 Condensation of Methylolmelamines


On heating, methylolmelamines condense to form resinous products,
which become increasingly hydrophobic until eventually a gel is formed.
104 Adhesives for Wood and Lignocellulosic Materials

The rate of resinification is strongly dependent on pH. An increase or


decrease from the pH level of 10 to 10.5 of liquid resin stability results in
considerable increase in the reaction rate. The chain extension occurring
during condensation can be represented as follows:

(a) Methylol-amine condensation methylene link is formed

R-NH-CH2OH + H2N-R R-NH-CH2-NH-R + H2O

(b) Condensation between two methylol groups Ether linkage is formed

R-NH-CH2OH + HO-CH2-NH-R R-NH-CH2-O-CH2-NH-R + H2O

(c) Condensation between methylol group and imino hydrogen

R-NH-CH2
R-NH-CH2OH + HO-CH2-NH-R
R-N-CH2OH + H2O

(d) Ether link is broken with evolution of CH2O and water and formation of mehylene

R-NH-CH2OH + HO-CH2-NH-R' R-NH-CH2-NH-R + H2O + CH2O

4.2.3 Cross-Linking
Conversion of liquid melamine–formaldehyde resin to a solid network
takes place by heating. The rate of cross-linking can be increased by the
addition of an acid or a salt, but for several applications, the rate of cure is
adequate without the addition of acids. Resin curing proceeds through the
participation of methylol groups and amino hydrogen in the same manner
as shown in Section 4.2.2 above. A simplified schematic structure of the
cured melamine formaldehyde resin is given below: The structure empha-
sizes the importance that there are many ether bridges besides unreacted
methylol groups and methylene bridges [1, 3]. This is because in curing
MF resins at temperatures up to 100°C, no substantial amounts of formal-
dehyde are liberated. Only small quantities are liberated during curing up
to 150°C. However, UF resins curing under the same conditions liberate a
great deal of formaldehyde.
Melamine–Formaldehyde Resin 105

N N N
C C NH CH2 NH C C NH CH2 CH2 NH C C NH
N N N N N N
C C C

NH NH NH

CH2 CH2

O O

CH2 CH2

N CH2OH NH

C C
N N N N
N
C C NH CH2 NH C C NH CH2 O CH2 NH C C NH
N N
N N
C

NH CH2OH

CH2

NH

C
N N
NH C C
N

4.3 Melamine–Urea–Formaldehyde (MUF) Resin


MUF resins are widely used in the wood working industry as adhesives
for the production of wood-based panels, namely, particleboard, ply-
wood, and various fiber boards. These resins are the products of the reac-
tion of urea and melamine with formaldehyde. The resin is set or cured
with a hardener such as ammonium salts or acids. A urea–formaldehyde
resin with a high melamine content is described in reference [4]. The
resin contained 9.5% to 34.4% melamine by weight. The resin had a shelf
life of less than 1 day [4]. Usable formulations are also disclosed in the
scientific literature [1, 5].
US Patents 4,536,245, 5,162,462, 4,603,191, and 5,008,365 have dis-
closed significant details of condensation of MUF resins [6–9].
106 Adhesives for Wood and Lignocellulosic Materials

US Patent 5,162,462 [7] describes a method for preparing aminoplast


resin based on 2-ureido-4,6-diaminotriazine-1,3,5. The resulting resin is
reported to have very low formaldehyde emission.

NH-C-NH2

C
N N
H2N C C NH2
N
2-ureido-4,6-diaminotriazine 1,3,5

2-ureido-4,6-diaminotriazine-1,3,5 is prepared by reacting melamine


and urea in all molar ratios in dimethyl formamide, the reaction being
maintained at reflux temperature for at least 6 h and washing the resulting
solution with an alkaline solution.
US Patent No. 5,681,917 discloses a method for preparing a MUF resin
of low formaldehyde content [10]. The resin has a formaldehyde-to-urea
ratio of 0.5:1 to 1.1:1.
US Patent 6,723,825 B2 relates to a liquid MUF resin that is stable for
a period of up to 4 weeks [11]. The relevant details have been discosed in
this patent.

4.3.1 Liquid MUF Resin Preparation


The resin is prepared by taking 955 g of non-methanol-stabilized formalin
of concentration 44% in a reaction vessel. The pH is adjusted to 8.0 to 8.5
using 1.1 g of 20% sodium hydroxide solution. The contents were heated to
50°C and after about 30 min, 423 g of urea was added. The temperature was
raised to 90°C and held for 15 min. The contents were cooled to 85°C and
pH was reduced to 4.9 to 5.1 using about 1.1 g of 10% formic acid. The
viscosity was monitored using a Brookfield cone plate viscometer at 25°C
using spindle # 42 and a shear rate of 10 rpm with 1-ml samples. Once the
viscosity reached 200 to 280 cps, the kettle was cooled to 60°C and the pH
increased to 8.4 to 8.8 with about 0.75 g of 20% sodium hydroxide solution.
A second formalin (610 g) was added to the mixture, and the pH was
adjusted to 8.5 to 8.8 with about 1.1 g of 20% sodium hydroxide solution.
Melamine (567 g) was then added, followed by 60 g of methanol. The pH
was then increased to 9.8 to 10.2 with about 1.1 g of 20% sodium hydroxide
solution. The mixture was then heated to 85°C and the 25°C hydrophobic
test was conducted. The hydrophobic test was performed by removing a
Melamine–Formaldehyde Resin 107

25-g sample of resin from the mixture and titrating the resin with distilled
water at 25°C to a hydrophobic cloudy end point. The volume of distilled
water is measured and the test repeated until the range is 22 to 25 ml/25 g
of the resin. The hydrophobic test provides an indication that the desired
viscosity has almost been reached. Once the hydrophobic end point was
reached, the viscosity of the mixture was monitored until it reached a
value of 200 to 280 cps. The mixture was then cooled to 25°C, the pH was
adjusted to 9.2 to 9.5 with about 0.35 g of 20% sodium hydroxide solution,
and the resin was filtered.

4.3.2 Phenol–MUF (PMUF) Resins


Most wood adhesives used for the production of wood panel products at
present are based on urea–formaldehyde resins that do not have acceptable
weatherability. Although melamine resins are increasingly used to meet
higher requirements for bond strength and moisture resistance, they were
thought to be not completely suitable for preparing weatherproof adhe-
sive joints. Weatherproof wood joints have been produced in the past
only with phenolic resins capable of alkaline cure; however, such joints
are dark in color. Thus, although nowadays really weatherproof melamine
and melamine–urea resins have been developed, there remains a feeling
in the industry for a commercial need for adhesive resins that are clear,
homogeneous, miscible with water, of low viscosity, suitable for easy and
even spraying or blending with wood composites, and stable, and that have
acceptable storage stability and a fast curing rate. They should be consistent
in quality from batch to batch, cure quickly and evenly, and meet the low
formaldehyde emission requirement. These properties are achieved by the
disclosure in European patent EP 0915141 B1 on PMUF resins prepared
by condensing phenol, formaldehyde, melamine, and urea in the sequence
set forth in the patent [12]. Unfortunately, for this type of PMUF adhesives,
resin engineering has progressed so much for MUF resins, which are eas-
ier to produce, to render much less attractive the preparation and use of
PMUF resins [1, 13].
According to PMUF preparation European Patent EP 0915141 B1, it
is imperative to condense formaldehyde with phenol and then condense
melamine with the phenol–formaldehyde condensate before introducing
urea [12]. If this condensing sequence is not observed, a resin will not be
obtained, which can be used to make adhesives with weatherproof quality.
The details of the process are disclosed in the above patent.
In preparing the condensate, formaldehyde and phenol are added to a
reaction vessel and thoroughly mixed. This will take the reaction between
108 Adhesives for Wood and Lignocellulosic Materials

phenol and formaldehyde to the methylolation state. Sufficient triethanol-


amine is added to raise the pH to 5.6–6.4, preferably 5.8–6.2, and then, with
caustic soda, the pH is increased to 8.5–9.5, preferably about 9.0. Cooling
of the reactor contents is commenced and when a temperature of 40°C or
below is reached, melamine is added with vigorous agitation to ensure that
all of the melamine is wetted.
At this point, the contents of the reactor are heated to 85°–100°C, pref-
erably 93°–95°C, and held at this temperature for 10–45 min, preferably
about 20 min. Condensation of the melamine is considered complete when
the solution becomes clear.
This is followed by the addition of the first portion of urea over a period
of 5–20 min, preferably 7–13 min, while holding the temperature at about
93°–95°C to obtain a polymer of desired chain length. After all of the urea
solubilizes, the reaction is held at 80°–100°C, preferably 87°–89°C, until a
water tolerance of 70–180%, preferably 100–150%, is reached.
The contents of the reactor at this stage are cooled to 45°–50°C and an
additional 0.2–1 mol of urea, preferably 0.65–0.75 mol, is added to react
with excess formaldehyde, with continued cooling to about 25°C. Lastly,
the pH is adjusted with triethanolamine to 7.5–8.5, preferably 8.0–8.4.
Frequently, in these resins, the phenol is not completely reacted. Thus,
the PMUF resin will be a poorer resin than just an equivalent MUF resin.
Reaction of the phenol with the MUF part depends exclusively on the con-
ditions under which the reaction is carried out. If the phenol addition is
done under the incorrect conditions, which is often the case, the phenol
may remain often unutilized and does not contribute to the final resin net-
work [1, 13].
It is then necessary to define the order how the reagents are added to
ensure that the phenol also participate to the strength and characteristics
of the hardened network [1, 13]. Good PMUF resins are produced today
by very reputable companies, but they do not appear to be better than
MUF resins where the phenol has been substituted mole by mole by just
melamine.

4.3.3 Melamine–Formaldehyde Resin Modification


by Acetoguanamine for Post-Formable High-Pressure
Laminate
Compared to the laminates of the usual commercial quality, post-moldable
laminates have the property of being still moldable after heating up. Using
such moldable sheets, it is possible to manufacture furniture compo-
nents with decorative melamine resin surfaces with rounded edges. Main
Melamine–Formaldehyde Resin 109

applications are for kitchen furniture. The advantages of such sheets are
obvious. By having seamless cover of the edges, penetration of water is
prevented. Several methods were used, e.g., during the hot pressing, the
resin does not cure completely. These sub-cured sheets could be post-
molded well. However, during storage, post-curing takes place and there-
fore the post-moldability is negatively affected or even lost. Unsuccessful
attempts were made to produce suitable resins by addition of plasticiz-
ers that will not react with formaldehyde or the methylol groups of the
melamine resin, e.g., PVC, polyacrylonitrile, and polyvinyl acetate. These,
called external plasticizers, have the defect to migrate to the surface, caus-
ing several technical problems. Subsequently, internal plasticizers, thus
plasticizers reacting and being co-condensed within the resin, were tried.
These modifying agents are built into the structures of the melamine
resin molecule. Acetoguanamine belongs to this category and is the most
effective for this purpose, but also others are sometimes used such as
ε-caprolactam.
CH3

N N

H2N N NH2

Acetoguanamine

Kuchler et al. [14] found that acetoguanamine is particularly suitable


as the modifying agent for melamine formaldehyde resin to improve the
post-moldability of high-pressure laminates made therefrom. Another
modifying agent used in this study was diethylene glycol or digly-
col. Seeholzer reported that acetoguanamine improved the elasticity of
melamine resins [15]. By condensing acetoguanamine together with
melamine resin, the resin remains plastic and formable over prolonged
periods of time. Seeholzer further found that diethylene glycol had a syn-
ergistic effect with the acetoguanamine and reduced cured cracking [15].
Paper laminates impregnated by melamine-based resins can today be
prepared in two different ways [1]: (1) either by double impregnating bath
machines in which the paper passes first in a cheaper UF resin solution
and then in a second bath of a more expensive MF resin, thus producing a
paper core impregnated of UF resin and surfaced with MF resin; (2) or in
single bath machines where either a traditional MF resin is impregnated,
110 Adhesives for Wood and Lignocellulosic Materials

this being more expensive, or a cheaper MUF resin of a relatively high


level of melamine. Both approaches are aimed to decrease the proportion
of expensive melamine without altering the impregnated paper approach.

4.3.4 MUF Adhesive Resins of Upgraded Performance


Decreasing the proportion of melamine in MUF resins at equal perfor-
mance has been a recent trend. This can be done by a number of different
approaches. These are as follows: (1) the use of melamine salts as additives
to UF resins, also performing, in this approach, the function of hardeners.
The proportion of melamine can then be decreased to 10% for an effect
normally obtained by 30% of melamine in the resin [16–19]. The mech-
anism by which addition of melamine salts to a UF adhesive is so much
more effective has been determined [16, 20]. Exposure to the weather of
panels prepared with this adhesive approach has confirmed the laboratory
results concerning the performance of these panel adhesives [18, 21].
(2) The progress in melamine and urea adhesive formulations has
caused that the postulate that degradation induced in the curing of these
resins by an excessively long panel press time at high temperature occurs
during board manufacture is no longer valid. The lower resin molar ratios
used today to counterbalance and decrease formaldehyde emission has led
to much slower curing resins, thus rendering questionable the industrial
practice of rapidly cooling the panels once out of the hot press to avoid the
degradation of the adhesive hardened network [22, 23]. This is so, as the
adhesives holding the boards together are not completely cured once com-
ing out of the hot press, with an estimate of around 70% only of possible
maximum cure. This means that while some cooling is still necessary once
the panels come out of the hot press, this can be reduced to just maintain
the panels at a hotter temperature than what is done in the past and thus
obtain a better strength at a shorter press time. Models defining the more
apt post-curing strategies have been determined [22].
(3) The sequence of manufacture of these MUF resins determines that
the chemical species produced are very different, and at minimum, their
relative proportions are also very different, causing also marked differences
in the cross-linked network induced by their hardening. For example, one
can prepare MUF resins according to the classical, traditional, sequential
addition of reagents, but also MUF resins according to nontraditional,
nonsequential formulation approaches. These two approaches have been
shown to produce very different resins [24]. Nonsequential MUF resin
formulations have only relatively recently been started to be investigated
for their potential in markedly decreasing the proportion of the expensive
Melamine–Formaldehyde Resin 111

melamine while maintaining resin performance. Such approaches are


based on additives having special effects. The main additives, used in 1% to
5% on resin solids, for this approach are the so-called acid anion-stabilized
iminomethylene basis [25–28] that can be prepared in a couple of different
ways such as reacting formaldehyde and ammonia [25–28] or by stabiliz-
ing the decomposition of hexamine to prepare the so-called “hexamine
sulfate” [29]. These approaches do achieve marked decreases in melamine
content at equal resin performance. For example, the strong effect on MUF
adhesives of the so-called “hexamine sulfate” was found to be due to the
strong buffering and stabilizing effect induced by the additive during adhe-
sive curing [26–28].

4.3.5 Cold-Setting MUF Adhesives


MUF cold-setting wood adhesives for glulam and fingerjointing are also
used especially in Europe. They are considered for use as protected exterior
structural applications and their clear bondline renders them particularly
acceptable from a visual, aesthetic point of view. The positive evolution
in their formulation and the forever-improving performance render them
today comparable to classical PRF adhesives [30], these latter being still
preferred in North America while the former are preferred in Europe.

References
1. Pizzi, A., Melamine–formaldehyde adhesives, in: Handbook of Adhesive
Technology, 2nd edn, A. Pizzi and K.L. Mittal (Eds.), pp. 653–680, Marcel
Dekker, New York, 2003.
2. Melamine–formaldehyde resins, Encyclopedia of Polymer Science and
Technology, vol. 10, H. Mark and N.M. Bikales (Eds.), pp. 369–378, John
Wiley & Sons, New York, 1965.
3. Saunders, K.J., Organic Polymer Chemistry, Chapman and Hall, London, 1973.
4. Melamine modified urea–formaldehyde resin for bonding flakeboards,
Proceedings of Symposium on Wood Adhesives, pp. 155–159, Forest Products
Research Society, Madison, Wisconsin, 1990.
5. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York,
1994.
6. D.W. Shiau and E. Smith, Low formaldehyde emission urea–formaldehyde
resins containing a melamine additive. US Patent 4,536,245, 1985.
7. J. Barthomieux, R. Garrigue, J. Lalo, Process for the preparation of amino-
plastic resin for use in the production of particle boards with low formalde-
hyde emission. US Patent 5,162,462, 1992.
112 Adhesives for Wood and Lignocellulosic Materials

8. H.C. Kong, Process for preparing a urea–formaldehyde resin having a very


low mole ratio of formaldehyde to urea. US Patent 4,603,191, 1986.
9. B. Druet and G. Rochet, Process for the manufacture of urea–formaldehyde
resins containing melamine. US Patent 5,008,365, 1991.
10. R.A. Breyer, S.G. Hollis, J.J. Jural, Low mole ratio melamine–urea–formaldehyde
resin. US Patent 5,681,917, 1997.
11. H.D. Zhu, W. Jacobs, R.G. Lees, L. Mendonca, R.C. Rasch, Stable liquid
melamine urea formaldehyde resins, hardeners, adhesive compositions, and
methods for making same. US Patent 6,723,825 B2, 2004.
12. M. Paventi, Phenol–melamine–urea–formaldehyde copolymer resin compo-
sitions, method of making and curing catalysts thereof. European Patent EP
0915141 B1, 2003.
13. Cremonini, C., Pizzi, A., Tekely, P., Influence of PMUF resins preparation
method on their molecular structure and performance as adhesives for ply-
wood. Holz Roh Werkst., 54, 85–88, 1996.
14. J. Kuchler, H. Erben, J. Seeholzer, Modifying agents for thermosetting resins
and thermosetting resins prepared using these agents. US Patent 5,346,937,
1994.
15. Seeholzer, J., Manufacturing of post-formable decorative laminates using
acetoguanamine. Kunststoffe, 69, 5, 263–265, 1979.
16. Prestifilippo, M., Pizzi, A., Norback, H., Lavisci, P., Low addition of melamine
salts for improved UF adhesives water resistance. Holz Roh Werkst., 54, 393–
398, 1996.
17. Cremonini, C. and Pizzi, A., Improved waterproofing of UF plywood adhe-
sives by melamine salts as glue-mix hardeners. Holzforsch Holzverwert., 49,
1, 11–15, 1997.
18. Cremonini, C. and Pizzi, A., Field weathering of plywood panels bonded
with UF adhesives and low proportion of melamine salts. Holz Roh Werkst.,
57, 318, 1999.
19. Kamoun, C. and Pizzi, A., Performance effectiveness of addition to UF of
melamine salts vs. melamine alone in MUF adhesives for plywood. Holz Roh
Werkst., 56, 86, 1998.
20. Pizzi, A. and Panamgama, L.A., Diffusion hindrance vs. wood-induced cata-
lytic activation of MUF adhesives polycondensation. J. Appl. Polym. Sci., 58,
109–115, 1995.
21. Pizzi, A., High performance MUF resins of low melamine content by a num-
ber of novel technique, in Proceedings, Wood Adhesives 2000, Forest Products
Society, Madison, Wisconsin, 2000.
22. Lu, X. and Pizzi, A., Curing conditions effects on the characteristics of ther-
mosetting adhesives-bonded wood joints—Part 2: Hot postcuring improve-
ment of UF particleboards and its temperature forecasting model. Holz Roh
Werkst., 56, 393–401, 1998.
Melamine–Formaldehyde Resin 113

23. Zhao, C. and Pizzi, A., Hot postcuring improvement of MUF-bonded par-
ticleboards and its temperature forecasting model. Holz Roh Werkst., 58,
307–308, 2000.
24. Zanetti, M. and Pizzi, A., Dependance on the adhesive formulation of the
upgrading of MUF particleboard adhesives and decrease of melamine con-
tent by buffer and additives. Holz Roh Werkst., 62, 451–455, 2004.
25. Pichelin, F., Kamoun, C., Pizzi, A., Hexamine hardener behaviour—Effects
on wood glueing, tannin and other wood adhesives. Holz Roh Werkst., 57,
305–317, 1999.
26. Zanetti, M. and Pizzi, A., Low addition of melamine salts for improved MUF
adhesives water resistance. J. Appl. Polym. Sci., 88, 287–292, 2003.
27. Zanetti, M. and Pizzi, A., Upgrading of MUF resins by buffering additives—
Part 2: Hexamine sulphate mechanisms and alternate buffers. J. Appl. Polym.
Sci., 90, 215–226, 2003.
28. Zanetti, M., Pizzi, A., Kamoun, C., Upgrading of MUF particleboard adhe-
sives and decrease of melamine content by buffer and additives. Holz Roh
Werkst., 61, 55–65, 2003.
29. Mouratidis, P., Dessipri, E., Pizzi, A., New adhesive system for improved
exterior-grade wood panels, in Proceedings, Wood Adhesives 2000, European
Union Final Contract Report, FAIR TC 96-01604. Tahoe, Nevada, 2000.
30. Properzi, M., Pizzi, A., Uzielli, L., Honeymoon MUF adhesives for exterior
grade glulam. Holz Roh Werkst., 59, 413–421, 2001.
5
Phenol–Formaldehyde Resins

5.1 Introduction
The extensive use of phenolic resin as adhesive for bonding wood and wood
panel products is due to such properties as their heat resistance, water resis-
tance, and the mechanical properties of the cured phenolic resins. Phenolic
resins can be prepared by two types of chemistry. The first, termed novolacs,
are phenolic resins that are prepared under acid catalysis with a molar excess
of phenol over formaldehyde. The polymerization, or cure, of the novolac
resins requires the addition of further formaldehyde or formaldehyde-
releasing agents such as hexamine or paraformaldehyde. The second type of
phenolic resins, known as resols, are resins prepared by employing a molar
excess of formaldehyde over phenol under alkaline conditions. In contrast
to novolac resins, there is no need to add additional formaldehyde to effect
the curing since formaldehyde is already in excess. Resols are cured either by
heating or by adding acid catalysts such as p-toluene sulfonic acid.

5.2 Historical
First, concrete experiments on the condensation of phenol–formaldehyde
(PF) condensation was conducted by Adolph von Baeyer in 1872. The first
synthetic resin “Laccain” as substitute for shellac was invented by C.H.
Meyer, Zwickau (Louis Blumer Co.) [1, 2].
In the early 1900s, Baekeland first disclosed the successful synthesis
of polymers, which he was able to commercialize initially as a resol and
shortly thereafter as a novolac [1–3].
The first patent on phenolic resins was filed by Baekeland in 1907, which
disclosed the process of curing of phenolic resins under “heat and pressure”
to produce the first member of thermosetting plastics ever known [3]. This
patent was followed by subsequent patents, namely, “base patent”, “varnish
patent”, and “grinding wheel patent”. Based on the above inventions, the

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (115–146) © 2019
Scrivener Publishing LLC

115
116 Adhesives for Wood and Lignocellulosic Materials

“Bakelite Company” was founded in Berlin in 1910 for the production of


thermosetting molding compounds and industrial (Bakelite) resins. The
abrasive industry started using the resol and novolac resins for the manu-
facture of coated and bonded abrasives, respectively [1, 2].

5.3 Definitions and Types of Phenolic Resins


PF resins are synthetic polymers obtained by the reaction of phenol or
substituted phenol with formaldehyde [4].

5.4 Basic Chemistry


Phenol and formaldehyde undergo step growth polymerization reaction
in the presence of acid or basic catalysts. Basic catalyzed reaction leads to
resol while the acid-catalyzed process leads to novolac.

5.4.1 Resols
As mentioned above, resols are synthetic resins produced from phenol
and formaldehyde under alkaline conditions with a formaldehyde-to-
phenol ratio of greater than 1 (usually around 1.5).

5.4.2 Novolacs
Novolacs are PF resins produced with phenol in molar excess over form-
aldehyde using acid catalysis. Oxalic acid, hydrochloric acid, or sulfonic
acids are normally used. Under both acidic or basic conditions, phe-
nol alcohols are first formed, which subsequently condense to produce
oligomers.
The formation of resol and novolac is summarized in Figure 5.1.

5.4.3 Difference between the Acid and Base Catalysis


The difference between the acid- and the base-catalyzed reactions lies in
the rates at which the addition and condensation reactions occur. In the
case of acid-catalyzed reaction, the rate of phenol alcohol formation is rel-
atively slow while the condensation of phenol alcohols and phenol leading
to dihydroxy-diphenyl methane derivatives is rapid. Hence, the formation
of phenol alcohol is the rate-controlling step.
Phenol–Formaldehyde Resins 117

Phenol + Formaldehyde (Excess) Phenol (Excess)s + Formaldehyde


+ Alkali + Acid

Resol Novolac

Heat/acid
Heat and Paraformaldehyde
or Hexamethylenetetramine
Resite

Figure 5.1 Resol, resitol, and resite.

Novolacs contain no free methylol groups and therefore cannot undergo


cross-linking by itself unless an additional cross-linking agent such as
hexamethylene tetramine or paraformaldehyde is added. Addition of the
cross-linking agents and heating, however, can lead to the formation of a
three-dimensional network structure through the formation of methylene
bridges.
In the case of base-catalyzed reaction between phenol and formalde-
hyde, the initial addition reaction leads to methylolphenols. This is the
reaction of Lederer–Manasse, which is an organic aromatic hydroxyalkyla-
tion reaction. It allows the introduction of a -CH2OH group on a phenolic
aromatic ring. It was independently published in the same period (1894)
by Manasse [5] and Lederer [6] studying the basic induction reaction of
formaldehyde with phenol. Baekeland was aware of the Lederer–Manasse
reaction during his initial experiments [7].
Methylolation reaction is faster than the subsequent condensation reac-
tion. The methylolphenols are stable under alkaline conditions and hence
predominate in the reaction product initially. The methyol groups present
in the phenol alcohols can either condense with methylol groups of other
phenol alcohols to form ether linkages or combine with the hydrogen atoms
of the unsubstituted ortho or para positions of methylolphenols [8, 9].

5.4.4 Reaction between Phenol and Formaldehyde


(Sodium Hydroxide Catalyzed)
5.4.4.1 Electron Delocalization in Phenol and Phenoxide Anion
The presence of hydroxyl group in the phenol molecule promotes the delo-
calization of electrons to enrich the electron densities at the 2, 4, and 6
positions in accordance with the resonance-stabilized structures given in
118 Adhesives for Wood and Lignocellulosic Materials

+ + +
: OH : OH : OH : OH
: : : :
: :

(5.1)

:
Figure 5.2 Resonance structures of phenol.

Figure  5.2. This opens up these positions for possible attack by an electro-
phile (-+CH2O-), through the electrophilic aromatic substitution process.
This phenomenon gives rise to the increase in the electron densities at
the o- and p- positions.
In the case of phenoxide anion, which is stabilized by resonance, the
following resonance structures give rise to still higher electron densities at
the o- and p- positions.
_
:O :O :O :O
: : : :
:
:
.. (5.2)

Hence, the ortho and para positions are the locations for attack by
electrophilic species.

5.4.4.2 Hydroxymethylation of Phenol and Further Condensation


(under Alkaline Conditions)
The ratio of formaldehyde to phenol is greater than 1 (usually 1.5) in the
case of base-catalyzed resol production.

1. Addition of hydroxymethyl groups to the ortho and para


free positions of phenol
(a) Mechanism based on polarized formaldehyde
Now, considering the structure of formaldehyde, it can
polarize as shown below:
+ −
CH2=O CH2 O
(5.3)
+
The electrophile CH2 — O will attack the phenoxide anion
as shown in Figure 5.3 to form the o-methylolphenol (I):
Phenol–Formaldehyde Resins 119

O O
: : H
_ Proton transfer
: + _
CH2O
+ CH2 O

OH
_
_
:O :O
OH :
: :
CH2OH CH2OH +
(5.4)

Figure 5.3 Formation of ortho-methylolphenol.

OH OH

CH2OH HO-CH2 CH2OH

OH I IV
OH

HO-CH2 CH2OH

OH OH
CH2OH
CH2OH V

CH2OH CH2OH
II III

Figure 5.4 Mononuclear phenol alcohols.

Similarly, p-methylolphenol (II), 2,4 dimethylolphenol


(III), 2,6 dimethylolphenol (IV), and 2,4,6 trimethylolphe-
nol (V) are formed as shown in Figure 5.4.
Although the para site is considered to be more reactive
than the ortho site, because of the presence of two ortho sites
for one para site, the ortho substitution predominates and
proceeds at a faster rate than para [10–12].
(b) Mechanism based on formaldehyde as methylene glycol
Aqueous formaldehyde exists in water as methylene gly-
col, the hydrated form of formaldehyde. The mechanism is
depicted as follows:
120 Adhesives for Wood and Lignocellulosic Materials

CH2 = O + H2O HO-CH2-OH (5.6)


OH O
O H
Na
H C H CH2OH + Na + OH
+
OH
(5.7)
O Na
O
H
CH2OH + H
- CH2OH (5.8)
-

From Equations 5.7 and 5.8,

H + OH H2O (5.9)
Figure 5.5 Mechanism of methylolphenol formation in alkaline medium.

The formed methylolphenols in Figure 5.5 are stable


under alkaline conditions and are the first intermediate
compounds under basic conditions. They are more stable
under these conditions since their formation is faster
than the subsequent condensation process described
below. Methylolphenols can be isolated [13]. Consequently,
phenol alcohols are initially predominant intermediate
compounds [8]. The o- or p-methylolphenols are more
reactive than the original phenol and therefore rapidly
undergo further methylolation to form di- and trimethy-
lolphenols [9].
A series of experimental results indicate that the ortho/
para substitution ratio depends on the type of catalysts.
The ortho/para ratio decreases from 1.1 at pH 8.7 to 0.38
at pH 13.0. The isomeric compositions of methylolphenols
change with different catalysts and mechanisms [11, 14]. For
instance, when metal hydroxides are utilized, a chelate ring
mechanism, at first proposed by Caesar and Sachanen [15],
favors ortho substitution [11,15–17].
Details of how ortho substitutions are influenced by
different metal hydroxides and the chelating strength
of the cations have been discussed by several authors
[18–21].
Phenol–Formaldehyde Resins 121

2. Further condensation and chain extension


The methylolphenols formed as described above can undergo
self-condensation to form dinuclear and polynuclear oligo-
mers. Two different types of self-condensation are possible:
(a) Involving two methylol groups
(b) Between methylol groups of one molecule and the
hydrogen atom of the unsubstituted o- or p- positions of the
next molecule. Examples are provided by the investigation of
Yeddanapalli and Francis [22]. When o- and p-methylophenol
are heated with aqueous sodium hydroxide at 80°C and the
products are analyzed by paper chromatography, it was clear
that the reaction between a methylol group and the p-hydrogen
atom as well as the reaction between the methylol group and
the o-hydrogen atom take place.
p-Methylolphenol gave 5-methylol-2,4’ dihydroxydiphe-
nyl methane (VIII) by reaction between methylol group and
the o-hydrogen atom and 4,4’ dihydroxydiphenyl methane by
reaction between two methylol groups [22].
The possible mechanism of the two types of condensation
(a) Between a methylol group of one molecule and the
hydrogen atom of unsubstituted o- or p- position of another
molecule (see Figures 5.6 and 5.7)
(b) Involving two methylol groups
Generally p-hydrogen atoms and p-methylol groups are
the cross-linking centers [9].
It is generally agreed that hydroxymethyl phenols con-
dense through quinone methide (QM) intermediates [23].
Quinone methides are formed from hydroxymethyl phenol

OH
OH
CH2OH
2 CH2 CH2OH
OH
VI (5.10)
OH
OH OH
CH2OH
2 CH2 CH2OH

VII (5.11)
Figure 5.6 Chain extension.
122 Adhesives for Wood and Lignocellulosic Materials

OH
OH

CH2
HO

CH2OH
CH2OH VIII

OH

CH2
HO OH

IX
CH2OH (5.12)

Figure 5.7 A dinuclear structure.

as depicted in Figure 5.8. They are strong electrophiles and


will readily substitute at the electron-rich sites of phenoxides
to form methylene bridges as follows:

O-
O=

=CH 2
CH2 OH OH
OH CH2OH
Quinone methide

OH OH OH
CH2 OH
CH2
CH2OH

Figure 5.8 Formation of quinine–methide as intermediate.

The above series of reactions yielding dinuclear phenol


alcohols may further lead to a chain extension to give trinu-
clear, tetranuclear, and so on. Hence, an industrial resol
resin is a complex mixture of mononuclear, dinuclear, and
polynuclear phenol alcohols together with unreacted phe-
nol. This mixture can be represented by a general structure
as shown in Figure 5.9.
n = 2 for typical commercial liquid resols and n = 3–4 for
solid resols.
Phenol–Formaldehyde Resins 123

OH
OH
CH3 OH OH
CH2

n
(CH2OH)0-3 (CH2OH)0-2 (CH2OH)0,1 (CH2OH)0-2
Unreacted phenol and Oigomer containing
Multifunctional methylol groups for
Monomers cross-linking

Figure 5.9 Typical structure of a commercial phenolic resol.

5.4.5 Formation of Chelate Ring


Different mechanisms of alkaline catalysis have been suggested according
to the alkali used. When caustic soda is used as the catalyst, Caesar and
Sachanen suggested the formation of a chelate ring involving the alcohol,
phenolic hydroxyl group, and the sodium in the PF alkali condensations [15].
Na
O O

CH2

The chelating mechanism was thought to initially cause the formation of


a sodium–formaldehyde complex or of a formaldehyde–sodium phenate
complex and is similar in concept to the mechanisms advanced for metal-
ion catalysis of phenolic resins in the pH range 3 to 7. However, while
the cyclic metallic ion catalysis ring complexes have even been isolated
[20, 21], this is not the case for the sodium ring complex, evidence for its
existence being rather controversial, the predominant indication being
that it may not form [8, 24].

5.4.6 Reaction between Phenol and Formaldehyde


(Ammonia and Amine Catalysis)
Ammonia-catalyzed PF resin is of great interest for producing electrical
grade wood laminates/composites since the resin is free from any conduct-
ing electrolytes. The compreg produced from ammonia-catalyzed phe-
nolic resin has very high dielectric strength, tracking resistance, and arc
124 Adhesives for Wood and Lignocellulosic Materials

OH OH

CH2-NH-CH2 HO CH2 NH CH2 OH

OH CH2NH2 OH OH

CH2-O CH2-N-CH2

CH2

OH

Figure 5.10 Structure of ammonia-catalyzed resol.

resistance. These composites are used for electrical applications such as


transformers, platforms, rings, studs, and rods.
The characteristics of ammonia-catalyzed phenolic resins are dif-
ferent from those of alkali-catalyzed resins. They are insoluble in water.
The mechanism of the reaction involved in the ammonia-catalyzed PF
resin differs from that of alkali-catalyzed PF resin [13] in that nitrogen-
containing intermediates are formed in the case of the former [25–31].
In the case of ammonia, the main intermediates are dihydroxybenzyl-
amines and trihydroxybenzylamines. Such benzylamine bridges have been
shown to be much more temperature stable to confer unique characteris-
tics to the resin. Such nitrogen-containing intermediates formed are shown
in Figure 5.10 [32].
Ammonia-, ammine-, and amide-catalyzed phenolic resins are char-
acterized by greater insolubility in water than that of sodium hydroxide-
catalyzed phenolic resins. The more ammonia that is used, the higher the
molecular weight (MW) without cross-linking. This could probably be
due to the inhibiting effect of the nitrogen-carrying groups (i.e., –CH2-
NH-CH3 or –CH2-NH2). These have a higher melting point. These resins
are soluble in alcohol [32].

5.4.7 Manufacture of Phenolic Resins


5.4.7.1 Principles of Manufacture
The following facts are to be kept in mind in order to understand the basic
principles and implement the same in the manufacturing procedure of
phenolic resols:
Phenol–Formaldehyde Resins 125

1. Thee alkali-catalyzed PF reaction is highly exothermic


and hence reaction temperature can increase at any time
to uncontrollable limits even without any external heating
2. The final product is a thermoset and hence the gelation in
the reaction kettle has to be avoided.
3. The heat of polymerization reaction increases with increase
of temperature.
4. Rate of polymerization reaction increases with temperature.
5. Because of the above reasons, it is imperative that the reac-
tion is controlled efficiently to prevent temperature runaway
of the reaction and consequent gelation inside the reaction
vessel.

If the gelation occurs in the reaction vessel, one has to wait for several
hours, depending on the size of the reactor for the contents (set solid) to
cool, and ready to be removed the from the reactor in order to continue the
production operation.
A typical phenolic resin is made in batches, in a jacketed, stainless steel
reactor equipped with an anchor-type or turbine-blade agitator, a reflux
condenser, vacuum equipment, and heating and cooling facilities. Molten
phenol and formalin (containing 37% to 42% formaldehyde or parafor-
maldehyde), in molar proportions between 1:1.1 and 1:2, along with water,
and methanol are charged into the reactor and mechanical stirring is
begun. To make a resol-type resin (such as those used in wood adhesives
manufacture), an alkaline catalyst such as sodium hydroxide is added to
the batch, which is then heated to 80°C to 100°C. Reaction temperatures
are kept under 95°C to 100°C by applying vacuum to the reactor or by
cooling water in the reactor jacket. Reaction times vary between 1 and 8 h
according to the pH, the phenol/formaldehyde ratio, the presence or
absence of reaction retarders (such as alcohols), and the temperature of
the reaction [8].
The quantity of exothermic heat generated in the PF reaction depends
on the following factors:

1. Ratio of phenol to formaldehyde


2. Concentration of alkali
3. Rate of heating of the reactants

One has to be extremely careful when vacuum is applied when the


reaction temperature is between 95°C and 100°C. In particular, if the
rate of heating of the reactants is very high, rate of heat generation due to
126 Adhesives for Wood and Lignocellulosic Materials

exothermicity is also very high. Under this circumstance, if the vacuum


is applied very fast, there can be intense agitation of the reaction mixture
and consequently the rate of polymerization reaction will be very high due
to high diffusion rate of the reactants. Sudden increase in the rate of heat
generation and increase of temperature may result in a thermal runaway,
which can lead to instant gelation of the resin.
It is possible to control the formulations and properties of phenolic
resins in a wide range of process variables. The following factors may be
mentioned:

1. Ratio of formaldehyde to phenol


2. Percentage of alkali
3. Degree of polymerisation prior to hardening

The reaction between phenol and formaldehyde under alkaline con-


ditions is highly exothermic, and if proper control of the reaction is not
exercised, the reaction will lead to thermal runaway, resulting in complete
solidification of the resin. Control of exothermicity is therefore vital.
Exothermicity can be controlled by a predictive fuzzy logic control sys-
tem with an adaptive loop for the manufacture of phenolic resin [33].
Fuzzy logic control (FLC) methodology was applied successfully for
the control of the reaction in tracking and regulating temperature profiles
under different operating conditions. An important advantage of apply-
ing FLC is that it does not require a mathematical model and the control
is based on human heuristic decisions. A laboratory-based reaction pro-
cess system and the required computer interface circuits were constructed
and the FLC software were developed and tested for the suitability in con-
trolling the temperature. Further, the above system was found to be an
ideal “Reaction Calorimeter” for the determination of heat of polymeriza-
tion under different operating conditions [34, 35].
Based on response surface methodology, experiments were conducted
by Kumar et al. [35] and the results on the heat of polymerization were
fitted to a second-degree polynomial containing the following process
variables:

1. Ratio of phenol to formaldehyde


2. Concentration of alkali
3. Rate of heating of the reactants

The heat of reaction under the above conditions was represented in a


three-dimensional response surface given in Figure 5.11 [35].
Phenol–Formaldehyde Resins 127

Estimated
Response
Function

1.5

1.2
Exothem × 10–5 Joules

0.9

0.6

2
0.3

ol) n
1

en tio
Ph tra
on cen
0

(% Con
–2 –1
–1 ali
Alk
Formaldehy
de to Phenol 1 2 –2
Ratio

Figure 5.11 Three-dimensional response surface relating the heat of P–F condensation
reaction under alkaline conditions to the F/P ratio and alkali concentration.

A typical phenolic resin is made in stainless steel batch reactor equipped


with an anchor-type or turbine-blade agitator, a reflux condenser, vacuum
equipment, and heating and cooling facilities.
Figure 5.12 shows the interaction of resin synthesis, structures, and
property relationships [36]. The resin synthesis variables are closely related
to the resultant resin structure parameters. The identification of variables
that are correlated to resin performance is the prime objective of resin syn-
thesis and characterization studies.
A typical phenolic resin is made in batches in a jacketed, stainless steel
reactor, which is equipped with an anchor-type or turbine-blade agitator,
a reflux condenser, vacuum equipment, and heating and cooling facili-
ties. In the formulation of PF resin, various resin synthesis variables such
as formaldehyde/phenol (FP) molar ratio, NaOH/phenol molar ratio,
viscosity, and reaction parameters such as pH, time, and temperature
were identified by many investigators [36]. All these variables influenced
the resin curing behavior and performance characteristics of manufac-
tured board.
Prior to the 1940s, PF resins for plywood were prepared with an F/P
molar ratio of 1.5 with a small amount of alkali. The finished resin had
128 Adhesives for Wood and Lignocellulosic Materials

Variables related to
Resin synthesis end use properties
Variable

- Cure rheology
- F/P Ratio
- Thermal properties
- NaOH/P Ratio (DSC)

- Types of catalyst - Thermo-mechanical


Properties (DMT)
- Viscosity

Resin Chemical
Structure

- Molecular weight and distribution

- Degree of branching

Figure 5.12 Interaction of resin synthesis, structure, and property relationships [36].

much formaldehyde and was slow to cure. Typical hot-press time was
15 min for 12.7-mm-thick plywood [37, 38].
In the early 1940s, van Epps [39] adopted a procedure by which more
alkali was added at the beginning of resin synthesis with a solids content
of 55–58%.
Hot-press time for the 12.7-mm plywood dropped to 7 min. Thus, the
resin doubled the productivity of plywood, but associated with a strong
formaldehyde odor, which may have come from the free formalde-
hyde present in the resin. In this method, formaldehyde was lost by the
Cannizzaro reaction because of the initial high alkalinity:

2 CH 2O OH CH3OH HCOO
40 60 C

The Cannizzaro reaction is usually considered to be an undesirable


pathway by which 2 mol of formaldehyde are replaced with 1 mol each
Phenol–Formaldehyde Resins 129

of methanol and formic acid. This reaction is favored at very high NaOH
concentrations.
This side reaction is considered most important because 3–5% formal-
dehyde loss is attributed to it [40].
Condensation reaction rate constants are about 1 × 10−5, but the rate
constant of the formaldehyde Cannizzaro reaction is 1.25 × 10−3 at 50°C
in NaOH. Thus, measuring the rate of condensation reaction by the loss of
formaldehyde gives erroneously high rates [41].
Another approach by Stephan et al. used a reduced initial NaOH to
suppress losses of formaldehyde [42]. The reaction was kept at reflux for
90–120 min until the desired viscosity was obtained. After cooling, a sec-
ond NaOH was added to the resin to lower viscosity. Hot-press time for a
1/2-inch panel was 6 min with the prepared resins that were relatively low
MW and had a solids content of 50–53%.
Redfern patents marked the beginning of the current technology in PF
resin manufacture [43, 44]. These methods consisted of a stepwise addi-
tion of the NaOH to get the viscosity as high as possible to still handle.
Each additional NaOH addition served to lower the viscosity and further
catalyze the condensation reaction. The resin had a solids content range of
40–44% and was cooked longer for a high-molecular-weight resin. These
resins are fast curing, giving 5 min of hot-pressing time for a 1/2-inch
panel. The drawback is that these resins with high MWs will not melt and
flow without water in the glueline. Lambuth used more NaOH in the first
addition than Redfem, and heated gently to reflux [45]. This condition
favored the formation of highly methylolated resins, but took more time
to prepare. Modern resins are a combination of Redfern and Lambuth
technology.

5.5 Effect of Process Variables


5.5.1 Catalyst Types and pH of Resin
As discussed earlier, the properties of PF resins can be modified using var-
ious catalysts such as sodium hydroxide, ammonia, amines, and amides
[24]. The possibility of involvement of the phenolic hydroxyl group in the
formation of methylene linkages in the PF resins was postulated by Caesar
and Sachanen [15].
The hypothesis is based on the fact that it was impossible for thio-
phene to react with formaldehyde in an alkaline environment in contrast
to the ready reaction of phenol with formaldehyde. To account for this
130 Adhesives for Wood and Lignocellulosic Materials

fact, Caesar and Sachanen postulated the formation of an intramolecular


chelate ring holding the Na+ between phenolic and o-methylol hydroxyl
groups of the PF resin.
The fact that acceleration or retardation of PF resin reaction in an acid,
neutral, and mildly alkaline environment can be effected by bivalent or
trivalent ions such as Zn2+, Ba2+, Cr3+, and others was first reported by
Fraser et al. [16, 17]. Fraser ascribed this to the ring mechanism similar
to the one proposed for Na+; i.e., the Na+ is replaced by the bivalent and
trivalent ions.
Thus, the reaction rate of PF phenolic resins can be improved by
metallic ion catalysis because of the higher proportion of more reactive
free para positions becoming available for the reaction during curing
of the resin. Most electropositive bivalent metallic ions are known to
accelerate PF reaction [20, 21, 24]. The accelerating effect depends on
the rate of replacement of the Na+ ions by bi- or trivalent metal ions in
solution.
The extent of acceleration depends on the type and the concentration
of metal ion. The effectiveness of acceleration follows the following order
[20, 21, 24]:

PbII, ZnII, CdII, NiII > MnII, MgII, CuII, CoIII, CoII > MnIII, FeIII >> BeII,
AlIII > CrIII, CoII

The ratio of o-methyl phenol to p-methylolphenol increased in the


following order with respect to the catalytic activity of metal hydrox-
ides: K < Na < Li < Ba < Sr < Ca < Mg [11, 19]. The valence of the
metal and the size of the hydrated metal cation influence the rate at
which phenol disappeared from the reaction mixture [12]. Accordingly,
magnesium, calcium, and barium hydroxides were found to be more
effective catalysts than lithium, sodium, or potassium hydroxides [11,
12, 19].
Fifty percent replacement of sodium hydroxide by potassium hydrox-
ide improved the curing rate of PF resin. Superior strandboard and
plywood can be made using potassium-modified PF resol resins when
compared to adhesives made from the unmodified resin [37]. Steiner
found that Ca(OH)2 was a more effective catalyst for PF resin systems
than NaOH [46].
Duval et al. employed gel permeation chromatography (GPC) to
determine the effect of reaction parameters such as nature of catalysts,
proportion of starting material, and post-treatments on the composition
of resols produced under these conditions [47]. Different commercial
Phenol–Formaldehyde Resins 131

products could be characterized by GPC. GPC was also found to provide


a clear understanding of the progress of the polycondensation reaction
as a function of time. Reactivity of different groups and unblocked ring
positions could be identified precisely by the method. The hydroxymethyl
group was found to be more reactive in the para position than in the
ortho position.
Pizzi showed that the polymerization time of PF resol is highly depen-
dent on its pH value [24]. The polymerization time increased to reach a
maximum and then decreased as the pH increased. In other words, a pH
of about 4 was the point of lowest reactivity for phenolic resin. However,
the greatest reactivity appeared to occur at pHs around 8–9, the reactiv-
ity becoming then progressively worse as the pH increased towards 10–13
[48]. Gel time of resorcinol–formaldehyde adhesives also followed a sim-
ilar trend as the pH of the resin increased but did not appear to show the
slowing down at very high pH shown by thermoset PF resins.

5.5.2 Effect of Viscosity


According to the Mark–Houwink equation, the MW of polymers is related
to their viscosity. The viscosity of a PF resin increases with the time of reac-
tion. A high formaldehyde-to-phenol ratio produces a high-viscosity PF
resin [49]. These authors also suggested that a lower reaction temperature
favors a more uniform MW distribution.
While evaluating the performance of flakeboard bonded with PF res-
ins of different formaldehyde-to-phenol ratios and MWs, Nieh and Sellers
reported that increased viscosity increased the internal bond (IB) strength
of flakeboard [50]. But MOR was not significantly influenced by viscosity
since the MOR depends essentially on the size and slenderness factor of
the wood flakes.
A study by Gollob showed that high-viscosity resins (i.e., high MW and
highly branched resin) produced low wood failure because of poor flow
and penetration [36].
Commercial phenolic resins currently used as binders for wood-
based materials contain relatively high MW PF condensates. Oldörp and
Marutzky found that the addition of urea to PF resin reduces the vis-
cosity of the resin [51]. It was also reported that the bending and the
IB strengths of the particleboard made from the resin increased and the
thickness swelling decreased. Particle boards made with urea-modified
PF glues showed the same resistance to hydrolysis as boards made with
unmodified PF glues. Interestingly, it was also found that almost all the
urea added to the glue could be extracted with water from the cured
132 Adhesives for Wood and Lignocellulosic Materials

resin, indicating that the urea does not become part of the resin during
the curing process.

5.5.3 MW and Its Distribution of PF Resin


One of the most important characteristics of commercial PF resins is the
MW and its distribution. It depends on a number of operating variables
employed in the resin synthesis such as phenol-to-formaldehyde ratio,
alkali concentration, degree of condensation, and temperature. In general,
MW of PF resins increases with increasing F/P ratio and initial NaOH/
phenol molar ratios [24, 36]. Gollob found that a high initial NaOH level
yielded a higher MW and less branched polymer network when cooked to
the same viscosity, i.e., same degree of condensation and with equivalent
final NaOH levels [36].
Wilson and Krahmer reported that phenolic resin of low to medium
MW decreased the IB strength of particleboard obviously due to glue line
starvation as a result of excessive penetration of the resin into the sub-
strate [52]. In contrast, high-MW PF increased the IB. Christiansen and
Gollob [53] found that in parallel-laminated panels, a higher-MW resin
gave higher wood failure than a low-MW resin.
On the contrary, Nieh and Sellers and Park reported that resins with
higher MW caused a decrease in IB strengths of the particleboard [37, 38,
50]. This is probably because of inadequate resin wetting of the substrate.
Stephens and Kutscha showed that the phenolic resin should contain
both low- and high-MW components in order to achieve optimum board
properties [54]. It is believed that the primary function of the low-MW
adhesive is to cross-link between the higher-MW molecules. In addition,
the low-MW fraction of the PF adhesive can have a high gross adhesive
penetration into the wood substrate and can efficiently wet the wood sur-
face. The high-MW components can enhance the cohesive strength of the
cured adhesive at the bond line.

5.6 Commercial Phenolic Resin for Wood Products


For the applications in the production of wood panel products, the follow-
ing types of phenolic adhesives are supplied by the resin producers:

1. Liquid resin, mostly dissolved in water or alcohol


2. Powdered resin (spray dried)
3. Resin films
Phenol–Formaldehyde Resins 133

5.6.1 Spray Drying of Phenolic Resin


The patent by Chiu discloses the method of spray drying a liquid resin
composition containing a low advanced PF resin [55]. The spray dryability
of the low-MW PF resin is enhanced by the addition of the oxo boron com-
pound. This resin is suitable for the production of waferboard/oriented
strandboard.
Three steps are involved for preparing the spray-dried resin: (a) prepa-
ration of a low-MW phenol–aldehyde resin, (b) addition of a water-soluble
oxo boron compound, and (c) spray drying of the liquid resin.

5.6.1.1 The Spray Drying Process


The phenolic resin with a solids content of 25% to 45% by weight and
a viscosity of about 30 to 300 Cps at 25°C is first converted into fine
spray by passing through nozzles. In feed resin solids content, surface
tensions, speed of the rotary atomizer, feed rate, and the temperature
differences of the inlet and outlet are to be precisely controlled. A
stream of hot air is employed to evaporate the water from the resin. The
powdered product is separated from the stream of hot air. Due to the
thermosetting nature of the product, the inlet temperature of the hot
air and the outlet temperature are carefully controlled. It is necessary
that the particle size distribution, moisture content, and thermal flow
of the spray-dried resins are controlled properly. Solving the caking
problem is very important. This is done by controlling the moisture
content below 2% to 3%. Drying agents such as calcium silicate and/or
lime at 0.5% to 2.0% based on powder resin weight are mixed with the
resin before bagging.

5.6.2 Phenolic Dry Resin Film


Phenolic dry film adhesives consist of a carrier paper (e.g., kraft paper),
which is impregnated with suitably formulated phenolic resol and dried to
procure a resistant state with a good shelf life.

Advantages of dry film adhesives over the wet glues


Basically, an ideal wood adhesive is one that can be spread as an even
coating on a substrate, and which gives a perfect glue bond when cured
subsequently either at room temperature or at an elevated temperature
in a hot press. A dry adhesive film meets all of these conditions in an
efficient way.
134 Adhesives for Wood and Lignocellulosic Materials

The following are the advantages of a dry film adhesive:

1. A dry adhesive film is simpler to apply than liquid adhesives.


2. All of the untidy and unpleasant mixing and spreading oper-
ations in wet gluing process can be eliminated by the use of
dry adhesive films.
3. Properly manufactured dry adhesive films contain precisely
the same quantity and quality of adhesive and thus can
ensure consistency and reproducibility in strength and other
properties of the bonded wood composites.

5.6.2.1 Types and Grades of Dry Glue Film


A “standard film” that meets all is a 60-g film. This means that over a given
area, the film contains by weight 40 g of resin and 20 g of paper, which is
the carrier required in producing the film. The carrier paper must be thin,
porous, and fibrous, so as to permit a uniform impregnation of adhesive on
both sides. It must be tough enough not to break in handling. In case more
than 40 g of coating is required, as in gluing metal to wood, it is practicable
to use two sheets of dry film glue in each joint. While this does not double
the bond strength, it still results in a substantial increase.

5.6.2.2 Process of Making the Dry Adhesive Film [7]


The impregnation of kraft paper is carried out in a continuous plant that
consists of an unwinding device from which paper is transported to an
impregnation bath filled with phenolic resin. With the help of squeezer
rolls, the quantity of the resin transferred onto the kraft paper is regulated
and controlled. The resin impregnated paper web then passes through a
dryer maintained at a pre-designed temperature and line speed and moves
on a cushion of hot air. The speed and temperature are so adjusted that
the impregnated paper when it emerges out of the machine acquires the
desired characteristics with respect to volatile content, flow, and degree of
condensation of the PF resin. After a cooling section, the paper is cut to
sheets for multi-daylight presses or rewound.

5.7 Curing of Phenolic Resin


The cure of phenolic resin is characterized by two events: gelation and
vitrification. Gelation occurs when the polymer attains infinite MW and
Phenol–Formaldehyde Resins 135

viscosity. The gel further cures to a cross-linked state. At this stage, the gel
undergoes a vitrification process. The polymer is transformed from a rub-
bery state to a glassy state and it is then defined by a glass transition tem-
perature (Tg). In other words, cooperative molecular motions are suddenly
frozen and the resin solidifies. At this point, the resin is insoluble, infusible,
and highly cross-linked into a three-dimensional network.

5.7.1 PF Cure Acceleration


Whereas PF has shown higher bonding efficiency than UF, its cure speed
is slower than that of UF. A slow resin cure rate would lead to a lower pro-
duction rate and higher production costs. This is perhaps the most import-
ant reason that PF resins are not used widely for MDF and PB. However,
if the cure speed of PF could be increased by a catalyst or accelerator, the
economic viability as an MDF/PB adhesive would be improved. There are
many ways to develop fast-curing PF resins: increasing MW, higher resin
solids content, using different catalysts, additives, and various synthesis
formulations. Both increasing MW and solids content of PF resin have
limitations as these prevent adequate penetration of the resin into a sub-
strate, which results in poor adhesive bonds.
Addition of resorcinol to PF resin can result in fast curing of PF res-
ins. Phenol–resorcinol–formaldehyde resins are now well known and
widely used commercially as cold-setting adhesives for finger jointing
and glulam production. The “honeymoon” type of adhesives, namely,
those based on separate application of fast-curing systems, has also
been developed for finger jointing purpose. But these adhesives never
reached the commercialization stage since a very expensive chemical,
namely, m-aminophenol, was used [56, 57]. It was the substitution of
m-aminophenol with the cheaper and more easily available resorcinol
that allowed fast-curing “honeymoon” adhesives to be used for finger
jointing and for the production of glulam on a commercial scale, first
in South Africa and then in other countries [58–61]. These resins were
further upgraded more recently by using higher content of biosourced
materials [62]. All these resins consist of two separate parts each being
applied on either side of the substrates and when joined together set at
ambient temperature.
PF resin also cures at ambient temperature under very acidic con-
ditions. Cameron and Pizzi developed a cold-setting PF resin by using
para-toluene sulfonic acid and a system to self-neutralize during curing
acid setting PF resins [63, 64]. The cure of PF resins can be accelerated
using various catalysts such as ammonia, amines, and amides [65–67].
136 Adhesives for Wood and Lignocellulosic Materials

Fast-curing phenolic resins may be prepared by employing metallic ion


catalysis since they facilitate availability of higher proportions of the free
reactive para positions for the subsequent curing resin. Most covalent
metal ions accelerate PF reaction, and the extent of acceleration depends
on the type and concentration of metal ion present. Also, the greater the
radius of hydrated cation, the faster the disappearance of formaldehyde in
the reaction [68].

5.7.2 PF Cure Acceleration by Additives


The so called α-set in liquid phase or β-set in the gas phase acceleration of
cure of very alkaline PF resins for foundry core binders was pioneered in
the early 1970s [69–71].
In this application, addition of some esters such as propylene carbonate,
methyl formate, and triacetin (glycerol triacetate) was found to accelerate
PF resin curing and reduce resin curing time to an extremely short dura-
tion. This technique is now used extensively for foundry core PF binders
(Lemon 1990) [71]. This mechanism of PF curing acceleration can also be
applied for wood adhesive applications such as for particleboard to reduce
the pressing time and increase the production. Thus, the curing agents may
be selected from the groups of compounds consisting of lactones, organic
carbonates, esters, or mixtures of these. The preferred lactone is gamma-
butyrolactone. Propylene carbonate is an example of a suitable carbonate.
Suitable esters include very low MW esters such as methyl formate and
higher-MW materials such as triacetin (glycerol triacetate) [8, 37, 48, 71,
72] and organic anhydrides [67].
Using ester-modified PF resol resin, Pizzi and Stephanou reported
that the press time could be reduced to 2.5 min in making particleboard
and the internal bond strength was good enough to pass exterior grade
board standard [65, 66]. Furthermore, they concluded that the preferred
ester was glycerol triacetate, which gave fast curing and long pot life of
adhesive.
Tohmura and Higuchi also studied the cure acceleration of PF resol
resin with propylene carbonate [73]. They ascribed the acceleration effect
of propylene carbonate to the hydrogen carbonate ion, which eventually
changed the NaOH/P molar ratio and to hydrolysis taking place even at
room temperature. They concluded that the cure acceleration by propylene
carbonate was catalytic in nature.
Thus, ester-accelerated PF resin may have potential in the manufactur-
ing of wood composite panel products [38, 74, 75].
Phenol–Formaldehyde Resins 137

Preparation of the complex catalyst calcium oxide, zirconium oxide,


sodium carbonate, copper oxide, and some other chemicals in fixed pro-
portion were mixed evenly and heated at 500–800°C for 30 min, then
cooled to room temperature, and used as a complex catalyst for PF resins
in synthesis and curing process [76].

5.7.3 Mechanism
Several mechanisms have been proposed to explain the accelerated cure
[77, 78]. Higuchi et al. [79] and Tohmura and Higuchi [73] proposed a
mechanism in which the bicarbonate anion derived from polyethylene car-
bonate coordinates with two hydroxymethylated phenol molecules form-
ing a transition state structure that facilitates faster reaction. Miller and
Detlefsen [80] proposed a mechanism in which the hydroxymethylated
phenol is transesterified by an organic ester facilitating a faster conversion
to a reactive quinone methide intermediate.
Pizzi and Stephanou instead [65–67] proposed a mechanism in which
carbon dioxide from propylene carbonate is incorporated into the poly-
meric structure of the cured resin. The existence of this latter mechanism
was first confirmed by indirect evidence by subsequent studies by Park et al.
[37, 38] and Pizzi et al. [81]. Its correct mechanism was finally clinched
and the intermediate species formed isolated and characterized by matrix-
assisted laser desorption/ionization time of flight (MALDI-TOF) and 13C
nuclear magnetic resonance (NMR) spectroscopy by Lei et al. [82] ending
the controversy that had arisen about the direct participation or not of
part of the ester to yield extra cross-linking or just act as a catalyst of the
traditional reaction route.
The proposed mechanisms are based on the carbanion behavior of
aromatic nuclei of phenate ions. The ester, or residue of its decomposi-
tion, attacks the negatively charged phenolic nuclei in a polycondensa-
tion reaction resulting in the formation of a higher functionality
(greater than 3) [37, 38, 65, 66, 81, 82]. This could lead to much earlier
gelling of PF resin. However, other acceleration mechanisms by propyl-
ene carbonate was suggested by a Japanese research group [73, 79].
Tohmura and Higuchi [73] proposed that the catalytic action of the
hydrogen carbonate ion produced by the reaction between sodium
hydroxide and propylene carbonate is responsible for the cure acceler-
ation of propylene carbonate.
The whole history and different progress on the mechanisms were well
summarized in detail in the background part of the article by Lei et al. [82].
138 Adhesives for Wood and Lignocellulosic Materials

O=

O=
O OH

=
C C
O O O

H3C HC CH2 H 3C HC CH2

OH
O

=
OH OH
O-

H3C HC CH2 OH
C OH
+
CO2 O O
=

H3C HC CH2
O=

OH OH
C +
O- Na
O=

OH
C OH
+
+
OH OH
H 3C HC CH2
Propylene
carbonate OH OH
H3C HC CH2
O=

OH
O=

C OH OH
C
O=

C
etc

This is quite different from the catalytic action of inorganic carbonates,


namely, sodium and potassium carbonates, in accelerating the cure of phe-
nolic resins as proposed by the Higuchi group. It must be further pointed
out that no proof of the mechanism proposed by the Higuchi group was
ever presented, with the mechanism proposed being simply a proposition
and no more [82]. The catalytic effect of the inorganic carbonates and
amines/amides was explained by Pizzi et al. [81].
However, it is of interest to examine the alternate studies. To eluci-
date the cure-acceleration mechanisms, Kamo et al. investigated and
compared the effects of propylenecarbonate (PC) with those of PC
Phenol–Formaldehyde Resins 139

hydrolysate, sodium bicarbonate (NaHCO3), and ethyl formate on the


condensation reactions of monomeric hydroxymethylphenols (HMPs)
[72]. Immediately after the reaction started, PC decomposed simultane-
ously and accelerated the formation of the ortho–para methylene-bonded
dimer of 2,4,6-trihydroxymethylphenol. This effect of PC was very sim-
ilar to that of ethyl formate. In contrast, PC hydrolysate accelerated the
formation of the para–para methylene-bonded dimer throughout the
course of the reaction. This effect of PC hydrolysate was identical to
that of NaHCO3. These results indicate that PC increases the reactivity
of the ortho-hydroxymethyl group, presumably through transesterifica-
tion. On the other hand, NaHCO3 is formed by the hydrolysis of PC or
decomposition of the transesterified HMPs, and it increases the reactivity
of the para-hydroxymethyl group.
2-HMP and 4-HMP were used as model compounds by Conner et al.
[78] to study the reactions that occur during cure of a PF resin to which
cure accelerators (ethyl formate, propylene carbonate, γ-butyrolactone,
and triacetin) were added. The addition of cure accelerators signifi-
cantly increased the rate of condensation reactions. The cure accelerators
were consumed during the reaction, indicating that they do not act as
true catalysts. A dimeric and a trimeric reaction product were isolated
by preparative thin-layer chromatography, and their structures were
determined by 13C-NMR spectroscopy. These results are consistent with
a mechanism in which the hydroxymethylated phenol of 2-HMP (or
4-HMP) is esterified by the cure accelerator, facilitating its conversion to
a reactive quinone methide intermediate.
A different approach to the acceleration of PF resins was taken by Zhao
et al. [74, 75]. Zhao et al. prepared fast-curing low-condensation PF resins
coreacted under alkaline conditions with up to 42% molar urea on phenol
during resin preparation, yielding PUF resins [74, 75]. The PUF resin so
produced was capable of faster hardening times than equivalent pure PF
resins prepared under identical conditions and presented better perfor-
mance than the latter. This principle was also applied to boards prepared
using the difficult to bond particles derived from palms, with very encour-
aging results [83].
Zhang et al. prepared a PF resin with a formaldehyde/phenol molar ratio
of 1.8, which was accelerated by the addition of 1% propylene carbonate
[84]. The curing acceleration was investigated by MALDI-TOF mass spec-
trometry. After addition of this small proportion of propylene carbonate,
the gel time decreased from 40 to 28 min. From the spectra of MALDI-
TOF, it was clear that the curing reaction of the PF resin was accelerated
by propylene carbonate; there was no obvious difference between the final
140 Adhesives for Wood and Lignocellulosic Materials

structures of the PF resin with and without the accelerator, indicating again
the complexity of the mechanism, and the reason why a controversy had
existed. The final proof of mechanism was that of Lei et al. [82].
Tohmura [85] studied the influence of merbau wood extractives on the
gelation rate of a phenolic adhesive and the effects of some cure accelera-
tors on the bond performance of merbau plywood. The addition of merbau
wood extractives slightly increased the gelation rate of the phenolic resin.
This increase in the gelation rate was revealed to be due to a fall in the resin
pH caused by addition of the extractives. The addition of cure accelerators,
sodium carbonate and propylene carbonate, caused a considerable reduc-
tion in the hot-pressing time required for the merbau plywood to achieve
sufficient bond qualities. Brushing veneer surfaces caused an increase in
bond qualities. The combination of the cure acceleration and the surface
brushing greatly improved the bondability of merbau wood. The main fac-
tor of gluing difficulty is considered to be the poor wettability of the veneer
surfaces resulting from the accumulation of migrating extractives.
Zhang et al. reported the preparation of complex catalysts [76]: calcium
oxide, zirconium oxide, sodium carbonate, copper oxide, and some other
chemicals in fixed proportion were mixed evenly and heated at 500–800°C
for 30 min, then cooled to room temperature, and used as a complex cat-
alyst for PF resins in synthesis and curing process. For increasing the cur-
ing rate and decreasing the curing temperature, modified PF resins were
synthesized under a complex catalyst. The bonding strength and formal-
dehyde emission of the plywood bonded by them were measured accord-
ing to Chinese National Standards methods. The curing behavior was
observed by differential scanning calorimetry measurement. The results
indicate that PF resins catalyzed by the complex catalyst show more mod-
erate pH values, lower curing temperature, and shorter gel time compared
with control ones. Plywood bonded with modified PF resins showed good
bonding strength and low formaldehyde emission even at low hot-press
temperature (110°C), which is closed to the plywood bonded with normal
control PF resin at high hot-press temperature (130°C).
Another article by Kamo et al. describes the catalytic effect of NaHCO3
on condensation reactions of monomeric HMPs to elucidate the cure-
acceleration mechanism [86]. By comparison of the kinetics of self-
condensations of HMPs, NaHCO3 was proved to increase the reactivity of
para-hydroxymethyl groups. The changes of 13C NMR chemical shifts on
each HMP system with the additive indicated that the addition of NaHCO3
enhanced some molecular interactions between HMPs and NaHCO3,
facilitating a resonance effect that might play a similar role in dissocia-
tion of the phenolic hydroxyl groups of HMPs. In addition, computational
Phenol–Formaldehyde Resins 141

modeling by molecular orbital calculations elucidated that hydrogen


carbonate anion (HCO3 ) forms an interaction between either the para-
hydroxymethyl group and the phenolic hydroxyl group or between the two
para-hydroxymethyl groups of HMPs by hydrogen bonds. From the exper-
imental results, the authors proposed the mechanism of the catalytic action
of NaHCO3: it appears to be due to the delocalization of an electron initi-
ated by the interaction of the para-hydroxymethyl groups and the phenolic
hydroxyl of HMPs with HCO3 through hydrogen bonds, which results in
facilitating the formation of active species.
Finally Pizzi, Mtsweni, and Parsons studied the acceleration induced on
the curing of PF resins by the wood substrate itself [87]. In this work, the
authors determined that the reaction of polycondensation/hardening of
PF resins in the presence of wood has a lower energy of activation than for
the PF resin alone. They determined that two effects appear to be present
when a PF resin cures on a wood surface, both induced by the polymeric
constituents of the wood substrate, namely, polymeric carbohydrates and
lignin. These appear to be (i) the catalytic activation of the resin self-
condensation induced particularly by carbohydrates such as crystalline
and amorphous cellulose and hemicelluloses, and (ii) the formation of
resin/substrate covalent bonding, particularly in the case of lignin. The
first of these causes appears to be, by far, the major cause of the lowering of
the activation energy of curing. The contribution of the second appeared
to be very small, practically negligible under the conditions prevalent in
the application of thermosetting PF adhesives. Molecular mechanics cal-
culations indicated that the marked catalytic activation of PF resin con-
densation and curing appears to be induced by the strong set of secondary
force interaction between the PF adhesive and the substrate. These appear
to weaken bonds, which, by cleavage, leads to PF adhesive condensation
and curing.

References
1. Gardziella, A., Pilato, L.A., Knop, A., Phenolic Resins: Chemistry, Applications,
Standardization, Safety and Ecology, 2nd edn., Springer Verlag, Berlin, 2000.
2. Economy, J. and Parkar, Z., Historical Perspectives of Phenolic Resins, vol. 1080,
ACS Symposium Series, American Chemical Society, Washington, US, 2011.
3. L.H. Baekeland, Condensation polymers of aldehydes or ketones with phe-
nols only of aldehydes of formaldehyde, e.g., of formaldehyde formed in situ
with phenol. US Patent 942,699A, 1907.
4. British Standard BS 1755-1:1982 BS EN 1755:2000+A1:2009. Glossary of
terms used in the plastics industry. Polymer and plastics technology.
142 Adhesives for Wood and Lignocellulosic Materials

5. Manasse, O., Ueber eine Synthese aromatischer Oxyalkohole. Chem. Ber., 27,
2409–2413, 1894.
6. Lederer, L., Eine neue Synthese von Phenolalkoholen. J. Prakt. Chem., 50, 2,
223–226, 1894.
7. Pilato L, L., Phenolic Resins: A Century of Progress, Springer-Verlag, Berlin,
2010.
8. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York,
1994.
9. Saunders, K.J., Organic Polymer Chemistry, Chapman and Hall, London, 1973.
10. Megson, N.J.L., Phenolic Resin Chemistry, Butterworth, London, 1958.
11. Knop, A. and Pilato, L.A., Phenolic Resins, Chemistry Applications and
Performance, Springer-Verlag, New York, 1985.
12. Grenier-Loustalot, M.F., Larroque, S., Grenier, P., Phenolic resins, 3. Study of
the reactivity of the initial monomers towards formaldehyde at constant ph,
temperature and catalysts type. Polymer, 37, 939–953, 1996.
13. Whitehouse, A.A.K., Pritchett, E.G.K., Barnett, G., Phenolic Resins, Iliffe
Books, London, 1967.
14. Astarloa-Aierbe, G., Echeverria, J.M., Martin, M.D., Mondragon, I., Kinetics
of phenolic resol resin formation by HPLC. 2. Barium hydroxide. Polymer,
39, 3467–3472, 1998.
15. Caesar, P.D. and Sachanen, A.N., Thiophene–formaldehyde condensation.
Ind. Eng. Chem., 40, 922–928, 1948.
16. Fraser, D.A., Hall, R.W., Raum, A.J.L., Preparation of ‘high-ortho’ novolak
resins. 1: Metal ion catalysis and orientation effect. J. Appl. Chem., 7, 676–
689, 1957.
17. Fraser, D.A., Hall, R.W., Jenkins, P.A., Raum, J.L., Preparation of ‘high-
ortho’ novolak resins, II: The course of the reaction. J. Appl. Chem., 7, 701–
712, 1957.
18. Peer, H.G., The reaction of phenol with formaldehyde: II. The ratio of ortho-
and para hydroxymethylphenol in the base-catalyzed hydroxymethylation of
phenol. Rec. Trav. Chem., 78, 851–863, 1959.
19. Knop, A. and Scheib, W., Chemistry and Application of Phenolic Resins,
Springer-Verlag, New York, 1979.
20. Pizzi, A., Phenol and tannin resins by reaction of coordinated metal ligands.
Part 1: Phenolic chelates. J. Appl. Polym. Sci., 24, 1247–1255, 1979.
21. Pizzi, A., Phenol and tannin resins by reaction of coordinated metal ligands.
Part 1: Phenolic chelates. J. Polym. Sci. Polym. Lett., 17, 489–491, 1979.
22. Yeddanapalli, L.M. and Francis, D.J., Kinetics and mechanism of the alkali
catalysed condensation of o- and p-methylol phenols by themselves and with
phenol. Makromolekulare Chem., 55, 74–86, 1962.
23. Jones, T.T., Some preliminary investigations of the phenol–formaldehyde
reaction. J. Soc. Chem. Ind., 65, 264–275, 1946.
24. Pizzi, A., Phenolic resin wood adhesives, in: Wood Adhesives; Chemistry and
Technology, vol. 1, A. Pizzi (Ed.), Marcel Dekker, New York, 1983.
Phenol–Formaldehyde Resins 143

25. Shono, T., Proc. World Eng. Congr. Tokyo, 31, 533, 1931.
26. Hultschz, K., Studien auf dem Gebiet der Phenol-Formaldehyd-Harze, XIV.
Mitteil.: Über die Ammoniak-Kondensation und die Reaktion von Phenolen
mit Hexamethylentetramin. Chem. Ber., 82, 16–25, 1949.
27. Zinke, A., The chemistry of phenolic resins and the process leading to their
formation. J. Appl. Chem., 1, 257–266, 1951.
28. Sojka, S., Wolfe, R.A., Guenther, G.D., Formation of phenolic resins:
Mechanism and time dependence of the reaction of phenol and hexamethy-
lenetetramine as studied by carbon-13 nuclear magnetic resonance and Fourier
transform infrared spectroscopy. Macromolecules, 14, 1539–1543, 1981.
29. Pichelin, F., Kamoun, C., Pizzi, A., Hexamine hardener behaviour—Effects
on wood glueing, tannin and other wood adhesives. Holz Roh Werkst., 57,
305–317, 1999.
30. Kamoun, C. and Pizzi, A., Mechanism of hexamine as a non-aldehyde poly-
condensation hardener, Part 1: Hexamine decomposition and reactive inter-
mediates. Holzforsch. Holzverwert., 52, 1, 16–19, 2000.
31. Kamoun, C. and Pizzi, A., Mechanism of hexamine as a non-aldehyde poly-
condensation hardener, Part 2: Recomposition of intermediate reactive com-
pounds. Holzforsch. Holzverwert., 52, 3, 66–67, 2000.
32. Pizzi, A., Phenolic resin adhesives, in: Handbook of Adhesives Technology,
2nd edn., A. Pizzi and K.L. Mittal (Eds.), Marcel Dekker, New York, 2003.
33. Nagarajan, R. and Kumar, R.N., A predictive fuzzy-logic controller with an
adaptive loop for the manufacture of resin adhesives. Comp. Industr. Eng., 39,
145–158, 2001.
34. Kumar, R.N. and Nagarajan, R., Fuzzy logic controlled reaction calorimeter
for the determination of heat of polymerization reactions. Eur. Polym. J., 34,
1801–1807, 1998.
35. Kumar, R.N., Nagarajan, R., Fun, F.C., Seng, P.L., Effect of process vari-
ables on the exothermicity during the production of PF resins-modeling by
response surface methodology. Eur. Polym. J., 36, 2491–2497, 2000.
36. Gollob, L., The correlation between preparation and properties in phenolic
resins, in: Wood Adhesives, Chemistry and Technology, vol. 2, A. Pizzi (Ed.),
Marcel Dekker, New York, 1989.
37. Park, B.-D., Cure acceleration of Phenol–Formaldehyde (PF) Adhesives for
Three-Layer Medium Density Fireboard (MDF), PhD. Dissertation, Université
Laval, Québec, Canada, 1999.
38. Park, B.-D., Riedl, B., Hsu, E.W., Shields, J., Differential scanning calorimetry
of phenol–formaldehyde resins cure-accelerated by carbonates. Polymer, 40,
1689–1699, 1999.
39. C.F. van Epps, Water soluble phenolic resin and methods. US Patent
2,360,376, 1994.
40. Parker, R.J., The Effect of Synthesis Variables on Composition and Reactivity
of Phenol–Formaldehyde Resins. PhD Dissertation, Oregon State University,
Corvallis, OR, 1982.
144 Adhesives for Wood and Lignocellulosic Materials

41. Paul, I.I. and Bochkareva, I.V., Chem. Abstr., 61, 13160B, 1964.
42. J.T. Stephan, R.A. Jarvi, J.R. Ash, Phenolic resin adhesive. US Patent 2,437,981,
1948.
43. D.V. Redfern, Art of making phenol-aldehyde reaction products and the
product thereof. US Patent 2,457,493, 1948.
44. D.V. Redfern, Production of themosetting phenol–formaldehyde resin con-
densation products. US Patent 2,631,098, 1953.
45. A.L. Lambuth, Highly reactive alkaline phenol-formaldehyde condensates.
US Patent 3,275,139, 1966.
46. Steiner, P.R., Phenol–formaldehyde wood adhesive characterization by pro-
ton magnetic resonance spectroscopy. J. Appl. Polym. Sci., 19, 215–225, 1975.
47. Duval, M., Bloch, B., Kohn, S., Analysis of phenol–formaldehyde resols by
gel permeation chromatography. J. Appl. Polym. Sci., 16, 1585–1602, 1972.
48. Pizzi, A. and Stephanou, A., On the chemistry, behaviour and cure accelera-
tion of phenol–formaldehyde resins under very alkaline conditions. J. Appl.
Polym. Sci., 49, 2157–2160, 1993.
49. Baker, D.E. and Honeyford, D.E., Adhesive requirements for overlaying
plywood, in: Adhesives for Wood; Research, Applications, and Needs, R.H.
Gillespie (Ed.), Noyes Pub., New Jersey, 1984.
50. Nieh, W.L.S. and Sellers, T., Jr., Performance of flakeboard bonded with three
PF resins of different mole ratios and molecular weights. For. Prod. J., 41, 6,
49–53, 1991.
51. Oldörp, K. and Marutzky, R., Untersuchungen an Spanplatten mit harnstoff-
modifizierten PF-Harzen. Holz Roh Werkst., 56, 75–77, 1998.
52. Wilson, G.L.J. and Krahmer, R.L., Using resin properties to predict bond
strength of oak particleboard. Adhes. Age, 22, 6, 26–30, 1979.
53. Christiansen, A.W. and Gollob, L., Differential scanning calorimetry of
phenol–formaldehyde resols. J. Appl. Polym. Sci., 30, 2279–2289, 1985.
54. Stephens, R.S. and Kutscha, N., Effect of resin molecular weight on bonding
flakeboard. Wood Fiber Sci., 19, 4, 353–361, 1987.
55. S.-T. Chiu, Spray-dried phenol–formaldehyde resin compositions. US Patent
5,019,618, 1991.
56. Baxter, G.F. and Kreibich, R.E., A fast-curing phenolic adhesive system.
Forest Prod. J., 23, 1, 17–22, 1973.
57. Kreibich, R.E., High speed adhesives for the wood-gluing industry. Adhes.
Age, 17, 1, 26–33, 1974.
58. Pizzi, A., Rossouw, D.duT., Knuffel, W., Singmin, M., Honeymoon” pheno-
lic and tannin-based fast setting adhesive systems for exterior grade finger-
joints. Holzforsch. Holzverwert., 32, 6, 140–151, 1980.
59. Pizzi, A. and Cameron, F.A., Fast-set adhesives for glulam. Forest Prod. J., 34,
9, 61–65, 1984.
60. Pizzi, A. and Cameron, F.A., Fast setting phenolic adhesives for glulam
taken to their limits. J. Appl. Polym. Sci. Appl. Polym. Symp., 40, Madison,
Wisconsin, 181–190, 1984.
Phenol–Formaldehyde Resins 145

61. Pizzi, A., Cameron, F.A., Goulding, T.M., Kes, E., van der Westhuizen,
P.K., “Honeymoon” fast-setting adhesives for timber laminating. Holz Roh
Werkst., 41, 61–63, 1983.
62. Mansouri, H.R., Pizzi, A., Fredon, E., Honeymoon fast-set adhesives for
glulam/finger joints of higher natural materials content. Eur. J. Wood. Prod.,
67, 207–210, 2009.
63. Pizzi, A., Vosloo, R., Cameron, F.A., Orovan, E., Self-neutralizing acid-set PF
wood adhesives. Holz Roh Werkst., 44, 229–234, 1986.
64. Cameron, F.A. and Pizzi, A., Acid-setting cold-setting phenolic adhesives
for glulam: A controversial issue. J. Appl. Polym. Sci. Appl. Polym. Symp., 40,
229–234, 1984.
65. Pizzi, A. and Stephanou, A., Phenol–formaldehyde wood adhesives under
very alkaline conditions. Part I: Behaviour and proposed mechanism.
Holzforschung, 48, 35–40, 1994.
66. Pizzi, A. and Stephanou, A., Phenol–formaldehyde wood adhesives under
very alkaline conditions. Part II: Esters acceleration mechanism and applied
results. Holzforschung, 48, 150–156, 1994.
67. Pizzi, A. and Stephanou, A., Completion of alkaline cure acceleration of
phenol–formaldehyde resins: Acceleration by organic anhydrides. J. Appl.
Polym. Sci., 51, 1351–1352, 1994.
68. Grenier-Loustalot, M.F., Larroque, F., Grande, D., Grenier, P., Phenolic
resins. 2. Influence of catalyst type on reaction mechanisms and kinetics.
Polymer, 37, 1363–1369, 1969.
69. G.W. Westwood and R. Higgins, British Patent GB2158448A1985.
70. Borden Inc., Japan Kokai Tokkyo Koho J.P.1-132650 A (1989). US Patent
priority 87-102665, 1987.
71. Lemon, P.H.R.B., An improved sand binder for steel castings. Int. J. Mater.
Prod. Technol., 5, 1, 25–55, 1990.
72. Kamo, N., Okamura, H., Higuchi, M., Morita, M., Condensation reactions of
phenolic resins V: Cure-acceleration effects of propylene carbonate. J. Wood
Sci., 50, 236–241, 2004.
73. Tohmura, S. and Higuchi, M., Acceleration of the cure of phenolic resin
adhesives VI: Cure-accelerating action of propylene carbonate. Mokuzai
Gakkaishi, 41, 1109–1114, 1995.
74. Zhao, C., Pizzi, A., Garnier, S., Fast advancement and hardening accelera-
tion of low-condensation alkaline PF resins by ester and copolymerized urea.
J. Appl. Polym. Sci., 74, 359–378, 1999.
75. Zhao, C., Pizzi, A., Kuhn, A., Garnier, S., Fast advancement and hardening
acceleration of low condensation alkaline PF resins by esters and copolymer-
ized urea. Part 2: Esters during resin reaction and effect of guanidine salts.
J. Appl. Polym. Sci., 77, 249–259, 2000.
76. Zhang, S., Li, J., Liu, X., Ou, Y.A., Gao, Q., Fast curing phenol–formaldehyde
resin catalyzed by a complex catalyst. Adv. Mater. Res., 113–116, 2124–2128,
2010.
146 Adhesives for Wood and Lignocellulosic Materials

77. Lorenz, L.F. and Conner, A.H., Accelerated cure of phenol–formalde-


hyde by the addition of cure accelerators: Studies with model compounds.
Proceedings Wood Adhesives 2000, Forest Products Society, Technical Forum
Poster Presentations, 2000.
78. Conner, A.H., Lorenz, L.F., Hirth, K.C., Accelerated cure of phenol–
formaldehyde resins: Studies with model compounds. J. Appl. Polym. Sci.,
86, 3256–3263, 2002.
79. Higuchi, M., Tohmura, S., Sakata, I., Acceleration of the cure of phenolic
resin adhesives V: Catalytic actions of carbonates and formamide. Mokuzai
Gakkaishi, 40, 604–611, 1994.
80. Miller, T.R. and Detlefsen, W.D., A primer on phenol–formaldehyde resin
for the wood products industry. Proceedings Forest Prod Soc Annual Meeting,
Forest Prod. Soc., Madison, WI, 1999.
81. Pizzi, A., Garcia, R., Wang, S., On the networking mechanisms of additives
accelerated PF polycondensates. J. Appl. Polym. Sci., 66, 255–266, 1997.
82. Lei, H., Pizzi, A., Despres, A., Pasch, H., Du, G., Esters acceleration mech-
anisms in phenol-formaldehyde resin adhesives. J. Appl. Polym. Sci., 100,
3075–3093, 2006.
83. Beng Hoong, Y., Pizzi, A., Chuah, L.Abd., Harun, J., Phenol–urea–formalde-
hyde resin co-polymer synthesis and its influence on Elaeis palm trunk ply-
wood mechanical performance evaluated by 13C-NMR and MALDI-ToF mass
spectrometry. Int. J. Adhes. Adhes., 63, 117–123, 2015.
84. Zhang, J., Pizzi, A., Li, J., Zhang, W., MALDI-TOF MS analysis of the accel-
eration of the curing of phenol–formaldehyde (PF) resins induced by propyl-
ene carbonate. Eur. J. Wood Prod., 73, 135–138, 2015.
85. Tohmura, S., Acceleration of the cure of phenolic resin adhesives VII:
Influence of extractives of merbau wood on bonding. J. Wood Sci., 44, 211–
216, 1998.
86. Kamo, N., Tanaka, J., Higuchi, M., Kondo, T., Morita, M., Condensation
reactions of phenolic resins VII: Catalytic effect of sodium bicarbonate for
the condensation of hydroxymethylphenols. J. Wood Sci., 52, 325–330, 2006.
87. Pizzi, A., Mtsweni, B., Parsons, W., Wood-induced catalytic activation of PF
adhesives autopolymerization vs. PF/wood covalent bonding. J. Appl. Polym.
Sci., 52, 1847–1856, 1994.
6
Resorcinol–Formaldehyde Resins
and Hydroxymethyl Resorcinol
(HMR and n-HMR)

6.1 Introduction
Resorcinol or meta-hydroxy benzene is a phenol with the following
structure:
OH

HO

The reaction of resorcinol and formaldehyde (RF) is much more rapid


than the reaction between phenol and formaldehyde. It is known that
the base- or acid-catalyzed resorcinol–formaldehyde (RF) reactions can
lead to polymeric resins, which are currently used as wood adhesives and
composites. RF phenol–resorcinol–formaldehyde (PRF) resin can cure at
ambient temperatures. Thus, they have an advantage over the plain phenol–
formaldehyde (PF) resin. Hence RF and PRF resins are ideally suited for
producing exterior grade glue-lam and finger joints. It has to be men-
tioned that pure RF resins are generally not used as wood adhesive due to
their high cost. To limit such costs, resorcinol is co-reacted with PF resols,
obtaining PRF co-condensed resins. The reaction between resorcinol and
formaldehyde can lead to the formation of linear polymer if the ratio of
resorcinol to formaldehyde is 1:1 or less.
It is to be noticed that the C4 and C6 sites of the aromatic ring of resorci-
nol are highly activated towards electrophilic substitution due to their high
electron density [1]. The C2 site, although strongly activated, suffers from
the steric hindrance induced by the two vicinal hydroxyl groups. The pro-
portion of the methylene (-CH2-) linkages in C4 plus C6 relative to those
in C2 is in the proportion 10.5:1 [2].

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (147–168) © 2019
Scrivener Publishing LLC

147
148 Adhesives for Wood and Lignocellulosic Materials

Due to the similarity of chemical structure of resorcinol to phenol, it


is natural to expect that the reaction products of resorcinol and formal-
dehyde are similar to those of phenol and formaldehyde in that methylol
resorcinols are first formed with the methylol groups in the o- and
p-positions initially. These methyolol derivatives can undergo further con-
densation to form dimers and oligomers in the same manner as their phe-
nolic counterparts though reactions between (1) the two hydroxymethyl
groups and/or (2) hydroxymethyl group and the hydrogen atom of the
unsubstituted o- or p-positions of another resorcinol molecule to form
methylene bridges.

6.2 Reaction between Resorcinol and Formaldehyde


The reaction between resorcinol and formaldehyde under alkaline condi-
tions leads to the formation of methylol groups (hydroxymethyl groups)
and, by further conversion, to quinone methides (QMs). This is given in
Figure 6.1.
QMs are the intermediates for subsequent chain extension and cross-
linking reactions similar to the reaction between phenol and formaldehyde.
From a kinetic point of view, the initial reaction to form dimers is much
faster than the subsequent condensation of these dimers and higher oligo-
mers. The reaction of resorcinol with formaldehyde on an equal molar
basis proceeds at a rate that is approximately 10 to 15 times that of the PF

OH O
O=
O=

CH2O
:
OH CH2O H
HO HO HO HO

OH
O=

O=

CH2 CH2OH CH2OH


=

HO HO
HO
Quinone methide

Figure 6.1 Reaction of resorcinol with formaldehyde through its different methylol
derivatives up to quinone methides.
RF Resins and Hydroxymethyl Resorcinol 149

system. Such a high reactivity of the RF system renders it impossible to


produce these adhesives in resol form. Hence, only the novolak-type resins
without free methylol groups are produced. In the novolak resin, all the
resorcinol nuclei are linked by methylene bridges with no methylol groups
or methylene–ether linkages [2].
In general, for the formulation of PRF resins, the formulator tries to
achieve the best performance with the minimum proportion of the very
expensive resorcinol. The industrial PRF resins have evolved with respect
to cost/performance in such a way that a general proportion of resorcinol
is between 16% and 18% of the total resin solids. In general, for the lower-
molecular-weight PF resins in which resorcinol needs to be grafted, the
proportion is around 1.15–1.2 molar resorcinol for 1 molar PF trimer. A
lower proportion may lead either to long linear polymers, or if even lower
may result in gelling of the resin. A higher proportion of resorcinol would
be an unnecessary addition of cost.
For a more detailed information on RF, PR, urea–resorcinol–
formaldehyde (URF), tannin–resorcinol–formaldehyde (TRF), and lignin–
resorcinol–formaldehyde (LRF), readers should refer to previous reviews [2].

6.3 Comparison between Resorcinol and Phenol


Significant difference exists between PF and RF resins in that the curing
of PF resin generally requires a hot-press temperature of above 160°C,
whereas RF and PRF cure at ambient temperature; i.e., they are cold-
setting resins. This indicates that RF condensations present a much lower
energy barrier and occur at much faster rate even at room temperature.
Superficially, higher reactivity may be attributed to the stronger nucleo-
philicity of resorcinol since it has two phenolic hydroxyl groups, which
lead to higher electron density on the aromatic ring.
Proposal of a suitable mechanism at the molecular level requires an
understanding of the formation of reactive intermediates since they deter-
mine the rate of the overall condensation reaction [3].
Thus, the comparison of the kinetics of the formation of intermediates
during the reactions between PF and RF will throw light on the difference
in the mechanism of the two reactions. However, the experimental kinetic
studies seem to be very difficult since the capture and quantitative eval-
uation of such short-lived intermediates are currently almost impossible
especially for solution reactions. Theoretical approach instead is expected
to give better clarity on the complex reactions involved. In the field of
wood resin adhesives, pioneering investigations employing a molecular
150 Adhesives for Wood and Lignocellulosic Materials

mechanics approach are on record [4–8]. These studies have focused on the
interactions between resin components and wood cellulose, with a view to
establish theoretical models for resin–wood surface interaction, thereby giv-
ing insight into the relationship between resin structure and performance.
Molecular mechanics is a technique based on force field theory and
mainly deals with all the secondary interactions such as hydrogen bond-
ing, van der Waals and electrostatic forces, and their influences on molec-
ular energy and conformations.
Li et al. on the other hand consider quantum chemistry calculations
to be necessary for analyzing the resorcinol–formaldehdye condensation
reaction [3]. This method is based on quantum mechanics and deals with
formation and breakage of chemical bonds, kinetics, and thermodynam-
ics of chemical reactions through rigorous calculations on the electronic
structures of the involved species.
By employing a quantum chemistry method, Li et al. recently studied
the base-catalyzed PF reactions. The QM was confirmed to be the key
intermediate that initiates condensations [9].
RF condensation reactions may share the same mechanism, but the
chemistry of RF resins is obviously more complex than that of the PF one.
Resorcinol has two hydroxyl groups and therefore dissociation of the pro-
tons can produce a singly charged or doubly charged anion (dianion) in
the presence of a base.
As a result, different QM intermediates may be formed.

OH O O
O
H2C CH2

HO HO HO HO
R 4 QM
2 QM CH2
6 QM

O O
H 2C CH2

O O
2 QMA 4 QMA

Figure 6.2 Possible quinone methide intermediates formed by reaction of resorcinol with
formaldehyde.
RF Resins and Hydroxymethyl Resorcinol 151

As can be seen in Figure 6.2, five possible intermediates may be formed


according to the proposed mechanism. 2-QM, 4-QM, and 6-QM are neu-
tral species that are similar to phenol QMs. However, 2-QMA and 4-QMA
are anions that may exhibit different reactivity.
According to the results for PF reactions, the formation of the QM is
the rate-determining step of the overall condensation reaction [9]. Thus,
it can be speculated that formations of RF QMs encounter lower energy
barriers to the reaction of condensation.

6.4 Reactive Positions and Types of Linkages


Comparison between Resorcinol and Phenol
Theoretically, resorcinol has three reactive sites on the ring, namely,
the three sites in ortho to the two hydroxyl groups. Thus, three types of
methylene linkages through 2,2’-, 2,4-, and 4,4’-condensations should be
formed. However, it is interesting that some 13C NMR studies indicated
that the 2,4’- and 4,4’-linkages were found to be dominant, whereas the
2,2’-linkage was observed to be minor or absent [10–13].
Further, the 4,4’-linkage appeared to be more favorable than
2,4’-linkages. These observations imply that the C4-site is more reac-
tive than the C2-site.
The results of the theoretical calculations of Li et al. showed that the
formation of 6-QM and 4-QMA have relatively lower energy barriers [9].
This rationalized previous experimental observations that the 2,4’- and
4,4’-methylene linkages were dominant, whereas the 2,2’-linkage was
almost absent.

6.5 Hydroxymethyl Resorcinol


6.5.1 Introduction
In 1995, a wood primer was developed, which mitigated the damaging
effects caused by stringent simulated weathering. This chemical treat-
ment is known as hydroxymethyl resorcinol (HMR) (Figure 6.3). HMR is
a 5% solids, dilute alkaline solution of resorcinol and formaldehyde, and
is currently the most well-known and effective wood primer available; it
remarkably enhances the moisture durability of wood–adhesive bonds.
The treatment is suggested to enhance the dimensional stability of wood,
152 Adhesives for Wood and Lignocellulosic Materials

OH
OH
HOH2C CH2OH
+ HCHO
NaOH OH
OH
CH2OH
Resorcinol Trihydroxymethyl resorcinol

OH
CH2 CH2OH

HO OH
CH2
HO

OH

Hydroxymethyl resorcinol trimer

Figure 6.3 Reaction of resorcinol with formaldehde to form HMR trimer.

and its effectiveness is attributed to its ability to cross-link within the


wood cell wall [14–17].
Both heat curable phenolic resin adhesives as well as room temperature
curable RF adhesives have produced durable bonds with wood for decades.
However, these resins do not form bonds of adequate durability in the case
of the following circumstances [18]:

(1) In the case of bonding wood treated with preservatives


such as chromated copper arsenate (CCA)
(2) In the case of bonding composites of wood with non-
wood materials such as plastics or metals that may require
epoxies or polyurethane adhesives
(3) In the case of bonding wood with epoxy resins to conform
to the requirements for resistance to delamination as spec-
ified in ASTM D2559 [19]

The above problems were addressed by researchers at the USDA


Forest Service, Forest Products Laboratory, Madison, WI, who first
RF Resins and Hydroxymethyl Resorcinol 153

published results on HMR as a new coupling agent for wood products


[20–22].

6.5.2 Normal HMR


HMR was disclosed in 1995 by Vick and co-workers at the United States
Forest Products Laboratory [21]. Ever since its introduction in the mid-
1990s, HMR gained in reputation among the wood adhesion community
as an effective coupling agent in producing durable bonds under stringent
exterior conditions for a number of resin-adhesive systems, although it
was originally developed for use with epoxies. Subsequently, the appli-
cation was extended to other systems such as emulsion polymer isocy-
anate, polymeric diphenylmethane diisocyanate (pMDI) [23], and PRF
resin [20, 22], for bonding wood to unsaturated polyester-based fiber-
reinforced plastic (FRP). HMR also promotes the exterior durability of
the joints bonded with adhesives such as epoxy, PRF resin [23], and vinyl
ester [24].
HMR was found to dramatically improve the moisture durability of
epoxy bonds to Sitka spruce [21]. The effectiveness of the HMR was
established since it conforms to the requirements of the ASTM D 2559
[19] delamination test. This procedure was a critical hurdle towards
the structural certification of wood adhesives, involving a harsh cyclic
water-weathering process. With no HMR treatment, bisphenol-A epoxy
adhesives failed the ASTM test, while HMR-treated samples passed this
rigorous test.
HMR has also been found to promote bonding of preservative-treated
wood, an extremely difficult substrate for adhesive bonding. Vick et al.
found that the HMR treatment significantly increased internal bond
strength of flakeboards made with CCA-treated wood flakes and a PF
adhesive [22]. However, the HMR treatment showed no effect on the
mechanical properties of flakeboards made with preservative-free flakes
[22]. Christiansen and Conner [25] found that the HMR treatment
enhanced the durability of bonds between CCA-treated lumber and epoxy,
pMDI, EPI, and PRF as confirmed by the conformity to the requirements
of the ASTM D 2559 delamination test. HMR was found to dramatically
improve the moisture durability of epoxy bonds to Sitka spruce [21].
HMR resin was found to function effectively as coupling agent between
wood and for bisphenol-A epoxy adhesive. It is of noteworthy signifi-
cance that the bonded wood joint passed the harsh ASTM D 2559 cyclic
water-weathering delamination test [19].
154 Adhesives for Wood and Lignocellulosic Materials

6.5.3 Formulation of HMR


The HMR coupling agent is applied to the wood surface as an aqueous
solution with low solids content (5%) and is analogous to a low-molecular-
weight RF resin. It contains four main components, which are resorcinol,
formaldehyde, sodium hydroxide, and water. The Ingredients of original
HMR formulation are given in Table 6.1.

6.5.3.1 Mixing Procedure


The original HMR coupling agent is prepared at the time of use by mixing
resorcinol with formaldehyde [formaldehyde-to-resorcinol (F/R) molar
ratio of 1.54] at pH 9 in a 5% solids content aqueous solution at room
temperature. This version of HMR becomes most effective 3 to 4 h after
mixing, and it can be used for up to about 8 h. The duration of reaction
time between preparing the solution and applying it to a wood surface
strongly affects the molecular weight distribution and residual reactivity
of HMR. Reaction times either shorter or longer than the optimum range
result in bonds that are less water resistant to delamination. A fresh mix
must be prepared for each new batch of HMR.
For wood-to-wood bonding using epoxy adhesives, the original ver-
sion of HMR was used by Vick et al. [21], Vick and Okkonen [26], and
Vick et  al. [27]. Accordingly, the ingredients in Table 6.1 are allowed to
react for 4 h at room temperature. After 4 h, the solution is applied onto a
freshly planed wood surface. Because of the amount of water contained in
the HMR solution (95%), the HMR-treated surface needs to dry out prior
to adhesive application. During drying, water evaporates from the wood
surface while the resorcinol and formaldehyde react. A drying time of 18
to 24 h is used under ambient conditions to bond non-aqueous adhesives,
such as epoxy resins [21].

Table 6.1 The ingredients of the original HMR formulation. F/R molar
ratio = 1.54.
Ingredient Percentage
Resorcinol (crystalline) 3.34
Water deionized 90.43
Sodium hydroxide 3 molar solution (10.8% by weight) 2.44
Formaldehyde solution 3.79
RF Resins and Hydroxymethyl Resorcinol 155

6.5.3.2 Limitations to the Use of HMR


Three main obstacles to the commercial use of HMR became apparent:

(1) HMR has no storage life. Thus, every batch has to be mixed
on-site from accurately measured proportions of starting
chemical, which is time-consuming and prone to error.
Each batch has only 3 to 4 h of storage life.
(2) The second obstacle was another time constraint, namely,
a waiting period of 3 to 4 h before the material can be used,
i.e., a duration almost equal to the storage life of the mix.
This makes HMR commercially cumbersome to use, too.
Care of freshly HMR-treated wood is required, because
contamination with, e.g., dust or chemical vapor can
decrease the effectiveness of the HMR.
(3) Furthermore, there was a need to reduce the reaction time
of the HMR solution.

6.6 Novolak-Based HMR


In an effort to make HMR more versatile, Christiansen et al. developed
the so-called “novolak-HMR” (n-HMR) [18, 28]. This is a stable novolak
prepolymer, in which formaldehyde and resorcinol are mixed at a low
mole ratio (0.39 F:R); at this stage, storage life is essentially infinite [18].
Prior to wood application, the mole ratio is increased to 1.54 [18]. Once
the additional formaldehyde is added, the n-HMR solution can be applied
to the wood substrate immediately and up to 7 h later [18]. Weather dura-
bility tests using n-HMR treatment and an epoxy adhesive yielded results
similar to those of Vick et al. [21]. The term “novolak” signifies the fact that
it is a two-stage resin system. In the first stage, the mole ratio of F/R is 0.39,
and in the second stage, the ratio is increased to 1.54 and the final cure is
achieved by a second addition of formaldehyde. Christiansen’s use of the
term novolak reflects the two-stage curing process, but n-HMR is prepared
under alkaline conditions.
As described above, the n-HMR system is carried out in two stages:

(1) The first stage is the preparation of an n-HMR solution.


(2) The second stage is the preparation of the actual coupling
agent by adding additional formaldehyde to n-HMR. This
is called the “activated stage” [18].
156 Adhesives for Wood and Lignocellulosic Materials

Christiansen et al. tested different F/R ratios from 0.23 to 0.46 for the
novolak stage [14, 18, 21]. In this work, an F/R ratio of 0.39 was used.
By using the n-HMR, the advantage is that the coupling agent can be
prepared in the factory and then shipped it to a wood-bonding plant where
the novolak would simply be mixed with the final formaldehyde.

6.6.1 Preparation of n-HMR


The mixing procedure used for n-HMR at 5% solids content is described
by Christiansen et al. [14, 18, 21].

6.7 Bonding Mechanism Using HMR


6.7.1 Mechanism based on the Material Properties of HMR
HMR was designed to be a low-solids-content, low-molecular-weight cou-
pling agent that could penetrate the wood cell wall at the molecular level
[29]. The low molecular weight nature of HMR solution was confirmed by
Vick et al. [27].
HMR in its reactive, application state mostly occurs as monomers,
dimers, and trimers that can penetrate the wood cell wall because the
molecular weight of these HMR molecules is less than 1000 [27].
Based on theoretical calculations of solubility parameters for wood cell
wall components and HMR, it is postulated that HMR most likely will
associate with lignin because of their similar solubility parameters [30].
Protonated HMR has a solubility parameter of 27.5 (J1/2 cm3/2) while lignin,
hemicellulose, and cellulose have solubility parameters of 31.1, 36.3, and
38.6 (J1/2 cm3/2), respectively. Further support for the preference of HMR
for lignin is the impact of HMR on possibly lowering the glass transition
temperature (Tg) of lignin in dynamic mechanical analysis (DMA) experi-
ments on HMR-soaked wood veneers [30].
In order to impart a durable wood–adhesive bond, HMR should form
a cross-linked structure at the cell wall. To prove the importance of the
cross-link density, Christiansen employed substituted resorcinol, namely,
2-methyl resorcinol, to decrease the cross-link density [14]. This substitu-
tion resulted in decreased wood–epoxy bond durability. By virtue of its low
molecular weight, HMR can diffuse into the cell wall to form an interpen-
etrating polymer network.
RF Resins and Hydroxymethyl Resorcinol 157

6.7.2 Mechanism Based on Surface Chemistry


Two primary means to measure surface chemical characteristics of wood
are X-ray photoelectron spectroscopy (XPS) and contact angle analysis. XPS
provides a measure of the elemental composition of the wood surface includ-
ing carbon and oxygen percentages as well as the carbon oxidation state [29].
Contact angle analysis can be used to probe the surface energetics (polar
and dispersive character) as well as adhesive wettability of wood [31, 32].
This XPS result correlates well with the surface energetics, where it was
found that HMR-treated wood reduced the dispersive (nonpolar) charac-
ter, but increased the polar character of the wood surface.
Upon treatment with HMR, treated wood surfaces were found to exhibit an
increase in non-oxidized carbons and a decrease in carbon–oxygen bonds [32].
Any carbonyl or carboxyl functionality remained unaltered or slightly
decreased compared to untreated wood. However, despite the increase in
non-oxidized carbon and the decrease in carbon–oxygen ratios, the oxygen/
carbon (O/C) ratios either remain unaffected or become higher. This XPS
result correlates well with the surface energetics, where it was found that
HMR-treated wood reduced the dispersive (nonpolar) character, but
increased the polar character of the wood surface [31]. The enhanced polar
interaction can be attributed to the hydroxymethyl groups of the HMR.

6.8 Applications of HMR and n-HMR


6.8.1 Bonding to Preservative-Treated Wood
It is known that bonding of wood treated with preservatives has been very
difficult. One of the important preservatives used for wood preservation is
CCA. In the past, it was essentially impossible to both treat wood and then
adhesively bond it to make bigger assemblies, even with durable PRF adhe-
sives since CCA preservatives interfere with bond formation of phenolic-
based adhesives on CCA-treated southern pine. Cr+3 ions and Cu+2 ions
are known to form a complex with phenol and formaldehyde and to affect
the rate of cure of the PF resin [33–36], thereby detracting from achieving
optimum bond strength. However, if CCA-treated southern pine is primed
with HMR, PRF adhesives could produce durable bonds that would meet
the requirements of the ASTM D 2559 delamination test [20]. Without the
primer, the PRF adhesive could meet the requirements. In the past, the
American Institute of Timber Construction also did not certify southern
158 Adhesives for Wood and Lignocellulosic Materials

pine lumber for glulam if the lumber had previously been treated with
CCA until HMR was also used as primer [37].
HMR was also used to make flakeboards from CCA-treated wood [23].
HMR treatment significantly increased internal bond strength of flake-
boards made with CCA-treated wood flakes and a PF adhesive. However,
the HMR treatment showed no effect on the mechanical properties of
flakeboards made with preservative-free flakes [22]. Christiansen and
Conner [25] also found the HMR treatment to enhance the durability of
epoxy, pMDI, EPI, and phenol–resorcinol–formaldehyde (PRF) bonds to
CCA-treated lumber in the ASTM D 2559 delamination test [17].
Alkaline copper quat (ACQ) and copper azole (CA-B), the most prom-
inent substitutes for CCA, are difficult to bond consistently using a PRF
adhesive formulated for bonding to CCA-treated wood.
The bond performance of ACQ- and CA-B-treated wood primed with
n-HMR was improved. Delamination of the PRF-bonded CA-B-treated
southern yellow pine samples was found to be reduced in the ASTM
delamination test [38].

6.8.2 Epoxy–Wood Adhesion


Epoxy resins develop strong bonds to wood in the dry state. Once they
are exposed to repeated soaking in water and drying, the bonds fail to
meet requirements of exterior grade structural wood adhesive. No epoxy
adhesives are known to meet the requirements of ASTM Specification D
2559 [19]. If wood surfaces were coated with HMR before bonding with
bisphenol-A epoxy adhesives, the durability of the bond was dramatically
increased in comparison with the bonding without the coupling agent [21].

6.8.3 Bonding of Fiber-Reinforced Polymer–Glulam Panels


In recent years, wood or wood panels are structurally upgraded by lami-
nation with a thinner layer of FRP composite to give higher strength and
defect-free surface for application as architectural beams and bridge mem-
bers. Such composites also provide a chemically resistant surface. HMR
can function as coupling agent between wood and FRP. For instance, HMR
was found to be effective in bonding E-glass/vinyl ester resin-based FRP
and eastern hemlock glulam panels.
Fiber-reinforced polymers (FRPs) for glulam reinforcement have been
thoroughly researched due to their light weight and high tensile strength
properties [17, 24, 39]. Such glulam panels approximately 4’ wide by 6”
thick have been employed in timber girder bridges [40].
RF Resins and Hydroxymethyl Resorcinol 159

Lopez-Anido et al. investigated the use of an E-glass/vinyl ester resin as


a means of reinforcing eastern hemlock glulam panels [24]. HMR was used
to improve the durability of the wood/vinyl ester resin bond.
The ASTM D 1101 (1997) [41] cyclic delamination test was used to test
the durability of the wood–FRP laminates [17, 24]. The HMR treatment
was found to promote strong, exterior-grade bonds between the wood and
FRP. These bonds were comparable in strength and durability to PRF adhe-
sive bonds; PRF adhesives are commonly considered as the most durable
structural wood adhesives [17, 24].
The wood–FRP composite requires bonding high-polarity wood to fairly
low-polarity thermoset “plastic”. In this application, HMR has proven to be
very useful [23, 42]. Though the PRF adhesive bonds well to wood, it can-
not maintain a durable bond to vinyl ester FRP. However, epoxy bonds well
to HMR-primed wood and to vinyl ester FRP.
Eisenheld employed IR heat for accelerating the drying of n-HMR-
treated wood before making the vinyl ester–glass–wood composites [43,
44]. The standard HMR drying procedure is 24 h at 23 ± 2°C. The rein-
forcement method adopted in the work was SCRIMPTM (an acronym that
stands for Seemann Composites Resin Infusion Molding Process), a resin
transfer-molding process.

6.8.4 Priming Agent for Bondability of Wax-Treated Wood


Various pretreatments have been used to improve the bonding proper-
ties of inactivated/hydrophobic wood surfaces. Among these were cou-
pling agents, bio-products, such as enzymes (xylanases), or other chemicals,
such as tris (polyoxyethylene) sorbitan monooleate [45], sodium hydroxide
(NaOH), calcium hydroxide, nitric acid, hydrogen peroxide, and borax [46].
Kurt et al. employed HMR priming wax-treated wood prior to bonding
with PVAc and MF and reported that it increased the shear strength of
PVAc-bonded specimens under wet conditions and of MF-bonded speci-
mens under dry and wet conditions [47]. The effect on MF-bonded speci-
mens, however, was much more pronounced under wet conditions.

6.9 Special Adhesives of Reduced Resorcinol Content


6.9.1 Fast-Setting Adhesive for Fingerjointing and Glulam
Together with the more traditional fingerjointing adhesives, a series of
ambient temperature fast-curing separate application systems have also
been developed. These eliminate the long delays caused by the use of more
160 Adhesives for Wood and Lignocellulosic Materials

conventional PRF adhesives, which require lengthy periods to set. These


types of resorcinol adhesives are applied separately. They were first devel-
oped in the United States to bond large components where presses were
impractical [48, 49]. Kreibich describes these separate application or
“honeymoon” systems as follows [49]:
Component A is a slow-reacting PRF resin with a reactive hardener.
Component B is a fast-reacting resin with a slow-reacting hardener.
When A and B are mated, the reactive parts of the component react within
minutes to form a joint that can be handled and processed further. Full
curing of the slow-reacting part of the system takes place with time. The
m-aminophenol used for component B is a frightfully expensive chemi-
cal, and for this reason, these systems were discarded and not used indus-
trially [50]. In their original concept, component A is a traditional PRF
cold-setting adhesive at its standard pH of between 8 and 8.5 to which
formaldehyde hardener has been added. Flour fillers may be added or
omitted from the glue mix. Component B is a phenol/meta-aminophenol/
formaldehyde resin with a very high pH (and therefore a high reactivity),
which contains no hardener or only a very slow hardener.
More recently, a modification of the system described by Kreibich has
been used extensively in industry with good success [51, 52]. Part A of the
adhesive is again a standard PRF cold-setting adhesive with powder hard-
ener added at its standard pH. Part B can be either the same PRF adhesive
with no hardener and the pH adjusted to 12 or a 50% to 55% tannin extract
solution at a pH of 12–13, provided that the tannin is of the condensed or
flavonoid type, such as mimosa, quebracho, or pine bark extract with no
hardener [50, 52]. The results obtained with these two systems are good
and the resin not only has all the advantages desired but also is consider-
ably cheaper as a result of the use of vegetable tannins and of the halving of
the resorcinol content of the entire adhesive system [50, 52, 53].
The adhesive works in the following manner. Once the component A
glue mix is spread on one fingerjoint profile and component B is spread
on the other fingerjoint profile and the two profiles are joined under pres-
sure, the reaction of component B with the hardener of part A is very fast.
In 30 min at 25°C, fingerjoints prepared with these adhesives generally
reach the levels of strength that fingerjoints glued with more conventional
phenolic adhesives are able to reach only after 6 h at 40 to 50°C or in 16
to 24 h at 25°C [50, 53]. Clamping of laminated beams (glulam) bonded
with these fast set honeymoon adhesives is in average of only 3 h at ambi-
ent temperature compared with the 16 to 24 h necessary with traditional
PRF resins [52, 53]. These adhesives also present two other advantages,
namely, (i) they are capable to bond without any decrease of performance
RF Resins and Hydroxymethyl Resorcinol 161

at temperatures down to 5°C only and (ii) they are able to bond “green”
timber at high moisture content, a feat that is used in industrial glulam
bonding since their commercial introduction in 1981. Several variations
on the theme exist, such as the “Greenweld” system from New Zealand in
which component B is a solution just composed of ammonia as a strong
accelerator of the PRF + hardener of component A and of a thickener; this
system, however, suffers from the presence of the odor of ammonia, which
is unacceptable in some sophisticated markets.

6.9.2 Branched PRF Adhesives [11, 12, 54]


Recently, another step forward has been taken in the formulation of PRF
adhesives of lower resorcinol content. Liquid RF resin or PRF resins appear
to be mostly linear [55]. The original concept in “branching” erroneously
maintained that if a chemical molecule capable of extensively branching
(three or more effective reaction sites with an aldehyde) the PF and PRF
resins is used after, before, or during, but particularly during or after, the
preparation of the PF resin, the polymer in the branched PRF adhesive
has (1) higher molecular weight than in normal PRF adhesives where
branching is not present and (2) higher viscosity in water or water/solvent
solutions of the same composition and of the same resin solids content
(concentration). It also needs a much lower resorcinol amount on total
phenol to present the same performance of normal linear PRF adhesives
[11, 12, 54, 55]. This can be explained schematically as follows:

Resorcinol −CH2 phenol−CH2 n resorcinol

Resorcinol −CH2 phenol−CH2 n resorcinol

Resorcinol −CH2 phenol−CH2 n resorcinol

n in integer numbers

Resorcinol CH2 − phenol n Phenol − CH2 Resorcinol


CH2 CH2 n

Branching
molecule

CH2
Phenol
CH2
n
Resorcinol
162 Adhesives for Wood and Lignocellulosic Materials

with n > 1 and an integer number and comparable to, similar to, or equal to
n in the preceding scheme for the production of PRF resins.
When comparing linear and branched resins, for every n molecules of
phenol used, a minimum of two molecules of resorcinol are used in the case
of a normal, traditional linear PRF adhesive, whereas only one molecule of
resorcinol for n molecules of phenol is used in the case of a “branched” PRF
adhesive. The amount of resorcinol has then been halved or approximately
halved in the case of the branched PRF resin. A second effect caused by the
branching is a noticeable increase in the degree of polymerization of the
resin. This causes a considerable increase in the viscosity of the liquid adhe-
sive solution. Because PRF adhesives must be used within fairly narrow vis-
cosity limits, to return the viscosity of the liquid PRF adhesive within these
limits, the resin solids content in the adhesive must be lowered, consider-
ably, with a consequent further decrease in total liquid resin of the amount
of resorcinol and of the other materials, except solvents and water. This
decreases the cost of the resin further without decreasing its performance.
Thus, to conclude, the decrease of resorcinol by branching of the resin
is based on two effects:

1. A decrease of resorcinol percentage in the polymer itself,


hence in the resin solids, due to the decrease in the num-
ber of the PF terminal sites onto which resorcinol is grafted
during PRF manufacture.
2. An increase in molecular weight of the resin, which, by the
need to decrease the percentage resin solids content to a
workable viscosity, decreases the percentage of resorcinol on
liquid resin (not on resin solids).

It is clear that, in a certain sense, a branched PRF will behave as a more


advanced, almost precured phenolic resin. While the first effect described is a
definite advance on the road to better engineered PRF resins, the second effect
can also be obtained with more advanced (reactionwise) linear resins. The con-
tribution of the second effect to the decrease in resorcinol is not less marked
than that of the first effect. It is, however, the second effect that accounts for
the difference in behavior between branched and linear PRF adhesives.
Branching molecules that can be used could be resorcinol, melamine,
urea, and others [56]. Urea is the favorite, because it is much cheaper than
the others and needs to be added in only 1.5% to 2% of total resin. When
urea is used as a brancher, the adhesive assumes an intense and unusual
(for resorcinol resins) blue color, after a few days, hence its nickname, “blue
glue”. However, later work has shown that tridimensional branching has
RF Resins and Hydroxymethyl Resorcinol 163

very little to do with the improved performance of these low-resorcinol-


content adhesives, with tridimensionally branched molecules contributing,
at best, no more than 8% to 9% of the total strength [11, 12, 54, 56]. In real-
ity, addition of urea causes the reaction as foreseen, but not in three points
branching but rather only in two sites of the branching molecule. This is
equivalent to saying that most of the resin doubles linearly in molecular
weight and degree of polymerization, while the final effect, good perfor-
mance at half the resin resorcinol content, is maintained [11, 12, 54, 56].
This effect is based on the relative reactivity for phenolic methylols of urea
and of unreacted phenol sites and thus while the macro effect is as wanted,
at the molecular level, it is only a kinetic effect due to the different relative
reactivities of urea and phenol under the reaction conditions used. Thus,

resorcinol–CH2–[–phenol-CH2–]n–resorcinol
resorcinol–CH2–[–phenol-CH2–]n–resorcinol

But

resorcinol–CH2–[–phenol–CH2–]n – UREA – CH2–[–phenol–


CH2–]n –resorcinol

HaIving of the resorcinol content is still obtained, but between 90% and
100% of the polymers in the resin are still linear.
It is noticeable that the same degree of polymerization and “doubling”
effect cannot be obtained by lengthening the reaction time of a PF resin
without urea addition [11, 12, 54, 56]. These liquid resins then work at a
resorcinol content of only 9% to 11%, hence considerably lower than that
of traditional PRF resins. These resins can also be used with good results
for honeymoon fast-setting adhesives in PRF tannin systems, thus further
decreasing the total content of resorcinol in the total resin system to a level
as low as 5% to 6%. This concept was extended to URF cold-setting adhe-
sives, these too giving good results [12].

6.9.3 Cold-Setting PF Adhesives Containing No Resorcinol


As the cost of cold-setting, exterior grade adhesives based on resorcinol is
very high due to the high cost of resorcinol itself, the tendency to decrease
the amount of resorcinol while maintaining unaltered the performance of
the adhesive, when brought to its ultimate conclusion, leads to the concept
of exterior cold-setting phenolic adhesives of zero-level resorcinol. As alka-
line PF resins have not an ambient temperature rate of reaction, which is
164 Adhesives for Wood and Lignocellulosic Materials

even vaguely sufficient to set and harden to a sufficient level the adhesive,
some modifications need to be introduced to overcome in this regard the
lack of resorcinol. This can be done in several manners: (i) by using stan-
dard PF thermosetting resol resins and hardening them by increasing the
glue line temperature by radiofrequency in fingerjointing and glulam man-
ufacture. The system is expensive and needs considerably higher capital
outlay and more careful handling of both the equipment and of the joint,
for results that are certainly not particularly exciting. (ii) By using resins
in which the PF resol of adhesives of type 2 above is terminated by the
terminal grafting of a resorcinol substitute, for example, a natural polyfla-
vonoid tannin, this system being truly cold-setting and yielding relatively
good results but at best just on the inferior limit of the standard require-
ments [57]. (iii) By using self-neutralizing acid setting PF resols. The term
“acid-setting” when used in the presence of a lignocellulosic substrate
makes wood technologists shudder, conjuring visions of extensive acid-
induced substrate degradation and early exterior joint failure. And this is
indeed the case! In reality, some exterior aminoplastic resins do harden in
the moderately acid range without any major substrate degradation prob-
lems. PF resins, however, while hardening very rapidly under acid condi-
tions, do need very acid conditions to give a hardened strong network, and
this elevated acidity is not really acceptable as regards long-term durabil-
ity of the substrate. The damage due to the acid hydrolysis of cellulose and
other wood carbohydrates is particularly aggravated and compounded by
the long-term effect of the glue line remaining acid after resin hardening.
However, the main negative effect due to acid-induced degradation of the
substrate has been overcome by using acid-setting PF resins containing
no resorcinol but hardened by the use of a self-neutralizing catalyst [58].
According to this principle, the adhesive first becomes acid to allow the PF
resin to cure, and after hardening, the hardened glue line self-neutralizes in
a very short time [58]. The greater majority of the effects of substrate deg-
radation are then avoided and very strong and durable exterior wood joints
are produced [58]. The system works well in radiofrequency cured joints,
yielding much better results than the alkaline resols of point (i) above, and
can work well under purely cold-setting conditions [58].

References
1. Raff, R.A.V. and Silverman, B.M., Kinetics of the uncatalyzed reactions
between resorcinol and formaldehyde. Ind. Eng. Chem., 43, 1423–1427, 1951.
2. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York, 1994.
RF Resins and Hydroxymethyl Resorcinol 165

3. Li, T., Cao, M., Liang, J., Xie, X., Du, G., Mechanism of base-catalyzed
resorcinol-formaldehyde and phenol-resorcinol-formaldehyde condensa-
tion reactions: A theoretical study. Polymers, 9, 426, 2017.
4. Pizzi, A. and Eaton, N.J.A., A conformational analysis approach to phenol-
formaldehyde resins adhesion to wood cellulose. J. Adhes. Sci. Technol., 1,
191–200, 1987.
5. Pizzi, A. and De Sousa, G., On the resolution of dihydroxydiphenylmethanes
on achiral crystalline Cellulose II—Correlation of experimental and calcu-
lated results. Chem. Phys., 164, 203–216, 1992.
6. Pizzi, A. and Maboka, S., Calculated values of the adhesion of phenol-
formaldehyde oligomers to crystalline Cellulose. J. Adhes. Sci. Technol., 7,
81–94, 1993.
7. Sedano-Mendoza, M., Lopez-Albarran, P., Pizzi, A., Natural lignans adhesion
to cellulose: Computational vs experimental results. J. Adhes. Sci. Technol.,
24, 1769–1786, 2010.
8. Lopez-Albarran, P., Pizzi, A., Navarro-Santos, P., Hernandes-Esparza, R.,
Garza, J., Oligolignols within lignin-adhesive formulations drive their
Young’s modulus: A ReaxFF-MD study. Int. J. Adhes. Adhes., 78, 227–233,
2017.
9. Li, T., Cao, M., Liang, J., Xie, X., Du, G., Theoretical confirmation of the quinone
methide hypothesis for the condensation reactions in phenol-formaldehyde
resin synthesis. Polymers, 9, 45, 2017.
10. Kim, M.G., Amos, W.L., Barnes, E.E., Investigation of a resorcinol-formaldehyde
resin by 13C-NMR spectroscopy and intrinsic viscosity measurement. J. Polym.
Sci., 31, 1871–1877, 1993.
11. Scopelitis, E. and Pizzi, A., The chemistry and development of branched PRF
wood adhesives of low resorcinol content. J. Appl. Polym. Sci., 47, 351–360,
1993.
12. Scopelitis, E. and Pizzi, A., Urea-resorcinol-formaldehyde adhesives of low
resorcinol content. J. Appl. Polym. Sci., 48, 2135–2146, 1993.
13. Christiansen, A.W., Resorcinol-formaldehyde reactions in dilute solution
observed by C13 NMR spectroscopy. J. Appl. Polym. Sci., 75, 1760–1768,
2000.
14. Christiansen, A.W., Chemical and mechanical aspects of HMR primer in
relationship to wood bonding. Forest Prod. J., 55, 73–78, 2005.
15. Son, J. and Gardner, D.J., Dimensional stability measurements of thin wood
veneers using the Wilhelmy plate technique. Wood Fiber Sci., 36, 98–106,
2004.
16. Sun, N. and Frazier, C.E., Probing the hydroxymethylated resorcinol cou-
pling mechanism with stress relaxation analysis. Wood Fiber Sci., 37, 673–
681, 2005.
17. Hosen, J.C., Fundamental Analysis of Wood Adhesion Primers, PhD
Dissertation, Virginia Polytechnic Institute and State University, Blacksburg,
VI, 2010.
166 Adhesives for Wood and Lignocellulosic Materials

18. Christiansen, A.W., Vick, C.B., Okkonen, E.A., A novolak-based hydroxymeth-


ylated resorcinol coupling agent for wood bonding, in: Proceedings, Wood
Adhesives 2000, Session 3B: Advances in Wood Adhesive Formulations, USDA
Forest Service, Forest Products Laboratory, Madison, WI, 2000.
19. ASTM D 2559 Standard Specification for Adhesives for Structural Laminated
Wood Products for Use Under Exterior (Wet Use) Exposure Conditions,
American Society for Testing and Materials, West Conshohocken,
Pennsylvania, USA 2004.
20. Vick, C.B., Coupling agent improves durability of PRF bonds to CCA-treated
southern pine. Forest Prod. J., 45, 3, 78–84, 1995.
21. Vick, C.B., Richter, K., River, B.H., Fried, A.R., Hydroxymethylated resorci-
nol coupling agent for enhanced durability of bisphenol-A epoxy bonds to
Sitka spruce. Wood Fiber Sci., 27, 1, 2–12, 1995.
22. Vick, C.B., Geimer, R.L., Wood, J.E., Jr., Flakeboards from recycled CCA-
treated southern pine lumber. Forest Prod. J., 46, 11/12, 89–91, 1996.
23. Vick, C.B., Hydroxymethylated resorcinol coupling agent for enhanced adhe-
sion of epoxy and other thermosetting adhesives to wood, in Proceedings of
Wood Adhesives 1995, A.W. Christiansen and A.H. Conner (Eds.), pp. 47–55,
Forest Products Society, Madison, WI, 1996.
24. Lopez-Anido, R., Gardner, D.J., Hensley, J.L., Adhesive bonding of eastern
hemlock glulam panels with E-glass/vinyl ester reinforcement. Forest Prod.
J., 50, 11/12, 43–47, 2000.
25. Christiansen, A.W. and Conner, A.H., Hydroxymethylated resorcinol cou-
pling agent for enhanced adhesion of epoxy and other thermosetting adhe-
sives to wood. Proceedings Nr. 7296 USDA Forest Service, Forest Products
Laboratory and Forest Products Society, pp. 47–55, 1995.
26. Vick, C.B. and Okkonen, E.A., Structurally durable epoxy bonds to aircraft
woods. Forest Prod. J., 47, 3, 71–77, 1997.
27. Vick, C.B., Christiansen, A.W., Okkonen, E.A., Reactivity of hydroxymeth-
ylated resorcinol coupling agent as it affects durability of epoxy bonds to
douglas-fir. Wood Fiber Sci., 30, 312–322, 1998.
28. Christiansen, A.W., Vick, C.B., Okkonen, E.A., Development of a novolak-
based hydroxymethylated resorcinol coupling agent for wood adhesives.
Forest Prod. J., 53, 2, 32–38, 2003.
29. Gardner, D.J., Frazier, C.E., Christiansen, A.W., Characteristics of the
wood adhesion bonding mechanism using Hydroxymethyl Resorcinol, in:
Wood Adhesives 2005, Frihart, C. (Ed.), pp. 93–97, 2005, Session 1B – Bond
Durability.
30. Son, J., Tze, W.T.Y., Gardner, D.J., Thermal behavior of hydroxymethylated
resorcinol (HMR)-treated wood. Wood Fiber Sci., 37, 220–231, 2005.
31. Gardner, D.J., Tze, W.T.Y., Shi, S.Q., Adhesive wettability of hydroxymethyl
resorcinol (HMR) treated wood, in: Proceedings Wood Adhesives 2000, pp.
321–327, Forest Products Society, Madison, WI, 2000.
RF Resins and Hydroxymethyl Resorcinol 167

32. Tze, W.T.Y., Bernhardt, G., Gardner, D.J., Christian, A.W., X-ray photo-
electron spectroscopy of wood treated with hydroxymethylated resorcinol
(HMR). Int. J. Adhes. Adhes., 26, 550–554, 2005.
33. Pizzi, A., Phenolic resins by reactions of coordinated metal ligands. J. Polym.
Sci., Polym. Lett., 17, 489, 1979.
34. Pizzi, A., Phenol and tannin-based adhesive resins by reactions of coordi-
nated metal ligands, Part 1: Phenolic chelates. J. Appl. Polym. Sci., 24, 1247–
1255, 1979.
35. Pizzi, A., Phenol and tannin-based adhesive resins by reactions of coordi-
nated metal ligands, Part II: Tannin adhesives preparation, characteristics
and application. J. Appl. Polym. Sci., 24, 1257–1268, 1979.
36. Vick, C.B. and Christiansen, A.W., Cure of phenol-fornaldehyde adhesive
in the presence of CCA-treated wood by differential scanning calorimetry.
Wood Fiber Sci., 25, 77–86, 1993.
37. AITC, American national standard for wood products- structural glued
laminated timber. ANSI/AITC A190.1-1992, American Institute of Timber
Construction, Vancouver, WA, 1992.
38. Lorenz, L.F. and Frihart, C.R., Adhesive bonding of wood treated with ACQ
and copper azole preservatives. Forest Prod. J., 56, 9, 90–93, 2006.
39. Dagher, H.J., Kimball, T.E., Shaler, S.M., Beckry, A.M., Effect of FRP rein-
forcement on low grade eastern hemlock glulams, pp. 207–214, National con-
ference on Wood Transportation Structures, Madison, WI, 1996.
40. Wipf, T.J., Ritter, M.A., Wood, D.L., Evaluation and testing of timber high-
way bridges, in: Pacific Timber Engineering Conference, G.B. Walford and D.J.
Gaunt (Eds.), pp. 333–340, Rotorua, New Zealand, 1999.
41. ASTM D 1101 97a, Standard Test Methods for Integrity of Adhesive Joints in
Structural Laminated Wood Products for Exterior Use, American Society for
Testing and Materials, West Conshohocken, Pennsylvania, USA, 1997.
42. Christansen, A.W. and Vick, C.B., Hydroxymethylated resorcinol coupling
agent for wood surfaces to produce exterior durable bonds, in: Silanes and
Other Coupling Agents, vol. 2, Mittal, K.L. (Ed.), pp. 193–208, 2000.
43. Eisenheld, L., Measuring the adhesive bond quality of vinyl ester-glass com-
posites on novolak HMR treated wood, MSc Dissertation, BOKU, University
of Agricultural Sciences, Vienna, Austria, 1997.
44. Eisenheld, L., Measuring the adhesive bond quality of vinyl ester–glass compos-
ites on novolak HMR treated wood, PhD Dissertation, University of Maine,
MN, Orono, Maine, 2003.
45. Christiansen, A.W., How overdrying wood reduces its bonding to PF adhe-
sives: A critical review of the literature Part 1 Physical responses. Wood Fiber
Sci., 22, 441–459, 1990.
46. Sernek, M., Comparative analysis of inactivated wood surfaces, PhD
Dissertation, Virginia Polytechnic Institute and State University, Blacksburg,
VI, 2002.
168 Adhesives for Wood and Lignocellulosic Materials

47. Kurt, R., Krause, A., Militz, H., Mai, C., Hydroxymethylated resorcinol
(HMR) priming agent for improved bondability of wax-treated wood. Holz.
Roh. Werkst., 66, 333–338, 2008.
48. Baxter, G.F. and Kreibich, R.E., Fast-curing phenolic adhesive system. Forest
Prod. J., 23, 1, 17–22, 1973.
49. Kreibich, R.E., High speed adhesives for the wood gluing industry. Adhes.
Age, 17, 1, 26–30, 1974.
50. Pizzi, A., Phenolic resins wood adhesives, in: Wood Adhesives Chemistry and
Technology, vol. 1, A. Pizzi (Ed.), pp. 105–178, Marcel Dekker, New York,
1983.
51. Pizzi, A., duT.Rossouw, D., Knuffel, W., Singmin, M., “Honeymoon” pheno-
lic and tannin-based fast setting adhesive systems for exterior grade finger-
joints. Holzforsch. Holzverwert., 32, 140–151, 1980.
52. Pizzi, A. and Cameron, F.A., Fast-set adhesives for glulam. Forest Prod. J., 34,
9, 61–65, 1984.
53. Pizzi, A. and Cameron, F.A., Fast setting adhesives for fingerjoints and glulam,
in: Wood Adhesives Chemistry and Technology, vol. 2, A. Pizzi (Ed.), pp. 229–
306, Marcel Dekker, New York, 1989.
54. Scopelitis, E., Synthetis, characteristics and applied aspects of cold-
setting urea-formaldehyde polymers, M.Sc. Dissertation, University of the
Witwatersrand, Johannesburg, South Africa, 1992.
55. Pizzi, A., Horak, R.M., Ferreira, D., Roux, D.G., Condensates of phenol,
resorcinol, phloroglucinol and pyrogallol, as flavonoids A- and B-rings
model compounds with formaldehyde, Part 1. J. Appl. Polym. Sci., 24, 1571–
1578, 1979.
56. Pizzi, A., Low resorcinol PRF cold set adhesives: The branching principle,
in: Wood Adhesives Chemistry and Technology, vol. 2, A. Pizzi (Ed.), pp. 190–
210, Marcel Dekker, New York, 1989.
57. Pizzi, A., Cameron, F.A., Orovan, E., Cold-set tannin-resorcinol-formaldehyde
adhesives of lower resorcinol content. Holz. Roh. Werkst., 46, 67–71, 1988.
58. Pizzi, A., Vosloo, R., Cameron, F.A., Orovan, E., Self-neutralizing acid-set
PF wood adhesives. Holz. Roh. Werkst., 44, 229–234, 1986.
7
Polyurethane Adhesives

Polyurethanes (PUs) are polymers that contain the urethane group


O

NH C O

in the polymer chain. Commercial development of PU dates from 1937


when O-Bayer (I.G. Farbenindustrie and later Farbenfabriken Bayer)
found that reactions of diisocyanate and glycols gave PUs.

7.1 Introduction
Urea–formaldehyde (UF)-based resins are the predominantly used resin
adhesives for the production of wood panel products, namely, particle-
boards and medium-density fiberboards. In recent years, there has been
increasing concerns about the health concerns due to the emission of
formaldehyde from the panels in actual service. Formaldehyde is asso-
ciated with the risks of cancer and bronchial health issues. This subject
has been discussed in detail in Chapter  13 (“Environmental Aspects of
Adhesives—Emission of Formaldehyde”).
In response to the pressure from US Green Building Council, which
established LEED (Leadership in Energy and Environmental Design) for
“green” or environmentally friendly structures, stringent emissions regu-
lations for formaldehyde emissions were proposed by the California Air
Resources Board. This opened the door for the development of no-added
formaldehyde (NAF) adhesives for the wood panel products.
Since isocyanate-based adhesives belong to the category of NAF adhe-
sives, there is a growing interest internationally in the use of isocyanate-
based adhesives for the production of wood panel products for application
in a number of different service-life environments, including full exterior

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (169–199) © 2019
Scrivener Publishing LLC

169
170 Adhesives for Wood and Lignocellulosic Materials

exposure. Further, these adhesives confer outstanding durability on wood


panels compared to the traditionally used amino resins.
Isocyanate resins were developed during World War II and quickly
became known as adhesives that can bond ‘‘anything to anything’’ [1].
Since they were first introduced successfully to the German particle-
board industry in the early 1970s, the use of MDI (4,4’-methylenediphenyl
isocyanate) binders in composite panels has grown significantly.
Over the last three decades, the forest products industry has increas-
ingly adopted isocyanate wood binders. This trend is partly due to the con-
siderable growth of the worldwide PU industry in general and the PUs
based on aromatic monomer 4,4’-diphenylmethane diisocyanate (MDI),
and the polymeric MDI (pMDI) in particular.
pMDI has established a reputation as an effective adhesive for particle-
board manufacture since the 1970s. The particleboard industry can be con-
sidered as the precursor to other panel products in the use of pMDI-based
adhesives. Concurrent with the growth of the PU industry, the forest prod-
ucts industry was also developing new wood composite materials such as
in general and oriented strand boards (OSB).
These new developments improved wood utilization efficiency because
lower-quality logs, alternative wood species, as well as non-wood lignocel-
lulosic materials could be used to manufacture panels. Following the OSB
production, other wood composite products laminated strand lumber and
I-beams with OSB webs followed suit.
pMDI is therefore an ideal alternative to the existing formaldehyde-
based adhesive systems and thus qualifies as NAF adhesives for the bond-
ing of reconstituted wood composites. pMDI has become an important
binder primarily for bonding OSB and similar particulate wood-based
composites. However, it must be born in mind that pMDI is at present also
under considerable public and government organizations’ pressure due to
health and environment protection concerns.

7.2 History
In the 1930s, organic isocyanates were developed in Germany by Bayer
Chemical Company. Initially, organic isocyanates were used as German air-
craft adhesives in World War II. In 1954, Goodyear synthesized pMDI by
phosgenation of the reaction products of aniline/formaldehyde reaction, but
this process was never patented [2]. Bayer, ICI, and Carwin/Upjohn separately
developed industrial processes for the production of 4,4’-MDI. Wood adhe-
sives consume relatively a small percentage of the total pMDI market [3].
Polyurethane Adhesives 171

7.3 Reactions of Isocyanates


These reactions are very important since they are involved in the adhesive
preparations as well as the applications.
Isocyanates can undergo a number of reactions, such as

1. Reaction with alcohol


O

R-NCO + R'-OH R-NH-C-O-R'


Isocyanate Alcohol Urethane
2. Reaction with amine
O

R-NCO + R'-NH2 R-NH-C-NH-R'


Isocyanate Amine A Substituted
Urea
3. Reaction with water
O

R-NCO + H 2O R-NH-C-O-H R-NH2 + CO2


Isocyanate Carbamic acid Amine
This is very important reaction for adhesion for wood
4. Reaction with ureas
O O O

R-NCO + R-NH-C-NH-R' R-NH-C-N-C-NH-R'


Isocyanate A Substituted
Urea R
Biuret
5. Reaction with urethane
O O O

R-NCO + R-NH-C-O-R' R-NH-C-N-C-O-R'


Isocyanate Urethane
R
Allophanate
6. Dimerization
O
C
2 R-NCO R N N R
C
O
Uretidione
7. Trimerization
O
C
3 R-NCO R N N R

OC CO
N

R
Isocyanurate

It is interesting to note the structural similarity between the isocyanu-


rate and allophenate. The presence of the intramolecular hydrogen bond
172 Adhesives for Wood and Lignocellulosic Materials

O O

R R R R1
N N N O

H
O N O O N
R R
Isocyanurate Allophenate

Figure 7.1 Isocyanurate and allophenate.

in the allophenate structure (Figure 7.1) makes the resin less viscous.
This will enable resins of higher solid contents and lower viscosity to be
obtained.
Among the various reactions mentioned in Figure 7.1 above, the fol-
lowing reactions are of noteworthy significance with respect to wood
bonding:

(1) Reaction of isocyanate with an amine to form a substituted


urea, i.e., reaction 2.
(2) Reaction of isocyanate with water to form an amine and
carbon dioxide, i.e., reaction 3.

Both reactions (2) and (3) occur when isocyanates function as wood
adhesives.
Thus cure of PU wood adhesives follows two steps: firstly, the isocy-
anate reacts with moisture in wood to create a primary amine terminus
and carbon dioxide; secondly, reactions between the terminal amine and
isocyanate yield urea-linked hard segments.
Isocyanates can also react with wood hydroxyl groups according to
reactions (2) and (3).
It has been recognized that moisture considerably influences the adhe-
sive properties and bond performance of PUs in wood [4–7].

7.4 Raw Materials


7.4.1 Isocyanates
7.4.1.1 Aliphatic Isocyanates (Figure 7.2)
The principal aliphatic isocyanates are 1,6 hexamethylene diisocyanate
(HDDI), BIS(4-isocyanatocyclohexyl)methane [H12 MDI], and isopho-
ronediisocyanate (IPDI).
Polyurethane Adhesives 173

NCO
OCN-(CH2)6-NCO
Hexamethylenediisocyanate NCO
HDI

OCN CH2 NCO

H12MDI

Figure 7.2 Aliphatic isocyanates.

NCO
OCN

+ H2O
–CO2

H
O N
NCO
H
N N
OCN NCO
O

Figure 7.3 Biuret of HDI.

Commercial IPDI is a mixture of Z-(cis) and E-(trans) isomers in a 75:


25 ratio. The isomers are difficult to separate. IPDI has two different types
of –NCO groups. Hence, they have different reactivities.
Hexamethylene diisocyanate (HDI) is especially hazardous. The
first less hazardous derivative was a biuret [8], which can be made by
reacting HDI with a small amount of water and removing excess HDI
(Figure 7.3).
The structure of HDI biuret as shown above is idealized. Commercial
products contain various fractions of oligomers. The presence of oligo-
meric biurets makes the average functionality higher than 3.

7.4.1.2 Aromatic Diisocyanates


7.4.1.2.1 Toluene Diisocyanate
There are two isomers of toluene diisocyanates (TDI), viz., 2,4 TDI and 2,6
TDI. Typically, two types of mixtures are commercially available (80% 2,4
174 Adhesives for Wood and Lignocellulosic Materials

CH3 CH3

NCO
OCN NCO

NCO
Toluene 2,4 diisicyanate Toluene 2,6 diisicyanate
65 : 35 TDI

Figure 7.4 TDIs 2,4 and 2,6 isomers (65:35).

CH3 CH3

NCO OCN NCO

NCO
Toluene 2,4 diisicyanate Toluene 2,6 diisicyanate
80 : 20 TDI

Figure 7.5 TDIs 2,4 and 2,6 isomers (80:20).

and 20% 2,6 isomers) or 65% 2,4 TDI and 20% 2,6 isomers as shown below
(Figures 7.4 and 7.5).
Due to high vapor pressure and toxicity of TDI, it is used as a deriva-
tive of high molecular weight and higher functionality. Higher molecular
weight reduces the volatility and hence the toxic hazard and higher func-
tionality leads to increased cross-link density in the cured material. TDI
has differential reactivity between the ortho and para isocyanate groups.
The para isocyanate group is seven times more reactive than the ortho
group. Furthermore, irrespective of which isocyanate group reacts first, the
second group is less reactive than the first [8].
The isocyanurate derived from TDI has a lower toxic hazard than the
monomeric TDI. The trimerization occurs exclusively through the para
isocyanate group.

7.4.1.2.2 Diphenylmethane 4,4’ Diisocyanate (MDI)


The forest products industry has been increasingly employing isocyanate
wood binders. As mentioned earlier, this trend is essentially due to the
Polyurethane Adhesives 175

tremendous growth of the worldwide PU industry. MDI and polyphenyl


polyisocyanates known as pMDI are central to the growth of the PU indus-
try and the forest products industry.
Important raw material for PU adhesives for wood is MDI. MDI is a
monomer and the associated oligomer is called pMDI. The term polymeric
MDI is, however, a misnomer. It is not a polymer. It comprises of a mix-
ture of monomeric and higher-molecular-weight species. Formerly, it was
called crude MDI or technical grade MDI. pMDI is applied as the neat
liquid in quantities of about 2–5% of the dry weight of wood [9]. A very
exhaustive account of how MDI began to be used extensively in the wood
industry has been dealt with by Frazier (2003). It is summarized as follows:
pMDI was first found to be effective as an adhesive for particleboard
manufacture in the late 1960s and early 1970s. Initially, a relatively small,
but significant, market was recognized. As the PU industry grew, simulta-
neously, the forest products industry was developing new types of wood
composites, in particular, OSB. Such developments effected an increased
wood utilization efficiency because lower-quality logs, alternate wood spe-
cies, and lignocellulosic residues could be used to manufacture panel prod-
ucts that were competitive to plywood markets. The growth of the OSB
stimulated development of other technologies, such as laminated strand
lumber and I-beams made with OSB webs.
This rapid expansion led to the growth of MDI-based polyisocyanates
as binders for wood panel products, which offered many advantages for
strand-based wood composites. However, the isocyanate binders have not
secured the entire market for reasons particularly of cost.
The commercial monomeric MDI comprises about 98% 4,4’ MDI,
with 2,4’ and 2,2’ MDI constituting most of the remainder (Figure 7.6).
The isomer ratio of the monomer fraction (the relative proportions of
4,4’-, 2,4’-, and 2,2’-MDI) is readily determined with gas chromatography.
The best measure of purity is through the determination of the isocyanate
content, or percentage–NCO. The isocyanate content is typically 31% to
32% for industrial pMDI. This is based on a simple reaction with excess
dibutylamine followed by back titration of the residual.
Molecular weights of typical pMDI wood binders are from about 255
to 280 g/mol number average molecular mass, and about 470 to 550 g/mol
weight average molecular mass. Typical viscosities are correspondingly
low, approximately 0.175–0.25 Pa s (175–250 cP).
Finally, pMDI surface tension is approximately 41–46 mN/m. In other
words, the 100% organic resin is very low in molecular weight, low in vis-
cosity, and low in surface tension. It is no surprise that this binder wets
readily and penetrates deeply into wood, as we shall discuss later.
176 Adhesives for Wood and Lignocellulosic Materials

OCN NCO NCO

C C
H2 H2
NCO
4,4' MDI 2,4' MDI

C
H2
NCO NCO NCO
2,2' MDI

CH2

OCN CH2

NCO

n = 1,2 3 etc
Polymeric MDI

Figure 7.6 Polymeric MDI.

7.4.1.2.3 Reactivity of MDI


NCO groups present in pMDI do not all have the same reactivity. The
reactivity of an isocyanate group depends on whether the position
ortho to the NCO group is occupied or free. The reactivity is more if
no substituent is present in the ortho position. Thus, 4,4’-MDI is more
reactive than 2,2’-MDI. Furthermore, on average, pMDI is less reactive
than the pure 4,4’-MDI because oligomeric polyisocyanates also have
ortho substituted NCO groups that will detract from its reactivity.

7.5 Catalysts
Reactions of isocyanates with alcohols are catalyzed by a variety of com-
pounds and organometallic compounds. The most widely used catalysts
are tertiary amines, commonly diazobicyclo(2,2,2) octane [DABCO,
Polyurethane Adhesives 177

Bu O C (CH2)10 CH3
N
Sn
H2C CH CH2
2 Bu O C (CH2)10 CH3

H2C CH2 CH2 O


N
DABCO Dibutyl tin dilaurate
DBTDL

Figure 7.7 Catalysts for reactions of isocyanates.

trademark of Air Products] and organotin (IV) compounds , most com-


monly dibutyl tin dilaurate (DBTDL) (Figure 7.7). Combinations of
DABCO and DBTDL often act synergistically.
As has been discussed, isocyanates react with water. It has been
reported [8] that catalyst selection affects the relative rates of reaction of
isocyanates with hydroxyl groups and water. In the reaction of n-butyl
alcohol or water with H12 MDI, dimethyl tin dilaurate gave a significantly
higher reaction rate than DBTDL [8]. A zirconium acetoacetate complex
is said to be more selective than DBTDL (anonymous). Triphenylbismuth
(and other organobismuth catalysts) also catalyze the reaction with alco-
hol rather than the reaction with water [10].

7.6 Blocked Isocyanates


The use of “blocked” polyisocyanates has many advantages in the coating
industry. It permits the formulation of stable one-package coatings, which,
on heating, deblock and lead to the formation of the highly reactive polyiso-
cyanate. Applications for these systems are in diversified areas such as pow-
der coatings, electro coating, wire coatings, and textile finishing. The nature
of the blocking agent has a significant effect on the deblocking temperature
of the isocyanate. Typical blocking agents used include malonates, triazoles,
ε-caprolactam, sulfite, phenols, ketoxime, pyrazoles, and alcohols [11].

7.7 Advantages of pMDI


pMDI is formaldehyde free and is recognized for fast cure rate [12–15]. As
will be discussed later, wood/PMDI bondline formation involves reactions
178 Adhesives for Wood and Lignocellulosic Materials

with water. pMDI can be cured at much higher wood moisture content
than resoles.
There are a number of advantages in using isocyanite adhesives [16]:

1. High adhesive and cohesive strength.


2. Flexibility in formulation: isocyanates phenol–formalde-
hyde (PF)/melamine–formaldehyde (MF)/UF can be made
emulsifiable to mix with UF, PF, or a number of other water-
based adhesives, thus providing a wide range of adhesive
properties and, consequently, end uses.
3. Isocyanates can be liquids at 100% resin, or they can be
made as oil-in-water emulsions.
4. Ability to bond with furnish having high moisture content.
5. No formaldehyde emissions.

The main disadvantage is their sticking to the steel platens and caul
plates of the hot press when used for bonding particleboards or OSB. This
problem has been solved by the use of release agents. But the use of releas-
ing agent will make the surfaces of the panels hydrophobic and hence will
interfere with the adhesion of paints and varnishes and make the surface
lamination with impregnated paper difficult. The most common way to
avoid the sticking problems is to use pMDI for the core of the board and a
different adhesive (UF or PF) for its surface layer.

7.8 PU Adhesive–Wood Interaction


Covalent bonds between an adhesive and wood can improve bond durability.
However, such a bond formation has not been unambiguously detected. The
formation of covalent bonds between wood and adhesive can occur under
the following conditions: (1) adhesive should have high reactivity with the
wood hydroxyls, (2) adhesive should be capable of penetrating the cell wall,
and (3) the adhesive should exhibit strong wettability to wood [17, 18].
Polymeric methylene diphenyl diisocyanate (pMDI) satisfies these char-
acteristics: (1) it reacts with wood hydroxyls [9], (2) it penetrates the cell
wall [19–21], (3) it flows into microvoids if given access by a fracture [22],
and (4) it travels ~1 mm from the applied radial wood surface [23].
Since its introduction in the early 1970s, pMDI has been considered
as the only adhesive for oriented strand board (OSB), laminated strand
lumber, and exterior grade particleboard that can chemically react with
wood cell wall polymers [24–26]. Isocyanate adhesives have indeed been
Polyurethane Adhesives 179

shown to create durable bonds with wood withstanding the swelling and
shrinkage stresses at the bondline interface [27], although there are clear
indications that, in the case of panel products, the time of pressing is far
too short for the isocyanate to be able to react with wood hydroxyl groups
and establish covalent bonds [28] at least in the board core, which deter-
mines the internal bond (IB) strength of the panel. However, it is not clear
whether covalent bonds are necessary for durable wood adhesion.
As mentioned earlier, pMDI is typically used for the core layer adhesive,
and phenol formaldehyde is often applied for the face layers in the indus-
trial OSB manufacturing process. Bonding with pMDI tolerates higher
moisture contents (MC > 16%) and relatively fast press cycles; this allows
for continuous pressing where mats can be injected with steam to acceler-
ate heat transfer and curing [29]. Hence, a major objective is to form strong
and durable bonds to wood within a short duration.
Several studies have been devoted to exploring how pMDI cures and
interacts with wood during adhesive application and hot pressing of OSB
[21, 28, 30–36].
In all these studies, cell wall moisture plays an important role. Bound
water in the cell wall rapidly reacts with isocyanates to form urea- and
biuret-type structures. Thus, this bound moisture has dramatic effect on
cure [30, 34] and, in some cases, the wood species also affect the kinetics
of the pMDI reaction [35]. In the absence of moisture, or under condi-
tions where isocyanate molar equivalents exceed that of water, there is the
possibility of carbamate and allophenate formation. This occurs because
pMDI (1) wets lumina surfaces more effectively than aqueous resins [37],
(2) flows into macrovoids if given access by a fracture [22, 38], and (3) has
the potential to infiltrate into the cell wall [20, 39].
Covalent bonds between an adhesive and wood are expected to improve
bond durability. It is believed that isocyanate-based adhesives are most likely
to form covalent bonds due to the following reasons: (1) isocyanates are well
known to react readily with alcohols to form carbamates [40], (2) wood con-
tains a substantial amount of hydroxyl groups with which isocyanates can
react, and (3) isocyanates have been thought to infiltrate the cell wall [20]. It
was therefore postulated [17] that the isocyanate-based adhesives can typically
form covalent bonds with the wood polymers in the cell wall (Figure 7.1) [40].
The examination of isocyanate chemistry (Figure 7.1) reveals that a vari-
ety of reaction products are possible when isocyanates react with moist wood

(1) The polyhydroxylic nature of wood provides ample oppor-


tunity for the formation of urethanes and thus establish
direct covalent linkages to wood.
180 Adhesives for Wood and Lignocellulosic Materials

(2) Due to the presence of moisture in wood, isocyanate can


react with water to form polyureas and biuret-type struc-
tures, and this reaction is expected to compete with ure-
thane formation.

The reactions between the pMDI and wood/pMDI and moisture result
in complex distribution of products. For reasons of simplicity, reaction
between the phenyl isocyanate with water and alcohol is considered and
Figure 7.8 depicts multiple pathways and products of this reaction.
The two main types of competing mechanisms (a) and (b) mentioned
above are expected during pMDI cure; isocyanates react with water
to form carbamic acid, which quickly expels CO2 to form an aromatic
amine at a rate ka. The reaction rate between the aromatic amine and
other isocyanate molecules exceeds that of water or wood, and a urea is
formed at a rate ku. Further reaction of the urea with phenyl isocyanate
can occur, forming a biuret at a rate kb. It has been demonstrated that
ku  > ka> kb and that ku is highly dependent on H2O concentration and
temperature [41, 42].

ka
(a) H2O + Ph-NCO Ph-NH2 + CO2
Aniline
O=

ku
Ph-NCO + Ph-NH2 Ph-NH-C-NH-Ph
Diphenyl urea
O=
O=
O=

kb
Ph-NH-C-NH-Ph + OCN-Ph Ph-NH-C-N-C-NH-Ph

Ph
Biuret
O=

R1
kc C Ph
(b) R-NCO + R1-OH O N

R2
Carbamate(General)

R1 = β-O-4; β-5; Phenolic or polysaccharide hydroxyl

NH (Allophanate)
R2 = H (Urethane); C R1
O=

Figure 7.8 Multiple pathways for the formation of a wood/isocyanate adhesive bond.
Polyurethane Adhesives 181

When the ratio of water to isocyanate is low, the isocyanate has the abil-
ity to react with hydroxyl groups of wood polymers leading to a carbamate
group at a rate kc. Thus, Wendler and Frazier found that when the wood is
dry (moisture content < 5%), pMDI reacts with wood hydroxyl groups to
form urethane (Figure 7.9) [30].
The carbamate can react with another isocyanate to form an allophanate
linkage. Because of their structural similarities, the difficulty has been in
distinguishing between urea, biuret, carbamate, and allophanate structures
in wood pMDI bondlines [9, 30, 43].
It is therefore clear that the adhesive bonding mechanism between
pMDI and wood cannot be recognized without confirmation of carba-
mate and/or urea formation. Only after confirmation of carbamate and
urea formation is it possible to confirm the formation of the covalent
bond as a contributing mechanism. Yelle et al. employed high-resolution
solution-state NMR spectroscopy to characterize carbamylated cell wall
polymers of wood [18, 44].
All the efforts to prove the reaction between isocyanate and wood, how-
ever, do not justify the existence of covalent bonds between wood and the
isocyanate. In the case of panel products such as particleboard, OSBs, and
other types of panels, the reaction time necessary to form covalent bonds
of this nature is not taken into account in the work of Yelle et al. [44]. In
fact, the core of a board pressed at 190°C–220°C for 3–5 min, as is indus-
trial practice, never reaches a temperature higher than 115–120°C for less
than a couple of minutes. This is too short a time to yield any significant
number of covalent bonds between wood and isocyanate. Pizzi and Owens
clearly indicate that due to the reaction time consideration, covalent bonds
of this type are very unlikely to contribute to the IB strength of isocyanate-
bonded panels [28]. Again, they distinguish the difference between the
board core, subjected to the above limitation, and the board surfaces in
direct contact with the hot press platen where some covalent bonds can
possibly form. Thus, there are clear indications that in the case of panel
products, the time of pressing is far too short for the isocyanate to be able
to react with wood at least in the board core, which alone determines the
IB strength of the panel [28].

H
wood

wood

-O-H +C=N- -CH2- -NCO -O-C-N- -CH2- -


=
=

O O

Figure 7.9 MDI reaction with wood hydroxyls.


182 Adhesives for Wood and Lignocellulosic Materials

7.9 PU–UF Hybrid Adhesives


In literature, attention has been focused on methods to modify and enhance
the performance of UF, MUF, and PF resins by combining with isocyanate
adhesives. Investigations conducted in this respect showed that isocyanates
can function as curing agents for UF resins, improve the glue line strength to
a considerable degree, and improve their water resistance. Pizzi and Walton
and Pizzi et al. showed that a 10–30% addition of pMDI to UF resin effectively
reduces its susceptibility to hydrolysis and accelerates cross-linking reaction,
which, in turn, may shorten the time of pressing of the boards [45, 46]. They
also showed that the addition of isocyanate to UF resins was the only approach
that could work well in the case of plywood. The same approach was shown
to be suitable for MUF and PF adhesives as well. This effect was determined
to be based on the much faster reactivity of the –NCO group of pMDI with
the droxymethyl groups of an aldehyde-based resin than with water. This still
constitutes the only way to prepare in water urethane bridges [45–47] and has
now been proven for aldehydes other than formaldehyde [48].
Furthermore, studies conducted by Mansouri et al. showed that an addi-
tion of pMDI to UF resin at 15% significantly improved resistance of glue
line to the action of hot water [49]. These authors showed that plywood
manufactured under such conditions was characterized by sufficient boil-
ing water resistance.
Lei et al. showed that an addition of 5% of pMDI to MUPF resin makes
it possible to manufacture particleboards with a 130% higher water resis-
tance than the board bonded with pure MUPF resin as measured by IB
strength after the boiling test [50].
Dziurka and Mirski reported that boards manufactured with a 10% addi-
tion of pMDI in the glue mixture, irrespective of their pressing time, passed the
boil test according to the V100 test (EN 1087-1: 1995) at the level required for
the standard for exterior grade boards that are non-load bearing (type P3) [51]
as found previously by other authors [45, 46, 52]. Moreover, the introduction
of pMDI to urea resin even at 2.5% level made it possible to reduce the pressing
time of boards from 22 s/mm to 16 s/mm. The manufactured boards had bet-
ter bending and IB strengths. Further, the introduction of pMDI to UF resin
resulted in an improvement of emission standard of manufactured boards.

7.10 PU–PF Hybrid Adhesives


PF resole and pMDI, the two most widely used thermosetting adhesives for
exterior wood-based composites, can be combined together so that their
Polyurethane Adhesives 183

unique individual qualities can be imparted on the hybrid system. Several


patents have been issued for resole/PMDI hybrid adhesives. The first of
such a patent has been used in industrial production for at least 10 years in
Chile for both plywood, particleboard, and MDF [46, 52, 53].
It is well known that PU resins generally have low thermal resistance
and undergo thermal degradation at temperatures above 200°C. Hence,
the application of PU adhesives is greatly restricted in their processing or
applications in high-temperature environments [54, 55].
Zhang et al. showed that water and heat resistances of one part mois-
ture curing PU could be enhanced by modifying the same with multi-
hydroxymethylated phenol [56]. The authors carried out a thorough
rheological study PF/pMDI blends at different shear rates, which are
characteristic of various operational processes. A study of adhesive–
wood penetration was performed to reveal the cured bondline struc-
ture. The relative cure speed was obtained using dynamical mechanical
analysis. The mechanical properties of PF/pMDI blends bonded wood
specimens were investigated with mode I fracture toughness testing. The
chemistry and morphological features of PF/pMDI blends were further
investigated with cross-polarization (CP) magic angle spinning (MAS)
nuclear magnetic resonance (NMR) with the use of 13C, 15N-doubly
labeled pMDI resin. Advanced differential scanning calorimetry tech-
niques were applied to further complement information obtained in CP/
MAS NMR studies regarding bondline morphology.

7.11 EMDI-Based Adhesives


MDI or pMDI is water incompatible but the NCO group can react with
water. Hydrophilic modification of poly(diisocyanates) has been used to
facilitate water dispersibility by incorporation of hydrophilic groups.
Water-based PU (WBPU/EMDI) adhesives are developed by incorporat-
ing ionic groups into PU backbone [57]. Depending on these ionic groups,
PU dispersions can be classified into cationic, anionic, and non-ionic types.
Non-ionic types contain hydrophilic soft segment pendant groups such
as polyethylene oxide. Anionic dispersions contain dimethylolpropanoic
acid (DMPA). PU cationomers used in aqueous are generally prepared by
incorporating tertiary amine functionality into the backbone. In non-ionic
dispersion necessary to build a high number of hydrophilic polyether seg-
ments into the PU, PU dispersions can be produced in a variety of ways. The
most important processes employed are the acetone process, pre-polymer
mixing process, the melt dispersion process, and the ketamine process [58].
184 Adhesives for Wood and Lignocellulosic Materials

PU–pre-polymer is first synthesized by reacting a macrodiol (e.g.,


ethylene/propylene glycol) and diisocyanate (e.g., pMDI/MDI), followed
by the introduction of hydrophilic groups by reacting with the reactant
such as dimelthylolpropanoic acid, which acts as an internal emulsifier
and becomes part of the main chain of polymer. PUs prepared above are
segmented polymers consisting of alternating soft and hard segments that
constitute a unique micro phase separated.
Acetone process: In this process, initially the isocyanateterminated
PU pre-polymer is prepared in a hydrophilic organic solvent, for exam-
ple, acetone. The chain extended with usually used is sulfonated functional
diamine. The solution is subsequently mixed with water and then the
organic solvent is removed by distillation.
Hot melt process: In the hot melt process, NCO-terminated ionic mod-
ified pre-polymer is capped with urea at high temperature (130°C) to form
biuret. This product is then dispersed in water at around 100°C, and chain
extension is carried out in the presence of water by reacting it with formal-
dehyde. These methylol groups can undergo self-condensation, producing
desired molecular weight.
Ketamine and ketazine process: It is similar to that of pre-polymer
mixing process. In this process, ketamine (ketone blocked diamine) or
ketazine (ketone blocked hydrazine) is used for chain extension [59].

7.11.1 Comparison between EMDI and pMDI


Galbraith et al. [59] reported that EMDI was more efficient than pMDI,
whereas Jones [60] and Lay and Cranley [61] found no benefit from using
EMDI. According to Papadopoulos et al., particleboards bonded with
EMDI were superior in strength to those bonded with pMDI [62]. This was
particularly noticeable when the resins were applied at a rate of 2%. For the
particleboard industry, a dosing rate of 2% EMDI can satisfy the standards
for P3 boards used in interior fitments while 4% EMDI is needed for the
more stringent P5 boards.

7.12 Emulsion Polymer Isocyanate (EPI) Adhesive


EPI adhesive is a two-component adhesive system that combines an emul-
sion component and an isocyanate functional cross-linking component.
They have been used since the early 1970s in Japan [63] and have been
described in the Japanese Industrial Standard JIS K6806 in 1985 [64].
Polyurethane Adhesives 185

EPI adhesives were introduced in the Japanese market for gluing of


wood-based products. The water-based emulsion adhesives with isocya-
nate as cross-linker are used in many parts of the world for production
of different types of wood-based products such as solid wood panels of
different types, parquet, window frames, furniture parts, plywood, finger-
joints, and load-bearing constructions like glulam beams and I-beams.
The curing characteristics of EPI adhesives are quite complex and include
film formation of the emulsion adhesive as well as chemical reactions of
the highly reactive isocyanate groups towards water, hydroxy-, amino-,
and carboxy-groups. The advantages obtained by the use of EPI adhesives
are fast-setting speed, cold curing, light-colored glueline, low creep of the
glueline, and high moisture resistance.
EPI adhesives give very good adhesion and are, because of this, ideally
suited for gluing difficult wood species [63, 64]. EPI adhesives can also be
used for gluing of wood to metal. In general, the basic adhesive (disper-
sion component) and the cross-linking agent (isocyanate component) have
to be mixed according to the prescribed mixing ratio until homogeneous
before application. After mixing, the adhesive must be processed within
the specified pot life.
Like almost all polymers, polymer dispersions have a poor biodegrad-
ability. Once hardened, they will remain in the environment and degrade
abiotically and biologically very slowly. However, they are not classified as
toxic to the environment, nor do they result in bioaccumulation. Therefore,
polymer dispersions are of less significance from an environmental point
of view [66].

7.13 Non-Isocyanate Polyurethanes and Biobased


PU Adhesives
The use of biobased vegetable oils for the preparation of PUs by their reac-
tion with polyisocyanate is a long-known practice. Dedicated and exten-
sive reviews on this aspect of PU synthesis already exist [67, 68], and this
chapter does not pretend to scan the whole literature on such reactions.
However, in this abundant literature, mostly dedicated to the preparation
of foams and lubricants, the references strictly aimed at the preparation
and application of PU adhesives are not too numerous [69, 70]. It must be
equally made clear that most of the developments made in this field for
applications other than for adhesives can also be used for adhesives as well.
The reaction is the same as for synthetic polyols (Figure 7.10).
186 Adhesives for Wood and Lignocellulosic Materials

O H
HO R OH + OCN R' NCO O R O N R' N O
H O
n

Figure 7.10 Schematic reaction of formation of a PU from a polyisocianate and a polyol.

The difference is that HO-R-OH is a natural, biosourced polyol that sub-


stitutes partially or totally a synthetic one.
The use of polyols of natural origin, such as castor oil, is a well-
established practice. It is due to the presence of hydroxyl groups on these
materials, groups that are the preferential sites for the reaction with the
isocyanate groups to form urethanes. Apart from castor oil and its deriva-
tives, which are relatively expensive, the use of other vegetable oils to react
with isocyanates to prepare PUs is increasing due to the demand for envi-
ronmentally friendly green products. The field of adhesives is no exception
to this trend. The adhesive field where these materials have made import-
ant inroads is first of all that of pressure-sensitive adhesives. For example,
epoxidized soybean oil has been cross-linked with dicarboxylic acids such
as sebacic, adipic, and other acids whereas a chromium (III) organometal-
lic derivative has been successfully used as the cross-linker [70, 71]. Such
a pressure-sensitive adhesive had good peel and shear strengths and good
resistance to aging (Figure 7.11).
The only problem that somehow disqualifies this as fully biosourced
was the use of epichloridrin for the preparation of the epoxidized vege-
table oil.
The second adhesive application is for wood adhesives. Polyester polyols
synthesized from potato starch and natural vegetable oils by transesterifi-
cation were used in the preparation of PU wood adhesives by reaction with
toluenediisocyanate. Their performance was reported as being comparable

O O O
O CH2 7
CH2 CH2 4CH3
O O O
O CH2 7
CH2 CH2 4 + HOOC R COOH GREEN PRESSURE-SENSITIVE ADHESIVE

O BIOSOURCED DICARBOXYLIC ACID


CH2 CH2 7CH3
7
O O
EPOXIDIZED VEGETABLE OIL

Figure 7.11 “Green” pressure-sensitive PU adhesive prepared from glycerol and a dicarboxylic
acid. Note that epichloridrin is still used to epoxidize the glycerol.
Polyurethane Adhesives 187

or even superior to commercial, fully synthetic PU adhesives for the same


application [72].
As regards the preparation of PU adhesives starting from condensed
or hydrolyzable tannins, two approaches have been taken: (i) modifica-
tion of the flavonoid tannin in order to render easier the reaction with
isocyanates. This is due to the difficulty in reacting the flavonoid hydroxyl
groups directly with isocyanates. In this type of application, the tannins
are in direct competition with more suitable natural polyols, of which the
literature abounds, to prepare semi-biosourced PUs; (ii) the use of a total
no-isocyanate approach to improve the environment-friendly character of
such adhesives.
For the first approach, benzoylation to reduce the number of hydroxyl
groups of the tannin before reaction with diisocyanate was tried a long
time ago [72]. A more studied approach has been to use lignin and ligno-
sulfonate hydroxypropylated to prepare urethanes [73–75], although this
approach has been more directed towards coatings than adhesives. The
same and more studied approach has been to prepare novel thermosetting
tannin-based PU adhesive resins using the hydroxypropyl and hydroxybu-
tyl derivatives of purified condensed tannins from Pinus pinaster bark and
other condensed tannin species by reaction with diisocyanates [76–79].
Hydroxypropyl tannins with a degree of substitution rising stepwise from
1 to 4 were cross-linked with either an aromatic (pMDI) or an aliphatic
isocyanate (HDI) with good results. Hydroxypropylation and hydroxybu-
tylation are one of the approaches used to react polyphenolic materials
with isocyanates to obtain PU adhesives and resins [76–79] (Figures 7.12
and 7.13). This also renders easier the reaction of flavonoid tannins with

(OR)
(OH)
5’ O
OR
OH
6’
4’ B
B 3’ RO O
HO 8
O CH3 OR
1’ OH
7 2 2’ A C
A C 3
6 NaOH(aq) OR
5 4 OH
RO
OH
H2
C

H
2-hydroxy propyl ether
O

n
R:
H
1-hydroxy propyl ether
O

CH
3
n

Figure 7.12 Reaction of P. pinaster bark tannin with propylene oxide to produce
hydroxypropyl ether tannin derivatives (considering full HP).
188 Adhesives for Wood and Lignocellulosic Materials

OH OH
5’ O
6’
H3C 4’ CH3
8 B 3’
O O
7 1’ O
2 2’
A C 3 + OCN–R–NCO POLYURETHANES
6 CH3
5 4 O
O CH3 OH

OH
HO CH3

Figure 7.13 Formation of PU adhesive and resins by reaction of a hydroxypropylated


polyphenolic tannin with an isocyanate.

isocyanates, due to the introduction of much more approachable hydroxyl


groups into the tannin structure, thus increasing reaction yield.
The same approach, thus to introduce in the tannin structure more
available –OH groups but through a reaction totally different from
hydroxypropylation, consisted in reacting an aldehyde with the tannin
and then use the hydroxyl groups formed by its addition onto the flavo-
noid structure, before their further condensation with other flavonoids, to
react with an isocyanate [45, 46, 53] (Figure 7.14). Such a system has been
used and is used industrially for wood adhesives [53].
More recently, the system has been adapted by eliminating formalde-
hyde and substituting it with glyoxal with good results, rendering the
approach even more environmentally interesting [48]. The reactions
involved are shown in the case of glyoxal in Figure 7.15.
The first reaction step, namely, the reaction of the tannin with the alde-
hyde, is easier than hydroxyalkylation.
However, isocyanates are harmful to human health. Thus, the prepara-
tion of PU adhesives of a high level of biosourced and environment-friendly
material involves by necessity the synthesis of non-isocyanate PUs. This
approach, which avoids the use of any isocyanate, is attracting increasing
interest. It is based on the polycondensation of diamines with dicyclocar-
bonates to lead to polyhydroxyurethanes. This reaction has been studied by
a few research groups [80–95] and leads to polyhydroxyurethanes of rela-
tively low glass transition temperatures and Mn lower than 30,000 g mol−1.
All the work on non-isocyanate PUs is concentrated on synthetic materi-
als. Thus, while isocyanates are definitely not used, synthetic diamines and
synthetic dicyclocarbonates, both of non-natural origin, are used, yielding
PU resins that are not biosourced. While use of glycerol has been reported,
the percentage of biosourced material is still lower than the approach of
using an isocyanate reacting with a natural polyol.
Polyurethane Adhesives 189

OH

HO O
OH

OH
OH

fast +HCHO
OH
HO O
OH

HOH2C OH
OH

+ Tannin oligomer units + OCN–R–NCO Isocyanate

HO OH OH
HH
O OO O HO O
HO OH OH

HO CH2 OH OCN R NH COO CH2 OH


OH OH OH

Tannin–formaldehyde
polymers

Figure 7.14 Reaction of the flavonoid tannin/formaldehyde system with isocyanates to


form PUs. Note that the reaction occurs also in water.

This second approach, i.e., the more radical elimination of the isocya-
nate in the preparation of PU resins, is based on the reaction of a double
cyclic carbonate with a diamine that has been proposed based exclusively
on synthetic, not biosourced materials [80–95] (Figure 7.16).
Only very recently were approaches to non-isocyanate PU adhesives
aimed not only at eliminating the toxic isocyanate but also at improving
the percentage of biosourced material composing the “green” PU adhesive.
The same reaction used for synthetic materials has been used with both

OH OH

HO O HO O
CHO OH OH
+ OCN R NCO
+
CHO HO OH
OH
OH OH OH
OH

HO O
OH

OCN NHCOO OH
OH OH

Figure 7.15 Formation of PU by reaction of a glyoxalated flavonoid tannin with a


polyisocyanate.
190 Adhesives for Wood and Lignocellulosic Materials

O O H
O O
+ N
O O O O N
R
H
OH OH O
N O O O O N
R N O
O O
n

Figure 7.16 Non-isocyanate PU formation by reaction of a dicyclic organic carbonate


with a diamine.

hydrolyzable tannins and condensed flavonoid tannins; these are natural


renewable materials, by reacting them first with dimethyl carbonate, a
non-cyclic carbonate, followed by reaction with hexamethylenediamine to
obtain non-isocyanate urethanes of the type shown in Figure 7.17 [96, 97].
This was the first approach to obtain a biosourced PU without using iso-
cyanates, although, here too as in approach (i) above, the percentage level of
biosourced material was only 45–50%, with the diamine and the carbonate
being still synthetic materials. As the amination of condensed tannins, and
thus the conversion of their hydroxyl groups into amino groups, is an easy
reaction [98], the aminated tannin was used to substitute the hexamethy-
lenediamine, further improving to more than 70% of biosourced material
in the final PU, and this without using any isocyanate to obtain flavonoid
oligomers linked by urethane bridges [99] as shown in Figure 7.18.
What was of further interest in this approach was that it was not nec-
essary to use a purified tannin, as the 10–12% carbohydrate fraction of
commercial, industrial tannin extracts was also found to undergo the same
two reactions leading to urethane bridges [96, 97]. Thus, species in which
either flavonoid tannin oligomers were linked by urethane bridges to a car-
bohydrate monomer when an aminated tannin was used as diamine (Figure
7.19) or where the urethane bridges were formed between the carbonated

OH
HO
O

OH O O O O
H
O N
N O
H
O O O O OH
OH
O–
OH
OH

Figure 7.17 Non-isocyanate diurethane obtained by reacting a precarbonated flavonoid


tannin with a diamine (from Ref. [97]).
Polyurethane Adhesives 191

OH
NH2 OH

O O O HO O
OH OH
O O NH
OH OH
O O O O O NH2
O HO
O
O O O HO
OH OH
O O

Figure 7.18 Non-isocyanate urethane bridge linking a precarbonated flavonoid tannin


dimer with an aminated tannin dimer. Note that only the dimethyl carbonate is of
synthetic origin, all the rest being renewable biosourced materials (from Ref. [99]).

HO
O
OH
O N
HO
H
O O OH
HO

NH O HO
OH
HO O
OH

OH
OH

Figure 7.19 Non-isocyanate diurethane obtained by reaction of a carbonated carbohydrate


monomer with two aminated flavonoid tannin oligomers (from Ref. [99]).

carbohydrate monomer and a synthetic diamine were also identified, such


as shown in Figure 7.20.
Such a finding opens new possibilities for the preparation of non-
isocyanate urethane adhesives of higher biosourced content, and this was
more recently realized by preparing glucose-based and sucrose-based
non-isocyanate PUs that gave very encouraging results for wood and metal
surface coatings and for wood adhesives [100, 101] as well as for rigid PU
foams [102]. The following linear and branched oligomers were identified
(Figure 7.21):
192 Adhesives for Wood and Lignocellulosic Materials

HO
61.60 O
OH 99.23 40.72 26.52 33.78
74.86
O O 154.25 NH NH2
29.76 26.90 42.06
HO 72.73
69.25 73.66

O
40.72 26.52 33.78
157.69 NH2
O NH 29.76 26.90 42.06

Figure 7.20 Non-isocyanate diurethane obtained by reaction of a carbonated carbohydrate


monomer with a diamine. The number indicates the 13C NMR shifts (from Ref. [99]).

OH OH
OH HO
O O
HO OH NH O HO O
O
O NH O HO O
HO O NH O NH
HN O OH
O O HN O O
O OH
OH OH
HO

And
OH
OH HO O
HO OH
O NH O HO O
HO O O NH OH
HN O
O O HN O
O O NH
NH OH
HO O
O O
OH OH
O
HO
OH

Figure 7.21 Examples of linear and branched oligomer structures identified for glucose-
based non-isocyanate PUs (from Ref. [100]).

References
1. Marra, A., Technology of Wood Bonding: Principles in Practice, Van Nostrand
Reinhold, New York, 1992.
2. Henri, U., Chemistry and Technology of Isocyanates, p. 385, J. Wiley & Sons,
New York, 1996.
3. Galbraith, C.J. and Newman, W.H., Reaction mechanisms and effects with
MDI isocyanate binders for wood composites. Proceedings of the Pacific Rim
Bio Based Composites Symposium, Rotorua, New Zealand, vol. 130, 1992.
4. Abbott, S.G. and Brumpton, N., The effect of moisture on polyurethane
adhesives. J. Adhes., 13, 1, 41–51, 1981.
5. Yang, B., Huang, W.M., Li, C., Lee, M., Li, L., On the effects of moisture in a
polyurethane shape memory polymer. Smart Mater. Struct., 13, 191–195, 2004.
6. Yang, B., Huang, W.M., Li, C., Chor, J.H., Effects of moisture on the glass
transition temperature of polyurethane shape memory polymer filled with
nano-carbon powder. Eur. Polym. J., 41, 1123–1128, 2005.
Polyurethane Adhesives 193

7. Yang, B., Huang, W.M., Li, C., Li, L., Effects of moisture on the thermo-
mechanical properties of a polyurethane shape memory polymer. Polymer,
47, 1348–1356, 2006.
8. Wicks, Z.W., Jones, F.N., Pappas, S.P., Organic Coatings Science and
Technology, 2nd edn., Wiley Interscience, New York, 1999.
9. Zhou, X. and Frazier, C.E., Double labelled isocyanate resins for the solid-
state NMR detection of urethane linkages to wood. Int. J. Adhes. Adhes., 21,
259–264, 2001.
10. Luo, S.G., Tan, H.-M., Zhang, J.-G., Wu, Y.-J., Pei, F.-K., Meng, X.-H., Catalytic
mechanisms of triphenyl bismuth, dibutyltin dilaurate, and their combination
in polyurethane-forming reaction. J. Appl. Polym. Sci., 65, 1217–1225, 1997.
11. Blank, W.J., He, Z.E., Picci, M.E., Catalysis of Blocked Isocyanates with
Non-Tin Catalysts, King Industries Inc. Norwalk, CT 06852, U.S.A. http://
www.wernerblank.com/pdfiles/paper18.pdf.
12. Hawke, R.N., Sun, B., Gale, M.R., Of fibre mat moisture content on strength
properties of polyisocyanate-bonded hardboard. For. Prod. J., 42, 11/12,
61–68, 1992.
13. Hawke, R.N., Sun, B., Gale, M.R., Effect of fibre mat moisture content on
physical properties of polyisocyanate-bonded hardboard. For. Prod. J., 43,
11/12, 15–20, 1993.
14. Hawke, R.N., Sun, B., Gale, M.R., Effect of polyisocyanate level on physical
properties of wood fibreboard. For. Prod. J., 44, 3, 34–40, 1994.
15. Hawke, R.N., Sun, B., Gale, M.R., Effect of polyisocyanate level on physical
properties of wood fibre composite materials. For. Prod. J., 44, 4, 53–58, 1994.
16. Wilson, J.B., Isocyanates adhesives as binders for composition boards.
Proceedings, Adhesives–Research, Application Needs, Sponsored by AC4 U.S.
Forest Products Laboratory and Washington State University, 1980.
17. Yelle, D.J., Jakes, J.E., Ralph, J., Characterizing polymeric methylene diphenyl
diisocyanate reactions with wood: 1. High-resolution solution-state NMR
spectroscopy. Proceedings, Wood Adhesives 2009, C.R. Frihart, C.G. Hunt,
R.J. Moon (Eds.) pp. 338–342, 2009.
18. Yelle, D.J., Ralph, J., Frihart, C.R., Delineating pMDI model reactions with
loblolly pine via solution-state NMR spectroscopy. Part 1. Catalyzed reactions
with wood models and wood polymers. Holzforschung, 65, 131–143, 2011.
19. Marcinko, J.J., Rinaldi, P.L., Bao, S., Exploring the physicochemical nature of
PMDI/wood structural composite adhesion. For. Prod. J., 49, 75–78, 1999.
20. Marcinko, J.J., Newman, W.H., Phanopoulos, C., The nature of the MDI/
wood bond. Second Pacific Rim Bio-Based Composites Symposium, Vancouver,
Canada, pp. 286–293, 1994.
21. Ni, J. and Frazier, C.E., On the occurrence of network interpenetration in the
wood–isocyanate adhesive interphase. Int. J. Adhes. Adhes., 18, 81–87, 1998.
22. Roll, H., Troger, G., Wegener, G., Grosser, D., Fruhwald, A., Untersuchung
zur PMDI-Verteilung in Spangemischen und Spanplatten. Holz Roh Werkst.,
48, 405–408, 1990.
194 Adhesives for Wood and Lignocellulosic Materials

23. Buckley, C.J., Phanopoulos, C., Khaleque, N., Engelen, A., Holwill, M.E.J.,
Michette, A.G., Examination of the penetration of polymeric methylene
di-phenyl-di-isocyanate (pMDI) into wood structures using chemical-state
x-ray microscopy. Holzforschung, 56, 215–222, 2002.
24. Deppe, H.J. and Ernst, K., Isocyanate als Spanplattenbindemittel. Holz Roh
Werkst., 29, 45–50, 1971.
25. Deppe, H.J., Technische Fortschritte bei der Isocyanatverleimung von
Holzspanplatten. Holz Roh Werkst., 35, 295–299, 1977.
26. Frisch, K.C., Rumao, L.P., Pizzi, A., Diisocyanates as wood adhesives, in:
Wood Adhesives, Chemistry and Technology, A. Pizzi (Ed.), pp. 289–317,
Marcel Dekker, New York, 1983.
27. Frazier, C.E., Isocyanate wood binders, in: Handbook of Adhesive Technology,
2nd edn., A. Pizzi and K.L. Mittal (Eds.), pp. 681–694, Marcel Dekker, New
York, 2003.
28. Pizzi, A. and Owens, N.A., Interface covalent bonding vs. wood-induced cat-
alytic autocondensation of diisocyanate wood adhesives. Holzforschung, 49,
269–272, 1995.
29. Sellers, T., Jr., Wood adhesive innovations and applications in North America.
For. Prod. J., 51, 12–22, 2001.
30. Wendler, S.L. and Frazier, C.E., Effect of moisture content on the isocyanate/
wood adhesive bond line by N-15 CP/MAS NMR. J. Appl. Polym. Sci., 61,
775–782, 1996.
31. Wendler, S.L. and Frazier, C.E., The effects of cure temperature and time
on the isocyanate-wood adhesive bond line by N-15 CP/MAS NMR. Int. J.
Adhes. Adhes., 16, 179–186, 1996.
32. Rosthauser, J.W., Haider, K.W., Hunt, R.N., Chemistry of PMDI wood bind-
ers: Model studies. Proceedings, Washington State University International
Particleboard/Composite Materials Symposium, Washington State University,
Pullman, WA, vol. 31, pp. 161–175, 1997.
33. Harper, D.P., Wolcott, M.P., Rials, T.G., Evaluation of the cure kinetics of the
wood/pMDI bondline. Int. J. Adhes. Adhes., 21, 137–144, 2001.
34. He, G. and Yan, N., Effect of moisture content on curing kinetics of pMDI
resin and wood mixtures. Int. J. Adhes. Adhes., 25, 450–455, 2005.
35. Das, S., Malmberg, M.J., Frazier, C.E., Cure chemistry of wood/polymeric
isocyanate (PMDI) bonds: Effect of wood species. Int. J. Adhes. Adhes., 27,
250–257, 2007.
36. Riedlinger, D.A., Sun, N., Frazier, C.E., Tg as an index of conversion in
PMDI-impregnated woo. Bioresources, 2, 605– 615, 2007.
37. Dunky, M. and Pizzi, A., Wood adhesives, in: Adhesion Science and
Engineering, Vol. 2, Surfaces, Chemistry and Applications, M. Chaudhury and
A.V. Pocius (Eds.), pp. 1039–1103, Elsevier, Amsterdam, 2002.
38. Roll, H., Mikrotechnologische Untersuchungen zum Verhalten des Klebstoffs
polymeres Diphenylmethan-4, 49-diisocyanat auf Holzoberflachen in Hinblick
Polyurethane Adhesives 195

auf seine Verteilung in Spanplatten, PhD. Dissertation, University of Munich,


Germany, 1993.
39. Frazier, C.E., The interphase in bio-based composites: What is it, what
should it be? Proceedings 6th Pacific Rim Bio-based Composites Symposium,
Portland, Oregon, Oregon State University, Corvallis, Oregon, vol. 1,
pp. 206–212, 2002.
40. Saunders, J.H. and Slocombe, R.J., The chemistry of organic isocyanates.
Chem. Rev., 43, 203–218, 1948.
41. Tiger, R.P., Bekhli, L.S., Entelis, S.G., Kinetics of the limiting stage of interac-
tion between isocyanates and water. Vysokomol. Soedin. Ser., B, 11, 460–462,
1969.
42. Tiger, R.P., Bekhli, L.S., Entelis, S.G., The kinetics and mechanism of isocya-
nate hydrolysis. Kinet. Katal., 12, 318–325, 1971.
43. Bao, S.C., Daunch, W.A., Sun, Y.H., Rinaldi, P., Marcinko, J.J., Phanopoulos,
C., Solid state two-dimensional NMR studies of polymeric diphenylmethane
diisocyanate (PMDI) reaction in wood. For. Prod. J., 53, 63–71, 2003.
44. Yelle, D.J., Ralph, J., Frihart, C.R., Investigating the reactivity of pMDI
with wood cell walls using high-resolution–solid state NMR spectroscopy.
Proceedings 32nd Annual Meeting of the Anderson Society, G., Anderson (Ed.),
Savannah, GA, Feb. 15–18, 2009.
45. Pizzi, A. and Walton, T., Non-emulsifiable, water-based diisocyanate adhe-
sives for exterior plywood, Part 1: Novel reaction mechanisms and their
chemical evidence. Holzforschung, 46, 541–547, 1992.
46. Pizzi, A., Valenzuela, J., Westermeyer, C., Non-emulsifiables, water-based,
diisocyanate adhesives for exterior plywood, Part 2: Industrial application.
Holzforschung, 47, 69–72, 1993.
47. Pizzi, A., von Leyser, E.P., Valenzuela, J., Clark, J.G., The chemistry and devel-
opment of pine tannin adhesives for exterior particleboard. Holzforschung,
47, 164–172, 1993.
48. Basso, M.C., Pizzi, A., Lacoste, C., Tannin–furanic–polyurethane foams for
industrial continuous plant lines. Polymers, 6, 2985–3004, 2014.
49. Mansouri, H.R., Pizzi, A., Leban, J.M., Improved water resistance of UF
adhesives for plywood by small pMDI additions. Eur. J. Wood Prod., 64,
218–220, 2006.
50. Lei, H., Pizzi, A., Du, G., Coreacting PMUF/isocyanate resins for wood panel
adhesives. Eur. J. Wood Prod., 64, 117–120, 2006.
51. Dziurka, D. and Mirski, R., UF-pMDI hybrid resin for waterproof particle-
boards manufactured at a shortened pressing time. Drvna Industrija, 61,
245–249, 2010.
52. A. Pizzi, E.P. von Leyser, C. Westermeyer, Adhesive composition compris-
ing isocyanate, phenol–formaldehyde and tannin, useful for manufacturing
plywood for exterior application, Chile Patent 1084-90 (1991); US Patent
5,407,980, 1995.
196 Adhesives for Wood and Lignocellulosic Materials

53. Valenzuela, J., von Leyser, E.P., Pizzi A., Westermayer, C., Gorrini, B.,
Industrial production of pine tannin-bonded particleboard and MDF. Eur. J.
Wood Prod., 70, 735–740, 2012.
54. Dominguez-Rosado, E., Liggat, J.J., Snape, C.E., Eling, B., Pichteld, J.,
Thermal degradation of urethane modified polyisocyanurate foams based on
aliphatic and aromatic polyester polyol. Polym. Degrad. Stabil., 78, 1–5, 2002.
55. Chattopadhyay, D.K. and Webster, D.C., Thermal stability and flame retar-
dancy of polyurethanes. Prog. Polym. Sci., 34, 1068–1133, 2009.
56. Zhang, X., Wei S, S., Gao, Z., Moisture-curing polyurethane resins modi-
fied by multi-hydroxymethylated phenol. Pigm. Resin. Technol., 43, 2, 69–74,
2014.
57. Mumtaz, F., Zuber, M., Zia, K.M., Jamil, T., Hussain, R., Synthesis and prop-
erties of aqueous polyurethane dispersions: Influence of molecular weight of
polyethylene glycol. Korean J. Chem. Eng., 30, 2259–2263, 2013.
58. Remya, V.R., Patil, D., Abitha, V.K., Rane, A.V., Mishra, R.K., Biobased mate-
rials for polyurethane dispersions. Chem. Int., 2, 3, 158–167, 2016.
59. Galbraith, C.J., Cohen, S.C., Ball, G.W., Self-releasing emulsifiable MDI iso-
cyanate: An easy approach for all-isocyanate bonded boards. Proceedings,
17th International Particleboard/Composite Symposium, Pullman, WA,
Washington State University, pp. 263–282, 1983.
60. Jones, N., Optimising the properties of straw based particleboard, PhD
Dissertation, University of Wales, Bangor, UK, 1997.
61. Lay, D.G. and Cranley, P., Polyurethane adhesives, in: Handbook of Adhesive
Technology, 2nd edn., A. Pizzi and K.L. Mittal (Eds.), Marcel-Decker, New
York, 2003.
62. Papadopoulos, A.N., Hill, C.A.S., Traboulay, E., Hague, J.R.B., Isocyanate
resins for particleboard: PMDI vs EMDI. Holz Roh Werkst., 60, 81–83, 2002.
63. Grøstad, K. and Bredesen, R., EPI for glued laminated timber, in: Materials
and Joints in Timber Structures, vol. 9, RILEM Bookseries, S. Aicher, H.X.
Reinhardt, H. Garrecht (Eds.), 2014.
64. Grøstad, K. and Pedersen, A., Emulsion polymer isocyanates as wood adhe-
sive: A review. J. Adhes. Sci. Technol., 24, 1357–1381, 2010.
65. Japanese Industrial Standard, JIS K6806, Water based polymer-isocyanate
adhesives for wood, Japanese Standards Association, 1985.
66. Technische Kommission Holzklebstoffe, (TKH Technical Committee on Wood
Adhesives) of Industrieverband Klebstoffe e.V., EPI adhesives, Düsseldorf,
Germany, 2015.
67. Desroches, M., Escouvois, M., Auvergne, R., From vegetable oils to polyure-
thanes: Synthetic routes to polyols and main industrial products. Polym. Rev.,
52, 1, 38–134, 2012.
68. Nohra, B., Candy, L., Blanco, J.-F., Guerin, C., Raoult, Y., Mouloungui, Z.,
From petrochemical polyurethanes to biobased polyhydroxyurethanes, a
review. Macromolecules, 46, 3771–3792, 2013.
Polyurethane Adhesives 197

69. Desai, S.D., Patel, J.V., Sinha, K., Polyurethane adhesive system from
biomaterial-based polyol for bonding wood. Int. J. Adhes. Adhes., 23, 393–
399, 2003.
70. Li, A. and Li, K., Pressure-sensitive adhesives based on epoxidized soybean
oil and dicarboxylic acids. ACS Sustainable Chem. Eng., 2, 2090–2096, 2014.
71. Ahn, B.K., Kraft, S., Wang, D., Sun, S., Thermally stable, transparent,
pressure-sensitive adhesives from epoxidized and dihydroxyl soybean oil.
Biomacromolecules, 1, 1839–1843, 2011.
72. Pizzi, A., Tannin-based polyurethane adhesives. J. Appl. Polym. Sci., 23,
1889–1890, 1979.
73. Glasser, W.G., Barnett, C.A., Rials, T.G., Saraf, V.P., Engineering plastics from
lignin II. Characterization of hydroxyalkyl lignin derivatives. J. Appl. Polym.
Sci., 29, 1815–1830, 1984.
74. Glasser, W.G., Kelley, S.S., Rials, T.G., Structure–property relationships of
engineering plastics from lignin. Proceedings, 1986 TAPPI Research and
Development Conference, pp. 157–161, 1986.
75. Glasser, W.G. and Leitheiser, R.H., Engineering plastics from lignin. XI.
Hydroxypropyl lignins as components of fire resistant foams. Polym. Bull.,
12, 1–5, 1984.
76. Garcia, D., Glasser, W.G., Pizzi, A., Osorio-Madrazo, A., Laborie, M.-P.,
Hydroxypropyl tannin derivatives from Pinus pinaster (Ait.) bark. Ind. Crops
Prod., 49, 730–739, 2013.
77. Garcia, D., Glasser W.G., Pizzi, A., Paczkowski, S., Laborie, M.-P., Substitu-
tion pattern elucidation of hydroxypropyl Pinus pinaster (Ait.) bark poly-
flavonoids derivatives by ESI(–)-MS/MS. J. Mass Spectrom., 49, 1050–1058,
2014.
78. Garcia, D., Glasser, W.G., Pizzi, A., Paczkowski, S., Laborie, M.-P.,
Hydroxypropyl tannin from Pinus pinaster bark as polyol source in urethane
chemistry. Eur. Polym. J., 67, 152–165, 2015.
79. Garcia, D., Glasser, W.G., Pizzi, A., Paczkowski, S., Laborie, M.-P.,
Modification of condensed tannins: From polyphenols chemistry to materi-
als engineering. New J. Chem., 40, 36–49, 2016.
80. Rokicki, G. and Piotrowska, A., A new route to polyurethanes from ethylene
carbonate, diamines and diols. Polymer, 43, 2927–2935, 2002.
81. J.M. Whelan, Jr., M. Hill, R.J. Cotter, Multiple cyclic carbonate polymers. US
Patent 3,072,613, 1963.
82. Kihara, N. and Endo, T., Synthesis and properties of poly(hydroxyurethane)s.
J. Polym. Sci., Part A, Polym. Chem., 31, 2765–2773, 1993.
83. Kihara, N., Kushida, Y., Endo, T., Optically active poly(hydroxyurethane)s
derived from cyclic carbonate and L-lysine derivatives. J. Polym. Sci., Part A,
Polym. Chem., 34, 2173–2179, 1996.
84. Tomita, H., Sanda, F., Endo, T., Structural analysis of polyhydroxyurethane
obtained by polyaddition of bifunctional five-membered cyclic carbonate
198 Adhesives for Wood and Lignocellulosic Materials

and diamine based on the model reaction. J. Polym. Sci., Part A, Polym.
Chem., 39, 851–859, 2001.
85. Tomita, H., Sanda, F., Endo, T., Polyaddition behavior of bis(five- and
six-membered cyclic carbonate)s with diamine. J. Polym. Sci., Part A, Polym.
Chem., 39, 860–867, 2001.
86. Tomita, H., Sanda, F., Endo, T., Model reaction for the synthesis of poly-
hydroxyurethanes from cyclic carbonates with amines: Substituent effect
on the reactivity and selectivity of ring-opening direction in the reaction of
five-membered cyclic carbonates with amine. J. Polym. Sci., Part A, Polym.
Chem., 39, 3678–3685, 2001.
87. Birukov, O., Potashnikova, R., Leykin, A., Figovsky, O., Shapovalov, L.,
Advantages in chemistry and technology of non-isocyanate polyurethane.
J. Sci. Israel-Technol. Adv., 11, 160–167, 2009.
88. Figovsky, O. and Shapovalov, L., Features of reaction amino-cyclocarbonate
for production of new type polyurethanes. Macromol. Symp., 187, 325–332,
2002.
89. Camara, F., Benyahya, S., Besse, B., Boutevin, G., Auvergne, R., Boutevin, B.,
Caillol, S., Reactivity of secondary amines for the synthesis of nonisocyanate
polyurethanes. Eur. Polym. J., 55, 17–26, 2014.
90. Blattmann, H., Fleischer, M., Bähr, M., Mulhaupt, L., Isocyanate- and
phosgene-free routes to polyfunctional cyclic carbonates and green polyure-
thanes by fixation of carbon dioxide. Macromol. Rapid Comm., 35, 1238–
1254, 2014.
91. Boyer, A., Cloutet, E., Tassaing, T., Gadenne, B., Alfos, C., Cramail, H.,
Solubility in CO2 and carbonation studies of epoxidized fatty acid diesters:
Towards novel precursors for polyurethane synthesis. Green Chem., 12,
2205–2213, 2010.
92. Kim, M.-R., Kim, H.-S., Ha, C.-S., Park, D.-W., Lee, J.-K., Syntheses and
thermal properties of poly(hydroxy)urethanes by polyaddition reaction of
bis(cyclic carbonate) and diamines. J. Appl. Polym. Sci., 81, 2735–2743, 2001.
93. Ochiai, B., Inoue, S., Endo, T., Salt effect on polyaddition of bifunctional
cyclic carbonate and diamine. J. Polym. Sci., Part A, Polym. Chem., 43, 6282–
6286, 2005.
94. Ubaghs, L., Fricke, N., Keul, H., Höcker, H., Polyurethanes with pendant
hydroxyl groups: Synthesis and characterization. Macromol. Rapid Comm.,
25, 517–521, 2004.
95. Fleischer, F., Blattmann, H., Mülhaupt, R., Glycerol-, pentaerythritol- and
trimethylolpropane-based polyurethanes and their cellulose carbonate com-
posites prepared via the non-isocyanate route with catalytic carbon dioxide
fixation. Green Chem., 15, 934–942, 2013.
96. Thebault, M., Pizzi, A., Dumarcay, S., Gerardin, P., Fredon, E., Delmotte, L.,
Polyurethanes from hydrolysable tannins obtained without using isocya-
nates. Ind. Crops Prod., 59, 329–336, 2014.
Polyurethane Adhesives 199

97. Thebault, M., Pizzi, A., Essawy, H., Baroum, A., van Assche, G., Isocyanate-
free condensed tannin-based polyurethanes. Eur. Polym. J., 67, 513–523,
2015.
98. Braghiroli, F., Fierro, V., Pizzi, A., Rode, K., Radke, W., Delmotte, L.,
Parmentier, J., Celzard, A., Condensation reaction of flavonoid tannins with
ammonia. Ind. Crops Prod., 44, 330–335, 2013.
99. Thebault, M., Pizzi, A., Santiago-Medina, F.J., Al-Marzouki, F.M., Abdalla,
S., Isocyanate-free polyurethanes by coreaction of condensed tannins with
aminated tannins. J. Renew. Mat., 51, 21–29, 2017.
100. Xi, X., Pizzi, A., Delmotte, L., Isocyanate-free polyurethane coatings and
adhesives from mono- and di-saccharides. Polymers, 10, 4, 402, 2–21, 2018.
101. Xi, X., Wu, Z., Pizzi, A., Gerardin, C., Lei, H., Zhang, B., Du, G., Non-
isocyanate polyurethane adhesive from sucrose used for particleboard. Wood
Sci. Technol., 53, 393–405, 2019.
102. Xi, X., Pizzi, A., Gerardin, C., Du, G., Glucose-based non-isocyanate poly-
urethane biofoams. J. Renew. Mat., 7, 301–312, 2019.
8
Wood Surface Inactivation (Thermal)

8.1 Introduction
The 1999 USDA Wood Handbook [1] defines surface inactivation as
“physical and chemical modifications of the wood surface that result in
reduced ability of an adhesive to properly wet, flow, penetrate, and cure.”
Accordingly, wood surface inactivation results in a poor bonding
strength. It is a time-dependent process and can be accelerated by increas-
ing temperature. Wood surface inactivation was originally detected when
difficulties were encountered in secondary bonding plywood surfaces after
hot pressing. De Bruyne attributed the inactivation phenomenon to reac-
tion between the pairs of cellulosic hydroxyl groups when a plywood sur-
face is exposed to high temperature (in the hot press) to produce water and
ether linkages [2]. Ether bonds are relatively more hydrophobic compared
to the hydroxyl groups.
De Bruyne’s explanation was based on the earlier hypothesis of Stamm
and Hansen that wood was heat stabilized though the formation of ether
bonds [3]. Later, Seborg et al. showed that cross-linking by ether bond for-
mation was not a good explanation for thermal stabilization [4]. Although
wood was stabilized against swelling by water, it was not stabilized against
pyridine or sodium hydroxide, none of which would break ether bonds.
The subject of inactivation of wood surface was discussed extensively by
Christiansen in two separate review articles [5, 6]. Part I deals with inacti-
vation of wood surfaces due to excessive drying (overdrying) to bond with
phenolic adhesives [5]. Three mechanisms were proposed: (1) migration of
extractives to the surface, which lowers the wettability or hides the surface;
(2) reorientation of wood surface molecules, which reduces wettability or
bonding sites, and (3) irreversible closure of large micropores in the wood
cell walls.
Part II of the review explains how chemical reactions related to over-
drying may cause inactivation of wood surfaces to bonding [6].

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (201–210) © 2019
Scrivener Publishing LLC

201
202 Adhesives for Wood and Lignocellulosic Materials

An American Plywood Association publication states that overdrying


of veneers can be due to the following reasons [7]: (1) drying temperature
can be too high, (2) drying time can be too long, (3) dampers open too
far (recycling less heated air), (4) low dryer humidity, and (5) low veneer
moisture content.
Hancock seems to have first used the term inactivation [8]. Previously,
Northcott et al. had defined casehardening as “a change in surface con-
dition of the wood, other than a coating from external source…, that is
induced or intensified by manufacturing process and that makes the sur-
face difficult to bond” [9].
Northcott et al. indicated a number of conditions that add to the prob-
lem, most of these involving insufficient resin moisture penetration into
the wood during the assembly period [10].
The problem due to Inactivation can be resolved by removing the sur-
face layer, for example, by sanding, so that the fresh surface is exposed [8,
9, 11–13]. However, some reports suggest that inactivation can extend to
a few layers below the surface [13]. Removal of inactivated surface of full-
size veneers will not be economically feasible. It would be impossible for
flake furnishes.

8.2 Causes and Sources of Inactivation


As mentioned above, the inactivation process is caused by the change of
the primarily hydrophilic wood surface into a hydrophobic wood surface.
This can be due to either migration of extractives to the surface or lig-
nin concentration and rearrangement at the surface. Both extractives and
lignin are hydrophobic relative to cellulose and hemicelluloses. Nonpolar,
hydrocarbon type of extractives should have the highest impact on severity
of surface inactivation.
Besides the causes of inactivation of wood surface as discussed earlier,
other sources of inactivation are air, light, heat, chemical treatments, and
machining [14, 15]. Typical processes that lead to wood surface inactiva-
tion are wood aging, wood weathering, wood seasoning, and wood heating
or drying. However, drying of veneers at high temperature, i.e., thermal
inactivation, is typically the most significant in the wood-based compos-
ite industry. Heat-induced inactivation comprises (1) primary inactivation
and (2) secondary inactivation.
The primary wood surface inactivation occurs when wood is exposed to
heat for the first time, e.g., wood drying. That includes kiln drying, drum
drying, jet drying, and hot-platen drying. The secondary wood surface
Wood Surface Inactivation (Thermal) 203

inactivation is caused during subsequent heat treatment, such as hot-


pressing a composite to cure the adhesive.
Surface inactivation can occur either at low temperature for long
duration (i.e., aging) or in short duration at high temperature. However,
high-temperature inactivation is reported to be more severe than aging.
According to Christiansen, physical and chemical inactivation mechanisms
can reduce the attractive forces on the wood surface, which were initially
available for bonding with the adhesive [6]. Each of the inactivation mech-
anisms can operate independently as well as function simultaneously [16].

8.3 Mechanisms of Inactivation


Christiansen reviewed the mechanisms that are most often proposed for
wood surface inactivation [5, 6]. The inactivation mechanisms involving
physical phenomena are (1) extractives-related non-wetting, (2) surface
molecular reorientation, and (3) micropore closure. A fourth possible
mechanism, which is seldom a problem, presents contamination by soot
or other airborne deposits [5]. The inactivation mechanisms involving
chemical phenomena include (1) elimination of surface hydroxyl bonding
sites by ether formation, (2) oxidation and/or pyrolysis of surface bonding
sites, and (3) chemical interference with the adhesive curing reaction [6].
Some other inactivation mechanism, especially operating in the case of
paper fibres, namely, “hornification”, has also been proposed. Hornification
is the increase in the degree of cross-linking between microfibrils due to
additional hydrogen bonds formed during drying and not broken during
rewetting [17].

8.4 Factors Affecting Wood Surface Inactivation


8.4.1 Effect of Wood Species
Some wood species are more susceptible to surface inactivation than others
[7, 18]. Wood-related factors influencing the inactivation are wood anatomy,
wood chemistry, and wood moisture content. A majority of the reported
inactivation problems are related to softwood species [5]. Softwoods are
usually more susceptible to inactivation than hardwoods. The temperature
at which inactivation occurs is also dependent on wood species.
Erb [18] and Christiansen [5] report that the maximum safe drying tem-
peratures for avoiding inactivation in several softwoods vary with species.
204 Adhesives for Wood and Lignocellulosic Materials

Southern pines were the most susceptible to inactivation, followed by


ponderosa pine, inland Douglas-fir, western white pine, larch, and coastal
Douglas-fir [15].
Heartwood and sapwood possess different susceptibility to inactivation
because

(1) Heartwood usually has a lower MC than sapwood [19].


(2) Heartwood contains a higher proportion of hydrophobic
extractives than sapwood [20]. Kajita and Skaar attributed
the greater wettability of sapwood compared with heart-
wood to the higher extractive content of the heartwood
[21]. Extractives’ deposition can also block the pit open-
ings between the wood cells. This reduces wood perme-
ability and prevents penetration of the adhesive into the
wood cellular structure.

Wood extractives tend to be a dominating factor in the inactivation of


the wood surface. However, all the chemical components of the wood con-
tribute to its surface chemistry [5, 22], and therefore, surface inactivation
can originate from different wood constituents, not from extractives alone.
Extractives’ deposition can also block the pit openings between the
wood cells. This reduces wood permeability and prevents penetration of
the adhesive into the wood cellular structure. The role of extractives in the
inactivation of wood surface is dealt with in detail in Chapter 9.
Under a given set of conditions of drying, heartwood dries more easily
than sapwood because it usually has a lower MC than sapwood [19]. Thus,
heartwood and sapwood possess different susceptibilities to inactivation.
Additionally, heartwood contains a higher proportion of the extractives
than sapwood [20]. The extractives can affect wood wettability and adhe-
sive spreading. Most of the extractives are hydrophobic in character; thus,
they interfere with wetting of the aqueous adhesives. Kajita and Skaar
attributed the greater wettability of sapwood compared with heartwood
to the higher extractive content of the heartwood [21]. Extractives’ depo-
sition can also block the pit openings between the wood cells. This reduces
wood permeability and prevents penetration of the adhesive into the wood
cellular structure.

8.4.2 Inactivation Due to High-Temperature Drying


As mentioned earlier, wood inactivation is a time-dependent process
accelerated by increasing temperature. Surface inactivation can occur
Wood Surface Inactivation (Thermal) 205

either at low temperature for a long time (i.e., aging) or in a short time at
high temperature. However, high temperature causes more severe inacti-
vation than aging.
The mechanism of inactivation changes with temperature. For exam-
ple, the inactivation by pyrolysis begins at temperatures >270°C [23],
while the inactivation due to migration of extractives occurs even at room
temperature. For the most sensitive American coniferous species, signifi-
cant wood surface inactivation occurs at the drying temperature of 160°C
and higher [5].
Even though wood surface inactivation can occur by a number of pro-
cesses as described above, the most significant inactivation problems are
associated with drying of veneer at high temperature. Veneer surface tem-
perature changes during drying. Initially, the wood surface is at the ambient
temperature. As wood dries, water starts evaporating and the water vapor
is transferred to the drying medium [24]. At some point, a steady state is
reached at which the rate of evaporation of water from the wood surface is
the same as the rate of capillary flow of water from the bulk of wood to the
wood surface. In this case, veneer surface temperature is lower (at wet bulb
temperature) than air temperature because of evaporative cooling. As the
MC decreases and falls below the FSP, wood contains only bound water.
This water is held more strongly to wood by hydrogen bonding and there-
fore the rate of flow/diffusion of water from the bulk of wood to the surface
is slower than the rate of evaporation of water from the surface. The evap-
orative cooling effect decreases and the surface temperature starts to climb
to temperatures near that of the air in the dryer (dry-bulb temperature)
[5]. This is the stage when typical wood surface inactivation occurs [25].

8.4.2.1 Effect of Drying Technique


The drying technique affects inactivation because drying parameters (e.g.,
air temperature, velocity, and direction) vary with the technique. Shupe et al.
found that contact angle was higher on oven-dried and air-dried wood sur-
faces than on freeze-dried wood surfaces [20]. Therefore, low-temperature
drying diminishes or eliminates wood surface inactivation.
In jet dryers, the high air velocity allows for faster water and heat trans-
port across the air–wood interface. The increased surface drying rates may
increase the chances for surface inactivation before the veneer is com-
pletely dried [5].
In a platen drying process, wood veneer is placed between two hot
plates. The contact between the plates and wood surface may influence
wood surface properties during the drying. The aluminum plate provides
206 Adhesives for Wood and Lignocellulosic Materials

an aluminum oxide surface and also excludes oxygen over much of the
surface. A few researchers investigated the effect of this drying method
on wettability and bondability of wood. Kadlec showed that the wet-
tability of Douglas-fir veneer generally decreased with increased plate
temperature [26].

8.5 Physical Mechanisms of Inactivation


8.5.1 Effect of Extractives on Wettability and Adhesion
The change in the wettability of wood as a result of change of tempera-
ture and time has been attributed to the migration of extractives to the
surface [5]. Podgorski et al. reported that wettability and adhesion were
adversely affected after thermal treatment of wood [27]. The extractable
compounds were responsible for poor wettability and adhesion. Gray
evaluated advancing and receding contact angles for 19 wood species
[28]. Sanding the surfaces of specimens produced lower contact angles,
i.e., improved wetting. Changes in contact angles were attributed to sur-
face contamination by low-molecular-weight fatty acids, high extractives
content, and high resin content. Wood extractives are polar and nonpolar
[23]. Nonpolar extractives are primarily responsible for low wettability of a
wood surface by waterborne adhesives.

8.5.2 Molecular Reorientation at Surfaces


It is known that synthetic polymer molecules on a surface can reorient
themselves to present a low energy (hydrophobic) surface against air
[5]. Wood surfaces consist of three natural polymers: cellulose, hemi-
celluloses, and lignin. Polymer surfaces are time-, temperature-, and
environment-dependent [15, 29]. The driving force for the molecular
reorientation is thermodynamic in nature, with the surface tending to
minimize its free energy. Amorphous and glassy polymers, such as hemi-
cellulose and lignin in wood, are not in thermodynamic equilibrium
[29]. If molecular motions are possible, glassy polymers may rearrange
to minimize surface free energy.
Surface reorientation can occur during the aging process in which
surface wettability is reduced. At high temperatures, reorientation
and other molecular movements are accelerated; thereby, faster for-
mation of a hydrophobic surface is facilitated. This is particularly pro-
nounced when the moisture contents are high enough to reduce the glass
Wood Surface Inactivation (Thermal) 207

transition temperatures of hemicelluloses and lignin so that reorienta-


tion is promoted at the temperature of exposure. The glass transition of
these two amorphous polymers strongly depends on moisture content.
Hemicelluloses have a glass transition temperature between −23 and
200°C [30], while lignin in softwoods and hardwoods has glass transi-
tion in the range of 65–85°C and 90–105°C, respectively [31]. Therefore,
structural rearrangement of the amorphous part of the wood surface can
occur when drying wood or curing wood-based composites. Compared
to extractives migration, molecular rearrangements at the wood surface
cause smaller changes in hydrophobicity than nonpolar extractives.

8.5.3 Micropore Closure


Many micropores between the lamellae of the cell wall are lost during
the drying process [5]. There is a loss of porosity as a result of increasing
drying temperature. One result of this is that the sorption and diffusion
properties of wood surfaces decrease after thermal exposure. The loss of
porosity is equivalent to micropore closure. Micropore closure affects the
adhesive penetration and wetting of the wood cell walls. Micropore clo-
sure as a result of drying has an effect for wetting; for example, Wellons
has shown that it is harder to wet wood that is at a lower moisture content
[32]. If anchoring of resins is important, as Nearn has stated for phenolic
adhesives [33], the closure of larger micropores should limit penetration
by larger resin molecules and lessen the bond strength and wood failure.

8.6 Chemical Mechanisms of Inactivation


8.6.1 Elimination of Surface Hydroxyl Bonding Sites
As stated earlier, the inactivation of wood surfaces was first noted in occa-
sional problems of bonding to plywood surfaces after hot pressing. The
original hypothesis put forward by De Bruyne for the mechanism of inac-
tivation was that water was eliminated from between pairs of cellulose
hydroxyl groups to form ether bonds [2].
Ether bonds are not as receptive as the original hydroxyl groups to
hydrogen bond formation with polar adhesives. De Bruyne’s hypothesis
was based on the earlier hypothesis of Stamm and Hansen that wood was
heat stabilized though the formation of ether bonds [3]. Later, Seborg et al.
[4] showed that cross-linking by ether bond formation was not a good
explanation for thermal stabilization; although wood was stabilized against
208 Adhesives for Wood and Lignocellulosic Materials

swelling by water, it was not stabilized against pyridine or sodium hydrox-


ide, none of which would break ether bonds [6].

8.6.2 Oxidation and/or Pyrolysis of Surface Bonding Sites


Oxidation and pyrolysis are important processes that occur at high tem-
peratures and long duration of exposures. Increasing temperature accel-
erates the rates of oxidation and pyrolysis reactions, and obviously, at
high temperatures, the time for degradation becomes shorter. At very
high temperatures, the hemicelluloses may be changed to furfural poly-
mers, which are less hygroscopic [34]. Also, moisture content increases
the depolymerization processes of wood constituents at high temperatures
[35]. Oxidation and pyrolysis were proposed as a prime cause of surface
inactivation for white spruce veneer [19]. Hemingway concluded that the
reduced wettability of yellow birchwood might be related to the oxidation
of some fatty acids [36].

8.6.3 Chemical Interference with Resin Cure or Bonding


The alkaline or acidic nature of the wood surface could impede bonding by
interfering with the cure reaction of the resin. The curing of adhesives could
be retarded or accelerated due to change in pH value of the wood surface.
The curing problem is more likely associated with species that have a high
amount of acid extractives such as tropical hardwood species, pine, and
oak. The acidity of oak surfaces significantly reduced the bond strength of
resorcinol adhesives [37]. Also, extractives often modify the cure of phe-
nolic adhesives [38]. The acidic extractives of oak and kapur prolonged the
curing of phenolic adhesives [39]. A low pH of extractives concentrated on
the wood surface accelerates curing of acid-catalyzed urea–formaldehyde
adhesives with the danger of pre-curing in the hot press.

References
1. United Stated Department of Agriculture (USDA), in: Wood Handbook:
Wood as an Engineering Material, USDA and Forest Products Society,
Madison, Wisconsin, 1999.
2. De Bruyne, A., The nature of adhesion. Suppl. Flight-The Aircraft Eng., 28,
51–54, 1939.
3. Stamm, A.J. and Hansen, A., Minimizing wood shrinkage and swelling.
Effect of heating in various gases. Ind. Eng. Chem., 29, 831–833, 1937.
Wood Surface Inactivation (Thermal) 209

4. Seborg, R., Tarkowa, H., Stamm, A.J., Effect of heat upon the dimensional
stabilization of wood. J. Forest Prod. Res. Soc., 3, 3, 59–67, 1953.
5. Christiansen, A.W., How overdrying wood reduces its bonding to phenol–
formaldehyde adhesives: A critical review of the literature. Part I. Physical
responses. Wood Fiber Sci., 22, 441–459, 1990.
6. Christiansen, A.W., How overdrying wood reduces its bonding to phenol–
formaldehyde adhesives: A critical review of the literature. Part II. Chemical
reactions. Wood Fiber Sci., 23, 69–84, 1991.
7. Erb, C., Dryers and veneer drying, Douglas Fir Plywood Association Technical
Association Report 112, Part I, p. 13, American Plywood Association, Tacoma,
WA, 1975.
8. Hancock, W.V., Effect of heat treatment on the surface of veneer. Forest
Prod. J., 13, 81–88, 1963.
9. Northcott, P.L., Colbeck, H.G.M., Han, W.V., Shen, K.C., Undercure.
Casehardening in plywood. Forest Prod. J., 9, 442–451, 1959.
10. Northcott, P.L., Hancock, W.V., Colbech, H.G.M., Water relations in pheno-
lic (plywood) bonds. Forest Prod. J., 12, 478–486, 1962.
11. Kaufert, F.H., Preliminary experiments to improve the gluing characteristics of
refractory plywood surfaces by sanding, p. 9, U.S. Department of Agriculture,
Forest Service, Forest Products Laboratory, Madison, WI, Report 1351, 1943.
12. Hancock, W.V., The influence of native fatty acids on the formation of glue
bonds with heat-treated wood, PhD Dissertation, University of British
Columbia, Vancouver, Canada, 1964.
13. Walters, E.O., The effects of green veneer water content dryer schedules, and
wettability on gluing results for southern pine veneer. Forest Prod. J., 23, 6,
46–53, 1973.
14. Marra, A.A., Technology of Wood Bonding: Principles in Practice, Van
Nostrand Reinhold, New York, 1992.
15. Sernek, M., Comparative Analysis of Inactivated Wood Surfaces. PhD Dissertation,
Virginia Polytechnic Institute and State University, Blacksburg, VI, 2002.
16. Carpenter, M.W., Characterizing the Chemistry of Yellow-poplar Surfaces
Exposed to Different Surface Energy Environments Using DCA, DSC and XPS,
PhD Dissertation, West Virginia University, Morgantown, WV, 1999.
17. Minor, J.L., Hornification—Its origin and meaning. Prog. Pap. Recycl., 3, 2,
93–95, 1994.
18. Erb, C., Effect of over drying on the gluability of softwood veneer, Laboratory
Report 103, p. 13, American Plywood Association, Tacoma, WA, 1965.
19. Troughton, G.E. and Chow, S.Z., Migration of fatty acids to white spruce
veneer surface during drying. Relevance to theories of inactivation. Wood
Sci., 3, 129–133, 1971.
20. Shupe, T.F., Hse, C.-Y., Wang, W.H., An investigation of selected factors that
influence hardwood wettability. Holzforschung, 55, 541–548, 2001.
21. Kajita, H. and Skaar, C., Wettability of the surfaces of some American soft-
woods. Mokuzai Gakkaishi, 38, 516–521, 1992.
210 Adhesives for Wood and Lignocellulosic Materials

22. Gardner, D.J., Wolcott, M.P., Wilson, L., Huang, Y., Our understanding of
wood surface chemistry, in: Proceedings Wood Adhesives 1995, pp. 29–36,
Forest Products Society, Madison, WI, 1995.
23. Fengel, D. and Wegener, G., Wood: Chemistry, Ultrastructure, Reactions,
Walter de Gruyter, Berlin, 1989.
24. Siau, J.F., Wood—Influence of moisture on physical properties, Dept. of
Wood Science and Forest Products, Virginia Polytechnic Institute and State
University, 1995.
25. Suchsland, O. and Stevens, R.R., Gluability of southern pine veneer dried at
high temperatures. Forest Prod. J., 18, l, 38–42, 1968.
26. Kadlec, K.M., Predicting Surface Inactivation after Platen Drying of Second-
Growth Douglas-Fir Veneer, PhD Dissertation, Oregon State University,
Corvallis, Oregon, 1980.
27. Podgorski, L., Chevet, B., Onic, L.L., Merlin, A., Modification of wood wet-
tability by plasma and corona treatments. Int. J. Adhes. Adhes., 20, 103–111,
2000.
28. Gray, V.R., The wettability of wood. Forest Prod. J., 12, 452–461, 1962.
29. Gunnells, D.W., Gardner, D.J., Wolcott, M.P., Temperature dependence of
wood surface energy. Wood Fiber Sci., 26, 447–455, 1994.
30. Kelley, S.S., Rials, T.G., Glasser, W.G., Relaxation Behavior of the Amorphous
Components of Wood, pp. 617–624, Chapman and Hall, London, 1987.
31. Glasser, W.G., Classification of lignin according to chemical and molecular
structure, in: Lignin: Historical, Biological, and Materials Perspectives, W.G.
Glasser, R.A. Northey, T.P. Schultz (Eds.), pp. 216–238, American Chemical
Society, Symposium Series 742, Washington, DC, 2000.
32. Wellons, J.D., Wettability and gluability of Douglas-fir veneer. Forest Prod. J.,
30, 7, 53–55, 1980.
33. Nearn, W.T., Application of the ultrastructure concept in industrial wood
products research. Wood Sci., 6, 285–293, 1974.
34. Hillis, W.E., High temperature and chemical effects on wood stability. Wood
Sci. Technol., 18, 281–293, 1984.
35. Zavarin, E., Activation of wood surface and nonconventional bonding, in:
The Chemistry of Solid Wood, R.M. Rowell (Ed.), pp. 349–400, American
Chemical Society, Washington, DC, 1984.
36. Hemingway, R.W., Thermal instability of fats relative to surface wettability of
yellow birchwood (Betula lutea). Tappi, 52, 2149–2155, 1969.
37. Subramanian, R.V., Chemistry of adhesion, in: The Chemistry of Solid Wood,
R.M. Rowell (Ed.), pp. 323–348, American Chemical Society, Washington,
DC, 1984.
38. Wellons, J.D., Adhesion to wood substrates, in: Wood Technology: Chemical
Aspects, R.F. Gould (Ed.), ACS Symposium Series, pp. 150–168, American
Chemical Society, Washington, DC, 1977.
39. Hse, C.-Y. and Kuo, M.-L., Influence of extractives on wood gluing and
finishing—A review. Forest Prod. J., 38, 52–56, 1988.
9
Wood Surface Inactivation
Due to Extractives

9.1 Introduction
One of the important differences between wood and other adherends is
the presence of extractives. They are common and important surface con-
taminants harmful to wood adhesion. Water-soluble extractives are trans-
ported to the wood surface along with water during the drying operation
and are deposited as solids when the water evaporates [1]. With regard to
wood as an adherend, the extractives are extremely important because of
their often undesirable and unpredictable effect upon adhesive bonding.
Innumerable opportunities exist for chemical reactions between extra-
neous materials and the atmosphere on the one hand, and between these
materials and adhesives on the other. Both these reactions can happen at
the wood–adhesive interphase [2].
These extraneous materials are not cell wall polymers. They are organic
or mineral substances found in the cell wall and cell lumen. These mate-
rials usually account for up to 5% of the dry weight of unextracted wood.
However, in some species or in certain locations within the tree, they may
constitute as much as or more than 30% of the weight of the wood [2].

9.2 Migration of Extractives to the Wood Surface


Extractives of wood have a profound effect on the adhesion. During pro-
cessing of wood such as kiln drying, extractives can diffuse from inside
the wood to the surface, thus altering the surface characteristics in a
manner that can deactivate the wood surface and interfere with bonding.
This migration phenomenon has been reported by several investigators.
Anderson and Fearing [3] found that the surface of kiln dried redwood
lumber contained five times the water-soluble extractives on the surface

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (211–222) ©  2019
Scrivener Publishing LLC

211
212 Adhesives for Wood and Lignocellulosic Materials

compared with green lumber. Accumulation of water-soluble extractives


was found on the wood surface during the high-temperature drying of
white oak [4]. Similarly, vapors of nonpolar extractives could migrate to the
surface during high-temperature drying process. Swanson and Cordingly
[5] reported that the surface characteristics of wood pulp significantly
changed to a less polar and more of a hydrophobic nature due to vapor
migration of resinous materials. Such migration of extractives to the wood
surface as a result of drying often causes deleterious effects on gluing.
However, Hancock [6] found that the removal of extractives in Douglas-
fir veneer prior to drying prevented vapor migration of nonpolar extractives.

9.3 Influence of Extractives on Bonding Properties


of Wood
As mentioned in the previous section, the surface characteristics of wood
can be altered by the migration of extractives from inside wood to the sur-
face and cause deleterious effect to adhesive bonding.
Hse and Kuo [1] identified the following reasons as to how extractives
can influence wood gluing.

(1) The polarity of wood surface can be severely changed in a


manner that will interfere with the wettability of the sur-
face by the adhesive.
(2) Permeability of adhesives into wood can be affected by the
extractives.
(3) Similarly, the heavy deposits of extractives in the wood
pores can function as a barrier to the penetration of adhe-
sives into the pores.

Both factors 2 and 3 will prevent the mechanical adhesion, which is very
important in wood adhesion. Finally, chemical incompatibility between
extractives and adhesives not only can affect the normal flow and spread-
ing of the adhesive but also chemically interfere with curing and setting
characteristics of adhesives.
The abovementioned effects may act individually or, more often, act
in combination. The chemical structures of extractives are such that they
can undergo several chemical reactions with atmosphere as well as adhe-
sives at the wood surface. The pH and buffering capacity of wood can be
affected by the type and amount of extractives. This in turn can affect the
rate of curing of the adhesives; i.e., increased rate of cure of the adhesive
Wood Surface Inactivation Due to Extractives 213

may result in precuring of the adhesive, which will lead to delamination.


The reactions triggered by the extractives will also affect the production
of wood products. Thus, wood species that are very acidic, such as oaks,
Douglas-fir, and kapur, are sometimes difficult to bond with adhesives that
are sensitive to extractives [2].
Thomas [7] studied the gluing characteristics of Determa (Ocotea rubra
Mez.), a tropical hardwood, and found that removal of the ether and
benzene solubles caused a considerable increase in the glue bond quality
obtained with a phenolic resin. The result indicated the chemical incom-
patibility between the nonpolar extractives present in unextracted wood
and the phenolic resin. The inability of the adhesive to wet the unextracted
wood and to secure adequate penetration was the cause of poor adhesion.
Extractives of high acidity accelerate the curing of acid curable UF res-
ins, and this may lead to the precure of the UF resin and the consequent
delamination of the bonded wood product produced there from unless
adequately lesser amount of hardener is added to the resin to compensate
for the excess acidity caused by the acidic extractives. Extractives of high
acidity adversely affect the cure of alkaline hardening PF resins. Water-
soluble extractives like free sugars are detrimental for bonding of wood
with cement. Polyphenolic extractives (tannins) can be used as a binder in
the wood-based industry but can also strongly affect the reactivity of phe-
nolic resins leading to their precuring. However pre-pressing of the glued
veneers before hot pressing can help in reducing this problem. Extractives
in wood can react with formaldehyde and reduce the formaldehyde emis-
sion of wood-based panels. Moreover, some wood extractives are volatile
organic compounds (VOC) and insofar also relevant to the emission of
VOC from wood and wood-based panels [8].
Extractives of high acidity accelerate the curing of acid curing UF resins
and decelerate bonding with alkaline hardening PF resins.
Since the extractives alter the polarity of the wood surface, the wettabil-
ity of the surface is accordingly affected. If the extractives are removed by
solvent extraction, the surface wettability may be regained. This method
has been successfully employed by Chen [9]. Wettability of many tropi-
cal woods could be improved by removing extractives from wood surfaces
with various solvents. Yellow birch is known to have poor wettability. This
is due to the oxidation of linoleic acid and its esters on the wood sur-
face [10]. It is well known (also discussed in Chapter 2, Fundamentals of
Adhesion) that efficient wetting of the substrate by the adhesive is imper-
ative for achieving good bond strength. Wood surface is hydrophilic and
the general adhesives used for bonding wood are water based. If nonpolar
extractives migrate to the wood surface as a result of thermal processing,
214 Adhesives for Wood and Lignocellulosic Materials

the wood surface becomes hydrophobic. Thus, water-repellent extractives


affect the surface tension of wood surfaces and water-based adhesives can-
not effectively wet the surface. This aspect has been studied by Hillis [11]
who summarized the influence of nonpolar extractives on gluing by point-
ing out that water-repellent extractives affect the surface tension of wood
surfaces inhibiting the wetting and penetration of adhesives, which subse-
quently affect bond strength. Resinous species, such as southern pine when
subjected to high-temperature drying process, concentrate and physically
block adhesive contact with wood. Thus, a barrier is created between wood
and the adhesive. The finding of Plomley [12] confirms this. Hoop pine
(Araucaria cunninghamii Hook) was dipped into solutions of commercial
tannins, crude wood extracts, and model compounds before bonding with
a phenol–formaldehyde resin adhesive. The bond quality was significantly
reduced. The formation of a physical barrier by the added compounds pre-
vented a good contact of the adhesive with the wood surface, resulting in a
reduced cohesive strength of the adhesive.
Further, the effect of extractives on wood adhesion depends on whether
the extractive is soluble or insoluble in the adhesive. Extractives that are
insoluble in the resin caused more adhesion problems than the extractives
that are soluble. Narayanamurti, Gupta, and Verna [13] found that the
extractives of teak (Tectona grandis) that are insoluble in water but solu-
ble in alcohol/benzene adversely affected the setting of water-based animal
and urea–formaldehyde adhesives. In contrast, the extractives of acacia that
are soluble in hot water did not interfere with either animal glue or urea–
formaldehyde. A better understanding of the manner in which extractives
affect adhesive bonding was possible by the investigation of Plomley [14]
through their studies on the effect of different commercial tannins, crude
wood extracts, and their components on bonding of phenol–formaldehyde
and resorcinol–formaldehyde adhesives on a number of wood species.
Plomley [12, 14] found that contamination of the bonding interface with
hydrolyzable tannins significantly reduced the bond quality of a phenol–
formaldehyde resin adhesive. Removal of a large part of the ether- and
benzene-soluble portion of the wax present in a tropical wood determa
(Ocotea rubra Mez) caused a considerable increase in the glue bond quality
obtained with a phenolic resin [7]. The result indicated chemical incom-
patibility between the wax and the phenol–formaldehyde resin.
In order to find the solution to the problem of interference of extractives
with the adhesion phenomenon in wood, it would be desirable to iden-
tify the specific components of extractives that are responsible for such
an interference. Research efforts have therefore been made in this direc-
tion. However, chemical complexity of the extractives together with the
Wood Surface Inactivation Due to Extractives 215

involvement of many operating variables in the gluing process make it dif-


ficult to define the compound or compounds that may cause inactivation
[1]. Similar problems due to extractives are encountered in the case of
the surface coatings on wood. An interesting test ascribed to Sanderman
et al. [15] for establishing the compatibility between the surface coating
and various extractives components has been described by Gardner [16].
A paper chromatogram of a solution of the extractives is first made and
then coated with the surface coating. The spots of the chromatogram
are observed for signs of failure. Specific types of interference, such as
interference with drying, discoloration, or cracking, can then be associ-
ated with the corresponding types of extractive components. The same
technique might be used to detect effects of extractives upon the curing
behavior of adhesives.
The curing of phenolic resins at the gluing surface was found to be
affected by a high concentration of extractives [17] in the following man-
ner. The acidity of extractives may reduce glueline pH, and this reduction
in pH requires a significant increase in press time to obtain a proper cure
of the resin. The reduction in glueline pH may also cause phenolic resin
solids to precipitate.
Nguyen [18] carried out a detailed study of effect of addition of
extractives of kapur on phenolic resin. It was observed that alcohol-, ether-,
and water-soluble extractives when added to a phenolic resin caused pre-
mature gelation of the resin and that the added extractives also caused a
substantial increase in the water solubility of the cured resin. Confirmation
to this effect has also been reported by Wellons et al. [17] who observed
that the premature gelation of resins altered the flow property of phenolic
resins and hindered their penetration into wood. The extractives were also
reported to cause an incomplete curing of the resins resulting in a weak
glue bond.
The extractives can alter the pH of an adhesive at the bond line, thereby
influencing the cure of the adhesive. For example, extractive of kapur is
sufficiently acidic to decrease the pH of the bond line from 11 to 9.5 imme-
diately after spreading of the adhesive [18]. This reduction in pH causes
the precipitation of phenolic resin solid and an increase in hot press time
significantly. Consequently, a cured resin of substantially lower cross-link
density and poor adhesive strength is produced. Difficult-to-bond kapur
and other veneers can be easily bonded with a PF resin to which polymeric
isocyanate (pMDI) is added between 10% and 30% of the phenolic resin
solids [19, 20].
The concentration of soluble extractives in the adhesive may be as
high as 10–20%, and this is enough to cause premature gelation of
216 Adhesives for Wood and Lignocellulosic Materials

phenol–formaldehyde adhesive. Wospakrik [21] found a high correla-


tion (correlation coefficient 0.87) between extractive content and the
shear strength of epoxy-resin-bonded joints and fairly strong correlations
(0.66 and 0.69) between extractive content and the wood failure obtained
with isocyanate– and phenol–resorcinol bonded joints. Although the
extractives were not identified, they were extracted by an alcohol-ben-
zene, alcohol, hot-water extraction sequence, which removes most types
of wood extractives.
The effect of extractives on the gel time of urea–formaldehyde resin
depends on whether the extractive is water soluble or ethanol soluble.
Albritton and Short, while investigating on the effect of the water-soluble
and alcohol-soluble extractives of southern hardwoods, found that the gel
time of urea–formaldehyde resin increased by water-soluble extractives
but shortened by the alcohol-soluble extractives [22].
An interesting finding of Roffael and Rauch [23] was that the old-
growth white oak (pH 3.5) wood had difficulty in gluing due to its high
acidity compared to young-growth oak wood, which has a pH of 4.5. The
difference in the pH values of old-growth and young-growth oaks woods
is attributed to the presence of extractives. Abe and Akimoto [24] reported
that a phenolic resin was changed from basic to neutral in the presence of
acidic extractives, resulting in the formation of a large quantity of dimeth-
ylene ether linkages and thus requiring a prolonged curing time. Abe and
Ono [25] also demonstrated that the acidity of extractives reduced glueline
pH, and this reduction in pH required a significant increase in press time
to properly cure the phenolic resins.
Extractives have been found to affect the energy of activation of the
curing reaction of urea–formaldehyde and phenol–formaldehyde resins as
confirmed by the measurements on 20 wood species [26, 27].
Extractives often change the surface characteristics of wood, thereby
affecting adhesion properties in the following ways: (1) It affects the wet-
tability and polarity of the wood surface; (2) heavy deposits of extractives
on the gluing surface form a physical barrier (i.e., a weak boundary layer),
which also prevents the anchoring of adhesives; and (3) chemical incom-
patibility between extractives and adhesives affects the normal flow, cur-
ing, and setting characteristics of adhesives. In addition, these gluing
interferences caused by extractives may act individually or in combina-
tion [1, 18, 28].
Migration of extractives to the surface causes the change in the wettabil-
ity of wood with time. Gray [29] carried out an extensive study on the wet-
tability of wood by measuring the advancing and receding contact angles
of 19 wood species with time. It was also found that sanding the surfaces of
Wood Surface Inactivation Due to Extractives 217

“contaminated” specimens resulted in the reduction of contact angle, the


extent of reduction depending on the species. Gray [29] attributed surface
contamination to low-molecular-weight fatty acids, high extractives con-
tent, and high resin content.
The contact angles with time for extracted and unextracted Douglas-
fir and redwood were studied by Nguyen and Johns [30]. Douglas-fir was
more wettable after extraction with benzene-alcohol, both initially and
after 84 h, whereas redwood showed a slight decrease of wettability after
extraction.

9.4 Effect of pH of Wood on the Adhesion


pH and the buffering capacity of wood are important properties that influ-
ence the adhesion. Most thermosetting wood adhesives are catalyzed by
acidic or alkaline additives to ensure rapid curing under hot pressing con-
ditions. For example, urea–formaldehyde cures under acid conditions and
phenolic resins cure under alkaline conditions. The initial wood pH could
be such that it can partially neutralize acidic and alkaline environments
at the bond line. If the wood is acidic, then appropriate adjustment in the
percentage of hardener like ammonium chloride is required in order to
compensate for the acidic environment already existing in wood. Only
then can an optimum cure be achieved [31, 32].
It is well known that, except for very few species, most of the wood spe-
cies exhibit an acidic behavior with a pH of between 4.0 and 5.5; strongly
acidic species like Douglas-fir has a pH of 3 and alkaline species is African
Blackwood (Dalbergia melanoxylon) [33]. The sources of acidity in wood
are acetyl groups and uronic acid residues present in hemicelluloses. Some
free organic acids and esters also occur in many species. Additionally, the
contribution of polyphenolic substances such as tannins is also of great
importance in some hardwoods, like the heartwoods of oak and chestnut,
for their acidic property [34].
Mizumachi and Morita [27] showed that the curing reaction of phenolic
resin could be delayed by some wood species, resulting in higher activation
energy during the curing process of PF resin with wood.
The pH, acidity, and alkalinity of wood have been estimated in differ-
ent ways. In most studies, wood is extracted with hot or ambient tem-
perature water and the pH is measured on the extract. Alkalinity and
acidity (buffering capacities) are estimated by titrating the extract with
acid or alkali to a specific pH or until there is a change in pH by one unit
[32, 35–39].
218 Adhesives for Wood and Lignocellulosic Materials

9.5 Effect of Extractive Migrations during Kiln


Seasoning on Adhesion
Wood is dried in heated, humidity-controlled kilns or in various types of
dryers. In kiln drying, wood stays relatively cool (at a constant rate period)
during the active loss of moisture, and the major effects of drying are the
movement and concentration of extractives at the surface. After moisture
loss, continued exposure of the wood to elevated temperatures induces
thermal changes in the extractives and in the wood substance. In the initial
stages, prior to drying, the concentration of extractives is relatively uniform
in wood. During the drying operation, extractives start to diffuse towards
the surface. As the drying proceeds, the concentration of extractives in the
core diminishes and the surface concentration of extractives uniformly
increases, resulting in a concentration gradient. The extractives accumu-
lated on the wood surface can thus influence the curing of adhesives.
Plomley, Hillis, and Hirst [14] found that hydrolyzable tannins decreased
the wet and, in some cases, the dry bond quality. They suggest that these
tannins must reach a threshold level of concentration at the wood surface
(0.4–2.0 g dry extract/m2) before they affect the cure of phenol or resor-
cinol adhesives. More specifically, they noted that the ellagic, but not the
gallic, acid moiety in the hydrolyzable tannin had a strong adverse effect on
wet bond quality. The absorbed tannin may form a weak boundary layer by
simply diluting the adhesive or by interfering with the polymerization and
cross-linking of the resin.
The drying process can also alter the chemical nature of the extractives.
Extractives exposed to high temperature during drying in kilns or ovens
may be converted from hydrophilic to hydrophobic substances.

9.6 Methods to Reduce the Influence of Extractives


on Wood Adhesion
9.6.1 Mechanical Method
The adverse effect of the extractives on the adhesion in wood can be mini-
mized by mechanical or chemical means. The surface contamination of all
kinds can rapidly and economically removed by sanding [1] or even more
effectively by planning.
Several investigators compared the performance of plywood made from
abrasive- and knife-planed veneers. Jokerst and Stewart [40] reported
that the strengths of the plywood made from the respective veneers were
Wood Surface Inactivation Due to Extractives 219

similar, but the durability of plywood made from abrasive-planed veneer


had inferior durability.
Dougal et al. [41] showed that a light knife-planing of bonding surfaces
removed surface contaminants and simultaneously exposed the highly
polar secondary cell walls to which adhesives bonded most effectively.
Abrasive-planing of wood produced rougher surfaces and caused more
cellular damages than the knife-planed surfaces [42]. Although mechani-
cal damages to the cell walls on the wood surfaces may increase percentage
of wood failure, the rough surfaces obtained during abrasive sanding often
produce discontinuous gluelines and thus reduce the durability.

9.6.2 Chemical Method


A surface treatment with chemicals such as sodium hydroxide solution or
prior removal of extractives of wood by suitable solvents may reactivate
the surface and improve glue bond quality. Chen [9] carried out surface
treatments of veneers of eight tropical species with NaOH solution and
found that an increased strength of the adhesive joints could be obtained
by using urea–formaldehyde and resorcinol–formaldehyde as adhesives.
He attributed the improvement in bond quality to the increased pH and
wettability of treated surfaces as a result of extractive removal. Hancock
[6] showed that the glue bond quality of Douglas-fir plywood could be
improved by extracting the veneers with methanol-benzene before drying
and gluing.
Chen [43] improved the glue bond quality of fire-retardant treated
plywood by treating the surfaces with an alcohol solution of NaOH and
by pressing at a higher press temperature and longer press time. Wellons
et al. [17] and Dougal et al. [41] also found that removal of water, alcohol,
and 1% NaOH solubles from kapur veneer greatly improved the bond
quality with phenol–formaldehyde adhesives.
Roffael and Rauch [23] attributed the difficulty in bonding white oak
particles with phenol–formaldehyde resins to low pH caused by extractives.
They found that extraction of white oak wood particles with hot water had
only a marginal improvement on bonding quality and that an extraction
with 1N sodium carbonate solution greatly improved the bond quality. In
addition, they found that adding extra amounts of NaOH to the phenolic
resin was also effective in improving the bond quality, especially the dura-
bility, of white oak particleboard. Kuo et al. [4] also reported that glue bond
quality of white oak plywood could be significantly improved by soaking
the veneers in 1% NaOH aqueous solution for 5 min. They attributed this
improvement to the removal of extractives on veneer surfaces and to the
220 Adhesives for Wood and Lignocellulosic Materials

increase in pH as a result of the treatment. According to Anderson and


Fearing [3, 44], solvent seasoning is a possible method of drying wood and
simultaneously removing a large quantity of extractives. The cost of sol-
vent seasoning, however, has been estimated to be at least three times more
expensive than conventional drying methods, thus limiting its economical
application to improve gluability of species rich in extractives.

References
1. Hse, C.-Y. and Kuo, M.L., Influence of extractives on wood gluing and
finishing—A review. Forest Prod. J., 38, 1, 53–58, 1988.
2. River, B.H., Vick, C.B., Gillespie, R.H., Chapter 1, in: Treatise on Adhesion
and Adhesives, vol. 7, J.D. Minford (Ed.), Marcel Dekker, New York, 1991.
3. Anderson, A.B. and Fearing, W.B., Jr., Distribution of extractives in solvent
seasoned redwood lumber. Forest Prod. J., 11, 240–242, 1961.
4. Kuo, M.L., De Carlo, D., Hse, C.-Y., Influence of extractives on bonding
properties of white and southern red oak. J. Adhesion, 16, 257–278, 1984.
5. Swanson, J.W. and Cordingly, S., Surface chemical studies on pitch. II. The
mechanism of the loss of absorbency and development of self-sizing in
papers made from wood pulps. Tappi, 42, 812–819, 1959.
6. Hancock, W.V., Effect of heat treatment on the surface of Douglas-fir veneer.
Forest Prod. J., 13, 81–88, 1963.
7. Thomas, R.J., Gluing characteristics on determa. Forest Prod. J., 9, 266–271,
1959.
8. Roffael, E., Significance of wood extractives for wood bonding. Appl.
Microbiol. Biotechnol., 100, 1589–1596, 2016.
9. Chen, C.-M., Effect of extractive removal on adhesion and wettability of
some tropical woods. Forest Prod. J., 20, 36–40, 1970.
10. Hemingway, R.W., Thermal instability of fats relative to surface wettability of
yellow birchwood (Betula luteal). Tappi, 52, 2149–2155, 1969.
11. Hillis, W.E., Forever amber: A story of the secondary wood components.
Wood Sci. Technol., 20, 203–227, 1986.
12. Plomley, K.F., Hillis, W.E., Hirst, K., The influence of wood extractives on the
glue-wood bond. I. The effect of kind and amount of commercial tannins and
crude wood extracts on phenolic bonding. Proc. IUFRO-5 Meeting, Vol. 2,
International Union of Forestry Research Organizations, Cape Town, South
Africa, September 22–October 12, 1973.
13. Narayanamurti, D., Gupta, R.C., Verna, G.M., Influence of extractives on the
setting of adhesives. Holzforsch Holzverwert, 14, 5/6, 85–88, 1962.
14. Plomley, K.F., Hillis, W.E., Hirst, K., The influence of wood extractives on the
glue-wood bond. I. The effect of kind and amount of commercial tannins and
crude wood extracts on phenolic bonding. Holzforschung, 30, 14–19, 1976.
Wood Surface Inactivation Due to Extractives 221

15. Sanderman, W., Dietrichs, H.H., Puth, M., On the inhibition of drying with
finished timbers. Holz. Roh. Werkst., 18, 63, 1960.
16. Gardner, J.A.F., Extractive chemistry of wood and its influence on finishing.
Official Digest, 698–707, June, 1965.
17. Wellons, J.D., Krahmer, R.L., Raymond, R., Sleet, G., Durability of exterior
siding plywood with Southeast Asian hardwood veneers. Forest Prod. J., 27,
2, 38–44, 1977.
18. Nguyen, D., Effect of Wood Extractives on Cure of Phenolic Resin. Dissertation,
Oregon State University, Corvallis, Oregon, 1975.
19. Pizzi, A., Valenzuela, J., Westermeyer, C., Non-emulsifiables, water-based,
diisocyanate adhesives for exterior plywood, Part 2: Industrial application.
Holzforschung, 47, 69–72, 1993.
20. Pizzi, A., unpublished results 1997.
21. Wospakrik, J.M., The effect of wood chemical characteristics and accelerated
test methods on bond durability. Dissertation, Washington State University,
Pullman, WA, 1984.
22. Albritton, R.O. and Short, P.H., Effects of extractives from pressure-refined
hardwood fibre on the gel time of urea–formaldehyde resin. Forest Prod. J.,
29, 2, 40–41, 1979.
23. Roffael, E. and Rauch, W., Extractives of oak and their influence on the glu-
ing with alkaline phenolic–formaldehyde resins. Holz. Roh. Werkst., 32, 182–
187, 1974.
24. Abe, I. and Akimoto, N., The inhibitory effect of kapur wood extractives
on the curing reaction of the resol. J. Japan Wood Res. Soc., 22, 191–196,
1976.
25. Abe, I. and Ono, K., Effect of the acidity of some tropical wood extractives on
the curing of the resol. J. Japan Wood Res. Soc., 26, 686–692, 1980.
26. Mizumachi, H., Activation energy of the curing reaction of urea resin in the
presence of wood. Wood Sci., 6, 14–18, 1973.
27. Mizumachi, H. and Morita, H., Activation energy of the curing reaction of
phenolic resin in the presence of wood. Wood Sci., 7, 256–260, 1975.
28. Wang, Y.-S., Influence of extractives on bonding properties of white oak.
Dissertation, Iowa State University, Ames, IA, 1992.
29. Gray, V.R., The wettability of wood. Forest Prod. J., 12, 452–461, 1962.
30. Nguyen, T. and Johns, W.E., The effect of aging and extraction on the sur-
face free energy of Douglas fir and redwood. Wood Sci. Technol., 13, 29–40,
1979.
31. Johns, W.E. and Niazi, K.A., Effect of pH and buffering capacity of wood
on the gelation time of urea–formaldehyde resins. Wood and Fiber, 12, 4,
255–263, 1980.
32. Xing, C., Zhang, S.Y., Deng, J., Effect of wood acidity and catalyst on UF resin
gel time. Holzforschung, 58, 408–412, 2004.
33. Irle, M., pH and why you need to know it. Wood Based Panels Int, 2012.
http://www.wbpionline.com/features/ph-and-why-you-need-to-know-it/.
222 Adhesives for Wood and Lignocellulosic Materials

34. Uçar, G. and Balaban Uçar, M., The estimation of acidic behavior of wood by
treatment with aqueous Na2HPO4 solution. J. Anal. Methods Chem., 9, 2012.
Article ID 496305.
35. Johns, W.E. and Niazi, K.A., Effect of pH and buffering capacity of wood on
the gelation time of urea formaldehyde resin. Wood Fiber Sci., 12, 255–263,
1980.
36. Hachmi, M.H. and Moslemi, A.A., Effect of wood pH and buffering capacity
on wood–cement compatibility. Holzforschung, 44, 425–430, 1990.
37. He, G. and Riedl, B., Curing kinetics of phenol formaldehyde resin and
wood–resin interactions in the presence of wood substrates. Wood Sci.
Technol., 38, 69–81, 2004.
38. Passialis, C., Voulgaridis, E., Adamopolous, S., Matsouka, M., Extractives,
acidity, buffering capacity, ash and inorganic elements of black locust wood
and bark of different clones and origin. Holz. Roh. Werkst., 66, 395–400,
2008.
39. Pedieu, R., Riedl, B., Pichette, A., Measurement of wood and bark particles
acidity and their impact on the curing of urea formaldehyde resin during the
hot pressing of mixed panels. Holz. Roh. Werkst., 66, 113–117, 2008.
40. Jokerst, R.W. and Stewart, H.A., Knife- versus abrasive-planed wood: Quality
of adhesive bonds. Wood Fiber Sci., 8, 107–113, 1976.
41. Dougal, E.F., Krahmer, R.L., Wellons, J.D., Kanarek, P., Glueline characteris-
tics and bond durability of Southeast Asian species after solvent extraction
and planing of veneers. Forest Prod. J., 30, 7, 48–53, 1980.
42. Caster, D., Kutscha, N., Leick, G., Gluability of sanded lumber. Forest Prod. J.,
35, 45–52, 1985.
43. Chen, C.-M., Gluing study of pre-treated fire-retardant plywoods—Part I.
Forest Prod. J., 25, 33–37, 1975.
44. Anderson, A.B. and Fearing, W.B., Jr., Solvent seasoning of tanoak. Forest
Prod. J., 10, 234–238, 1960.
10
Surface Modification of Wood

10.1 Introduction
Due to increasing concerns on global warming, the forestry policies have
been laid down globally to conserve forests and to halt the loss and deg-
radation of forest ecosystems. In order to achieve this goal to promote
and restore forests, environmentally sound forest harvesting practices are
adopted. This is reflected in the restricted availability high quality of wood
raw materials to the wood industry. As a result, wood of large diameters
are no longer easily available for the production of veneer-based compos-
ites. The industry is therefore compelled to use secondary species of poor
quality with gums or resins on the surface, which have adverse effect on
bonding. In addition, drying veneers at high temperature to increase pro-
duction and save energy results in the inactivation of veneer surfaces due
to reasons discussed in Chapter 8. In order to obtain a better performance
with the available raw materials, the surface modification of veneers is
necessary [1].
Surface modification of wood is defined as the application of a chemi-
cal, physical, or biological agents to the wood surface in order to obtain a
desired improvement of performance [2, 3]. The improvement of perfor-
mance may be effected by modification of the surface energy in order to
improve adhesion. Wood surfaces can be treated by plasmas from various
sources, by chemical or enzymatic grafting of functional groups or by coat-
ing by either sol–gel methods, or deposition of nanoparticles. The target
properties to be improved are surface activation for better wettability, glu-
ing and adhesion of surface coatings, and resistance to weathering.
Surface modifications are important not only for solid wood and veneers
but also for wooden particles, wood-based fibers, and other non-wood lig-
nocellulosic fibers. They are mainly intended to improve adhesion between
the lignocellulosic particles/fibers and thermoplastic matrices in wood–
polymer composites.

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (223–238) ©  2019
Scrivener Publishing LLC

223
224 Adhesives for Wood and Lignocellulosic Materials

10.2 Surface Modification Methods


10.2.1 Plasma and Corona Treatments
One of the technologies to improve the surface characteristics of wood is
the plasma treatment [4]. The purpose of plasma treatment is to improve
adhesion. The plasma technology has recently attracted research interest
in many countries [5, 6]. It has also been applied to lignocellulosic fiber
panels [7].
Plasma can be described as partially ionized gas consisting of a complex
mixture of ions, electrons, ultraviolet photons, and reactive neutrals such
as radicals, excited and ground state molecules, etc. [8]. In other words,
plasma is a collection of all the abovementioned species. Plasma is usually
generated from gases by the absorption of enough energy to cause the elec-
trons to become separated from their nuclei, resulting in ionization and
fragmentation. When this happens, the plasma no longer acts like a gas.
Plasma, as characterised by particles with very low degrees of ioniza-
tion and little penetrating energy (called cold plasma), is often used for
surface modification or activation [18]. Excited particles in cold plasma
have energy levels (0.5–3 eV) sufficient to break chemical bonds on the
surface of organic and inorganic substrates [9, 10]. These broken chemical
bonds are thermodynamically unstable and combine with ambient mole-
cules to re-engineer the surface of the material [11]. This type of modifica-
tion (cleaning, etching, or cross-linking), however, is limited to the surface,
typically to a depth of a few molecular layers. Plasma is also considered to
be the fourth state of matter because it is more highly activated than solid,
liquid, or gaseous states.
More detailed definitions of plasma have appeared in the literature
[12–14].
The characteristic behavior of plasma depends on the ratio of ionized to
neutral particles and the particle energies. These characteristics lead to a
broad spectrum of plasma types.
Plasmas can be classified either as:

(1) Equilibrium plasmas (also called thermal or hot plasmas,


with high degrees of ionization) or
(2) Non-equilibrium plasmas (also named non-thermal or
cold plasmas, with low degree of ionization).

In equilibrium plasmas, the temperature of electrons and of the gas is


the same, and can be as high as 10,000°C. In non-equilibrium plasmas, the
Surface Modification of Wood 225

gas is at ambient temperature, but the temperature of electrons is very high,


of about 10,000°C [6, 7, 12, 15].
Thus, a plasma is referred to as being “hot” if it is nearly fully ionized,
or “cold” if only a small fraction (for instance 1%) of the gas molecules are
ionized, but other definitions of the terms “hot plasma” and “cold plasma”
are common. As mentioned above, even in cold plasma, the electron tem-
perature is still typically several thousand degrees centigrade [16].
Plasmas transfer energy to the surfaces, thereby effecting surface modi-
fication. In order to sustain the plasma state, energy must be continuously
injected into the system. The easiest way to ensure this requirement is by
using electrical energy [12, 17]. An electrical field is necessary to generate
a plasma. In a vacuum chamber, the ions and electrons have long lifetimes.
Radio-frequency power can be applied to two metal plates immersed in the
vessel, creating a capacitive discharge.
Gas discharge plasmas, also known as low-temperature plasmas, have
drawn much attention in the past few decades because of their importance
in many technological developments. A dielectric barrier discharge (DBD)
plasma generator functions due to electrical discharge between two elec-
trodes separated by an insulating dielectric barrier. It was originally called
silent discharge and is also known as ozone production discharge.
Because of the unique characteristics of plasmas, a growing number of
applications have emerged [18]. The species in plasma that modify materi-
als are electrons, ions, and radicals. The latter two species are generated by
collisions between electrons and gas molecules existing in plasma.
Two categories of plasma, the thermal and non-thermal ones, can be
defined according to the conditions in which they are created.
Thermal plasmas are obtained at high pressure (≥105 Pa) and need sub-
stantial power (up to 50 MW). Non-thermal plasmas are obtained at lower
pressures and use less power. Such plasma can be generated by electric
discharges in gases at low pressure.
Over recent years, atmospheric plasmas have been developed that are
now available for industry use. This technique is already used in the plas-
tic and glass industries to improve adhesion between coating and sub-
strate by modifying substrate surface energy or/and grafting new chemical
functions.
Plasma treatment technology is simple and has low operating costs. In
addition, there is no environmental pollution created by this treatment
process, and it has a very promising future for engineering applications.
Plasma treatment only affects the near surface of a material without chang-
ing bulk material properties, as opposed to chemical modifications [19,
20]. The plasma treatment system is described by Aydin [4] and Petric [2].
226 Adhesives for Wood and Lignocellulosic Materials

The application of plasma in wood and wood-based materials has two


main goals: (1) the improvement of adhesion [21–25] and (2) the depo-
sition of low-surface-energy thin films to increase the barrier properties
against polar and nonpolar liquids and gases. It is a dry and a clean process
with little environmental impact.
Plasma modification causes increased surface polarity induced by oxi-
dation reaction, leading to the formation of hydroxyl, carboxyl, aldehyde,
and other polar functional groups [26].
Improvement of specific surface properties in the case of plasma treat-
ment depends on the plasma reactor’s design, the plasma gases used, and
plasma treatment parameters. Examples of such properties are wood wet-
tability, water repellence, surface free energy, the adhesion strength, and
biocompatibility [14] and coating adhesion. Plasma can also alter the sur-
face properties even of normally inert materials such as ceramics [13]. The
plasma gas is probably one of the most important parameters. Inorganic
plasma gases such as nitrogen, oxygen, carbon dioxide, air, water vapor,
argon, and others are generally used to alter surface properties such as
wettability.
Over recent years, atmospheric plasmas have been developed that are
now available for industry use. This technique is already used in the plas-
tic and glass industries to improve adhesion between coating and sub-
strate by modifying substrate surface energy or/and grafting new chemical
functions.
It is well known that a wood surface exposed to contaminants such as
dust or atmospheric grime can experience surface inactivation that can
reduce the attractive forces on the wood surface and lead to a decrease in
wettability. But it was shown that low-pressure O2 plasma treatment can
reactivate the surfaces of spruce wood for adhesive bonding and to increase
wettability even after a 9-year period of natural surface inactivation. Thus,
the target properties to be improved or even introduced are mostly surface
activation for better gluing and adhesion of surface coatings, wettability,
and resistance to weathering.
Different gas species can be used in plasmas in order to obtain the
desired surface properties, by influencing the interactions between the
gas and the solid surface depending on the nature of the gas species used
in plasma. Some gas species can change the surfaces into hydrophobic,
whereas other ones create hydrophilic surfaces [27]. One of the most pop-
ular gas species used in plasma influencing wood surfaces is oxygen, with
a target to improve adhesion [4, 28]. Oxygen plasma has also been used
for cleaning different areas such as micro-electro-mechanical systems.
Literature contains some studies about the effects of plasma treatment with
Surface Modification of Wood 227

different gas species on adhesion properties [29], contact angle [4], wetta-
bility, and bonding.

10.2.2 Corona Treatment


The non-equilibrium plasmas are classified roughly into two categories:
(a) ordinary low-temperature plasmas generated at low pressure, also called
“glow discharge”, and (b) corona discharges at atmospheric pressure [12].
Glow discharges are the most widely used technique for low-temperature
plasmas. The reason is that the mean free path of activated gas molecules
is longer in vacuum, which allows the use of a bigger distance between the
electrodes and the samples. This kind of plasma has been used as a highly
effective pretreatment for the surface preparation of low-surface-energy
polymers for adhesive bonding [6].
The corona discharge treatment (CDT) has a more restricted appli-
cation range, since it is limited to the materials that are responsive to
this method of surface treatment and the material configuration (geom-
etry). Complex shapes cannot easily be treated because the treatment
quality is a function of the distance of the surface to the electrode. Also,
since corona treatments are normally conducted in ambient air, they
can be affected by environmental changes in the location where it takes
place. The main advantages of CDTs are that, since no vacuum system is
needed, the equipment investment is much lower than that for ordinary
low-temperature plasma installations. In CDT, the discharge is generated
by applying a high-frequency, high-voltage signal to an electrode that is
separated from an earthed plane, usually by only a few millimeters, by an
air gap, the substrate, and a layer of dielectric material [6].
Uehara and Jodai [30] found that the joint strength of corona-treated api-
tong (Dipterocarpus grandiflorus Blanco) bonded with urea–formaldehyde
adhesive increased with the corona treatment. However, the results for the
phenol–resorcinol–formaldehyde (PRF) resin were different. After the treat-
ment, the wettability by the PRF resin did not improve and no improvement
in the joint strength was observed.
Back studied the effects of the corona discharge on the wettability of
teak, birch, and pine (Pinus sylvestris L.) [31]. The treatment resulted in
a significant increase in the wettability of teak and birch, but the effect
on pine was minimal. It was concluded that the CDT could be used to
improve the wettability of timbers with high resin or oil content and also
to treat wood surfaces that have undergone thermal aging.
Sakata et al. (1993) have shown that the treatment by corona of the
surface of several softwood, hardwood, and tropical wood veneers caused
228 Adhesives for Wood and Lignocellulosic Materials

a considerable increase in the wettability of the surface by the urea resin


and the improvement in the adhesion of veneers with water-based adhe-
sives [32]. An increase in the wettability of corona-treated wood veneers of
hardwoods, softwoods, and tropical woods was observed. The wettability
increased with an increase in the degree of treatment. This is attributed
to the oxidation of the highly hydrophobic surface layer. To elucidate the
nature of any chemical change occurring on the wood surface, the dyeing
examination of the wood and its components with Schiff ’s reagent was car-
ried out, and the results showed a high degree of dyeing ability for corona-
treated samples as compared to untreated ones, thereby indicating that
aldehyde groups increased as a result of the corona treatment.
Thus, the wettability of corona-treated wood veneers resulted mainly
from the oxidation of the high hydrophobic surface layer.

10.2.3 Plasma Applications for Wood Surface Plasma Treatments


Some applications have been investigated. Podgorski et al. studied the
influence of various plasma treatments on several fir species [33]. They
showed that adhesion of coatings could be improved. Rehn et al. showed
that the fracture strength of glued black locust (Robinia pseudoacacia)
could be increased and coating delamination reduced [34]. The fire and
moisture resistance of different Philippine species were improved after
hydrogen plasma treatment by Blantocas and Al-Aboodi [35]. Evans et al.
investigated the impact of plasma treatment on the wettability and glue
bond strength of four eucalyptus species [36]. Finally, Wolkenhauer and
coworkers studied several properties of wood after plasma treatments [29,
37]. In 2009, they demonstrated that a plasma treatment of wood surfaces
was superior to sanding for increasing surface energy [29].
Wolkenhauer et al. [29, 37] carried out the surface treatment of beech,
oak, spruce, and Oregon pine wood by plasma produced by DBD and com-
pared the surface characteristics of the plasma-treated wood with those
of sanded surface. The surface energy of aged, freshly sanded, or plasma-
treated surfaces was examined by contact angle measurement and calcu-
lation of work of adhesion. Plasma treatment turned out to be superior to
sanding.
The increase in surface energy was caused by an increased polar group
on the surface. To see whether a combined treatment amplified this effect,
a combination of sanding and plasma treatment was also investigated.
Avramidis et al. [38] carried out the plasma treatment at atmospheric
pressure using a DBD in order to increase the surface hydrophilicity of
wood and wood-based materials. Surface energy determination by contact
Surface Modification of Wood 229

angle measurement revealed an increase in the polar component of sur-


face energy and in total surface energy following plasma treatment. X-ray
photoelectron spectroscopy revealed an increase in O/C ratio, indicating
the generation of polar groups. An atmospheric-pressure plasma jet using
hexamethyldisiloxane as precursor and air as process gas was used for
thin-layer deposition.
The feasibility of plasma polymerization on wooden substrates at
atmospheric pressure to create water-repellent characteristics was also
investigated. Atomic force microscopy revealed a closed surface layer con-
sisting of silicon, oxygen, carbon, and hydrogen that exhibited low water
permeability.
Aydin et al. [4] enhanced the ability of the inactivated wood surfaces
for better adhesion and bonding by employing plasma treatment. In their
studies, a low-pressure plasma treatment was applied to reactivate the sur-
faces of spruce wood for glue bonding and to increase wettability after a
9-year period of natural surface inactivation. The inactivation was found
to have been caused by contaminants such as dust or atmospheric grime
due to the exposure of the surface to atmosphere over a long period of
time. The inactivation mechanisms were proposed to be due to reduction
of the attractive forces on the wood surface that leads to a decrease in
wettability. Changes in contact angles, surface energy, surface color, and
bonding strength of inactivated and oxygen plasma-treated wood surfaces
were studied. Wettability, bonding, and other mechanical properties of
plywood panels produced from the wood veneers treated with the oxy-
gen plasma were increased .Thus, plasma treatment has been applied to
recover inactivated wood surfaces for better adhesion and bonding. As
mentioned earlier, the plasma treatment technology is very simple and the
cost is rather low. In addition, this treatment produces no environmental
pollution.
In their study on the surface treatment of Pinus yunnanensis wood,
Fang et al. used microwave plasma [39].The microwave plasma presented
very significant treatment effects on the treated surface, even under weak
treatment conditions. The treated surface showed better surface wettabil-
ity. The contact angles on the treated surface measured from deionized
water, glycerin, and diiodomethane decreased sharply, even to 0°. The
treated surface also presented higher surface free energy, for example,
61.4–62.8 mJ m−2, being greatly improved by microwave plasma compared
to that of untreated surface of 46.5 mJ m−2. The bond strength of the treated
surface was 7.34 MPa, about 16% higher than 6.31 MPa for the untreated
surface. The authors recommend that this technique might be widely used
in wood modification and wood processing.
230 Adhesives for Wood and Lignocellulosic Materials

Acda et al. reported on the effects of plasma modification on surface


properties of three species of wood [26, 28]. The authors made use of DBD
to modify the surface properties of Shorea contorta, Gmelina arborea, and
Acacia mangium commonly used in the Philippines and Southeast Asia for
furniture and wood-based panel manufacture. Wood specimens exposed
to oxygen plasma resulted in a several-fold increase in surface free energy.
They also reported that cold plasma treatment induces physical and chemi-
cal changes to the depth of a few micrometers on the wood surface without
any change in bulk properties [14]. Cold oxygen plasma, which is com-
monly used to improve wood surface, can react with wood to produce a
variety of oxygen functional chemical species such as C–O, C=O, O–C=O,
C–O–O, and CO3 at the surface [40].
No apparent surface defects or changes in bulk properties were
detected. The increased surface free energy could be positively correlated
with the development of good joint strength during gluing with polar
adhesives.
Cold plasma has been found to be a powerful tool that can be used to
improve the wetting and adhesion properties of wood [26, 28, 36, 41, 42]
and wood composites [37, 43, 44]. It is believed that plasma modification
results in increased surface polarity brought about by the formation of
hydroxyl, carboxyl, aldehyde, and other polar functional groups [45]. The
increased surface polarity improves wettability and penetration behavior
(hydrophilicity) of wood to liquid [34]. The wettability assists in establish-
ing extensive and molecular-scale contact with the wood surface, which is
critical to the development of strong adhesion at the adhesive–wood inter-
face and also for the coating adhesion [33, 34, 36, 41, 42, 46–48].
Wascher et al. reported that the five-layer plywood boards made from
plasma pretreated veneers exhibited up to 2.7-fold improvement in shear
strength compared to plywood made of untreated veneers [49].
Wolkenhauer et al. investigated the adhesion of PVA glue on fiberboard
and particle board in the untreated state and plasma-treated with a DBD
at atmospheric pressure [37]. A force-sensitive peel test was carried out
and confirmed increased adhesion after plasma treatment.

10.3 Enzymatic Modification for Hydrophobicity


Preparation of wood thermoplastic polymer matrix composites requires
hydrophobization of the lignocellulosic fibers by suitable surface
modification so as to impart compatibility with thermoplastic poly-
mers. Chemical modifications and plasma treatment of the fibers are
Surface Modification of Wood 231

currently adopted. Enzymatic modification can be an attractive alterna-


tive for increasing lignocellulose hydrophobicity. An example is laccase-
mediated grafting of laurylgallate onto pulp, which has increased its
hydrophobicity [50].
Laccase-catalyzed coupling of fluorophenols was shown to increase
hydrophobicity of wood veneers [51]. The authors also provided the
first in vitro evidence for the possibility of establishing covalent bonds
between complex lignin model compounds and the fluorophenols, which
were found to bond to sinapyl units via 4–0–5 linkages while coupling
to guaicyl units occurred through 5–5 linkages. The same authors also
elegantly demonstrated the ability of laccase to couple long-chain alkyl-
amines onto wood veneers, resulting in an increase in hydrophobicity
[52]. Research has been focused on the application of laccases toward
functionalization of pulps in order to produce novel paper products. The
pioneering works of Yamaguchi and coauthors [53–55] demonstrated
the ability of laccase to polymerize various phenolic compounds to form
dehydrogenative polymers, which were subsequently coupled to thermo-
mechanical pulps in the presence of peroxidase. Incorporation of phe-
nolics (vanillic acid, catechol, mimosa tannin, and tannic acid) in the
presence of laccase enabled production of paper with increased tensile
strength and water-resistant properties attributed to the coupling of free
phenolic groups on the fiber surface with the added dehydrogenative
polymers. The progress made in the field of laccase-assisted biograft-
ing has been outlined above in order to show the potential for further
exploitation of laccases to the future development of environmentally
friendly adhesively bonded materials.

10.4 Modification of Wood Surface by Chemical


Treatment—Functionalization of Wood
The term grafting is considered to represent covalent bonding of various
substances, very often polymers or monomers.
In the case of wood, these are predominantly cellulose and lignin.
Roman-Aguirre et al. and Wang and Piao define grafting as a molecular
technique that chemically bonds foreign molecules to a surface [56, 57].
An example is to use a silane coupling agent. The coupling agents contain
alkoxy groups that react with wood hydroxyl groups by initially converting
the polar OH groups of wood via Si–O–C (wood) linkages into the hydro-
phobic moiety. An analogous approach with alkoxy-terminated silicones
has been described for the hydrophobization of proteins [58–61].
232 Adhesives for Wood and Lignocellulosic Materials

The disadvantage of the use of such silane coupling agents is that


Si–O–C linkages are formed between wood and the hydrophobic moi-
ety. However, Si–O–C bonds are normally subject to hydrolysis, giving
silsesquioxanes (R-SiO3/2), limiting the degree of hydrophobization [62].
Sebe and Brook have adopted a method in which grafting relies on the
formation of covalent bond, viz., Si–C–O–C (wood), linkages that are
not so susceptible to hydrolysis [63].
In this method, the wood is initially treated with maleic anhydride or
allyl glycidyl ether. This reaction resulted in oligoesterified wood bear-
ing terminal alkenes (see the reaction scheme, Figure 10.1). The terminal
alkenes are reacted with hydride terminated silicones. Hydrosilylation
reaction leads to a wood–silicone moiety containing Si–C–O–C (wood)

O
OH CH2 O CH2 CH=CH2
Wood + H2C CH
Allylglycidylether

OH

O CH2 CH CH2 O CH2 CH=CH2


Wood

H-Si Hydrosilylation

OH

O CH2 CH CH2 O CH2 CH=CH2


Wood

H-Si Hydrosilylation

OH

O CH2 CH CH2 O CH2 CH2 CH2 Si


Wood

Reaction scheme for hydrophobization of wood

Figure 10.1 Hydrophobization of wood.


Surface Modification of Wood 233

linkages, which are hydrolytically more stable than the Si–O–C linkages
formed by using the conventional silane coupling agents. Hydrosilylation
results in the formation of hydrophobic surfaces that remained unim-
paired even after extensive soxhlet extraction with good solvents for
silicones.
In order to convert the hydrophilic surface of bamboo flour into a
hydrophobic surface, Yu et al. employed atom transfer radical polymer-
ization method by which methyl methacrylate was grafted onto the bam-
boo flour surface [64].

10.5 Sol–Gel Method


Another method for the surface modification of wood is to conduct the
chemical reaction of the reagent Si(OEt)4 or Cl-Si(OEt)3, which is more
reactive than Si(OEt)4 with lignocellulosic substrate by base/acid-catalyzed
hydrolysis to form a wood-bound silica composite with glass-like surface
properties. This method is a better alternative to reacting the lignocelulosic
substrates with silane coupling agents. As mentioned in Section 10.4, silane
coupling agents form C–O–Si linkages.
Although application of silane coupling reagents for improvement of
wood flour and wood-derived fiber polymer composites dates from 1983,
the hydrolytic stability of the C–O–Si has not been considered seriously.
The silane coupling mechanism per se depends on the hydrolytic suscepti-
bility of the Si–O–C bonds in Si–OC2H5, which readily hydrolyzes to form
the -Si–OH groups in the presence of acids or alkalis. The result is the for-
mation of a C–O–Si covalent bond formed between cellulose and silicon
that is not hydrolytically stable.
A new sol–gel strategy can be applied to enhance adhesion between
wood/lignocellulosic fibers and polymeric adhesives or matrix mate-
rials. By this new process, glass-like surface properties can be first
imparted onto the wood or lignocellulosic fibers followed by the addi-
tion of silane coupling reagents, which will result in the formation of a
Si–O–Si–C bond between wood and the organic polymeric adhesives or
matrix materials.
The traditional sol–gel process involves the chemistry of metal alkox-
ide precursors, such as silicon alkoxides, and similar metal alkoxides, such
as titanium, aluminum, and zirconium, resulting in amorphous inorganic
materials [2, 65].
The sol–gel reactions are given in the Scheme in Figure 10.2.
234 Adhesives for Wood and Lignocellulosic Materials

1. Formation of reactive species

OC2H5 OC2H5

C2H5-O Si OC2H5 + H2O C2H5-O Si OH

OC2H5 OC2H5

2. Sol formation-Polycondensation
OC2H5 OC2H5 OC2H5
OC2H5
C2H5-O Si O* Si O Si OH
nC2H5-O Si OH
H2O
OC2H5 OC2H5 OC2H5
OC2H5
n-1
Sol

3. Gel formation

OC2H5 OC2H5 OC2H5

C2H5-O Si O* Si O Si OH

OC2H5 OC2H5 OC2H5


n-1
Sol

O O O

O Si O* Si O Si O

O O O
n-1

Gel

Figure 10.2 Sol–gel reaction scheme.

References
1. Tang, L., Zhang, R., Zhou, X., Pan, M., Chen, M., Yang, X., Zhou, P., Chen,
Z., Dynamic adhesive wettability of poplar veneer with cold oxygen plasma
treatment. Bioresources, 7, 3, 3327–3339, 2012.
2. Petric, M., Surface modification of wood: A critical review. Rev. Adhes.
Adhes., 1, 216–247, 2013.
3. Hill, C.A.S., Wood Modification: Chemical, Thermal and Other Processes,
John Wiley & Sons, Chichester, England, 2006.
Surface Modification of Wood 235

4. Aydin, I. and Demirkir, C., Activation of spruce wood surfaces by plasma


treatment after long terms of natural surface inactivation. Plasma Chem.
Plasma Proc., 30, 697–706, 2010.
5. Kogelschatz, U., Dielectric-barrier discharges: Their history, discharge phys-
ics, and industrial applications. Plasma Chem. Plasma Proc., 23, 1–46, 2003.
6. Custodio, J., Broughton, J., Cruz, H., Hutchinson, A., A review of adhesion
promotion techniques for solid timber substrates. J. Adhes., 84, 502–529, 2008.
7. Pizzi, A., Kueny, R., Lecoanet, F., Massetau, B., Carpentier, D., Krebs, A.,
Loiseau, F., Molina, S., Ragoubi, M., High resin content natural matrix-natural
fibre biocomposites. Ind. Crops Prod., 30, 235–240, 2009.
8. Boenig, H.V., Plasma Science and Technology, Cornell University Press, New
York, 1982.
9. Lide, D.R., Handbook of Chemistry and Physics, 84th edition, CRC Press,
Boca Raton, FL, 1993.
10. Denes, A.R., Tshabalala, M.A., Rowell, R., Denes, F., Young, R.A.,
Hexamethyldisiloxane-plasma coating of wood surfaces for creating water
repellent characteristics. Holzforschung, 53, 318–326, 1999.
11. Setoyama, K., Surface modification of wood by plasma treatment and plasma
polymerization. J. Photopolym. Sci. Technol., 9, 243–250, 1996.
12. Kinloch, A.J., Adhesion and Adhesives: Science and Technology, Chapman and
Hall, London, UK, 1987.
13. Moreau, M., Orange, N., Feulloley, M.G.J., Non-thermal plasma technolo-
gies: New tools for bio-decontamination. Biotechnol. Adv., 26, 610–617, 2008.
14. Chu, P.K., Chen, J., Wang, L.P., Huang, N., Plasma surface modification of
bio-materials. Mater. Sci. Eng., R, 36, 143–206, 2002.
15. Pizzi, A. and Mittal, K.L., Handbook of Adhesive Technology, 2nd edn.,
Marcel Dekker, New York, 2003.
16. Hamrang, A., Advanced Non-Classical Materials with Complex Behavior:
Modeling and Applications, vol. 1, CRC Press, Boca Raton, FL, 2014.
17. Pocius, A.V., Adhesion and Adhesives Technology: An Introduction, 2nd edn.,
Hanser Gardner Publications, Cincinnati, OH, 2002.
18. Eliezer, S. and Eliezer, Y., The Fourth State of Matter: An Introduction to
Plasma Science, 2nd edn., Institute of Physics Publishing, Philadelphia, 2001.
19. Demirkir, C., Aydin, I., Colak, S., Colakoglu, C., Effects of plasma treatment
and sanding process on surface roughness of wood veneers. Turk. J. Agric.
For., 38, 663–667, 2014.
20. Demirkir, C., Aydın, I., Colak, S., Ozturk, H., Effect of plasma surface treat-
ment on bending strength and modulus of elasticity of beech and poplar
plywood. Maderas Cienc. Tecnol., 19, 195–202, 2017.
21. de Cademartori, P.H.G., de Muniz, G.I.B., Magalhaes, W.L.E., Changes of
wettability of medium density fiberboard (MDF) treated with He-DBD
plasma. Holzforschung, 69, 187–192, 2015.
22. de Cademartori, P.H.G., Stafford L, L., Blanchet, P., de Muniz, G.I.B.,
Magalhaes, W.L.E., Enhancing the water repellency of wood surfaces by
236 Adhesives for Wood and Lignocellulosic Materials

atmospheric pressure cold plasma deposition of fluorocarbon film. RSC Adv.,


29159–29169, 2017.
23. Hünnekens, B., Peters, F., Avramidis, G., Krause, A., Militz, H., Viöl, W.,
Plasma treatment of wood–polymer composites: A comparison of three
different discharge types and their effect on surface properties. J. Appl.
Polym. Sci., 133, 43376, 2016.
24. Novak, I., Popelka, A., Spitalsky, Z., Mičušík, M., Omastová, M., Valentin,
M., Sedliačik, J., Janigová, I., Kleinová, A., Šlouf, M., Investigation of beech
wood modified by radio-frequency discharge plasma. Vacuum, 119, 88–94,
2015.
25. Kral, P., Rahel, J., Stupavska, M., Šrajer, J., Klímek, P., Mishra, P.K.,
Wimmer, R., XPS depth profile of plasma-activated surface of beech
wood (Fagus sylvatica) and its impact on polyvinyl acetate tensile shear
bond strength. Wood Sci. Technol., 49, 319–330, 2015.
26. Acda, M.N., Devera, E.E., Cabangon, R.J., Pabelina, K.G., Ramos, H.J., Effects
of dielectric barrier discharge plasma modification on surface properties of
tropical hardwoods at low pressure. J. Tropic. For. Sci., 24, 416–425, 2012.
27. Mortazavi, M. and Nosonovsky, M., A model for diffusion-driven hydropho-
bic recovery in plasma treated polymers. Appl. Surf. Sci., 258, 6876–6883,
2012.
28. Acda, M.N., Devera, E.E., Cabangon, R.J., Ramos, H.J., Effects of plasma
modification on adhesion properties of wood. Int. J. Adhes. Adhes., 32, 70–75,
2012.
29. Wolkenhauer, A., Avramidis, G., Hauswald, E., Militz, H., Viol, W., Sanding
vs. plasma treatment of aged wood: A comparison with respect to surface
energy. Int. J. Adhes. Adhes., 29, 18–22, 2009.
30. Uehara, T. and Jodai, S., Gluing of wood by corona-treatment. Mokuzai
Gakkaishi, 33, 777–784, 1987.
31. Back, E.L., Oxidative activation of wood surfaces for glue bonding. For. Prod.
J., 412, 30–36, 1991.
32. Sakata, I., Morita, M., Tsuruta, N., Morita, K., Activation of wood surfaces
by corona treatment to improve adhesive bonding. J. Appl. Polym. Sci., 49,
1251–1258, 1993.
33. Podgorski, L., Bousta, C., Schambourg, F., Maguin, J., Chevet, B., Surface
modification of wood by plasma polymerisation. Pigm. Resin Technol., 31, 1,
33–40, 2002.
34. Rehn, P., Wolkenhauer, A., Bente, M., Förster, S., Viöl, W., Wood surface
modification in dielectric barrier discharges at atmospheric pressure. Surf.
Coat. Technol., 174–175, 515–518, 2003.
35. Blantocas, G.Q. and Al-Aboodi, A.S., Physicochemically modifying wood by
low energy hydrogen ion shower: An alternative plasma-based antitermite
method. Wood Fiber Sci., 43, 4, 449–456, 2011.
36. Evans, P., Ramos, M., Senden, T., Modification of wood using a glow-
discharge plasma derived from water, in: Proceedings of the Third European
Surface Modification of Wood 237

Conference on Wood Modification, C.A.S. Hill, D. Jones, H. Militz, G.A.


Ormondroyd (Eds.), Bangor, UK, 2007.
37. Wolkenhauer, A., Militz, H., Viöl, W., Increased PVA glue adhesion on
particle board and fibre board by plasma treatment. Holz Roh Werkst., 66,
143–145, 2008.
38. Avramidis, G., Hauswald, E., Lyapin, A., Militz, H., Viöl, W., Wolkenhauer,
A., Plasma treatment of wood and wood-based materials to generate hydro-
philic or hydrophobic surface characteristics. Wood Mater. Sci. Eng., 4, 52–60,
2009.
39. Fang, Q., Wang, H., Du, G., Surface wettability, surface free energy, and
surface adhesion of microwave plasma-treated Pinus yunnanensis wood.
Wood Sci. Technol., 50, 285–296, 2016.
40. Belgacem, M.N., Czeremuzhkin, G., Supieha, S., Gandini, A., Surface char-
acterization of cellulose fibres by XPS and inverse gas chromatography.
Cellulose, 2, 145–157, 1995.
41. Blanchard, V., Blanchet, P., Riedl, B., Surface energy modification by RF
inductive and capacitive plasmas at low pressures on sugar maple: An explor-
atory study. Wood Fibre Sci., 41, 245–254, 2009.
42. Blanchard, V., Riedl, B., Blanchet, P., Evans, P., Modification of Sugar
Maple Wood Board Surface by Plasma Treatments at Low Pressure. Contact
Angle, Wettability and Adhesion, vol. 6, Koninklijke Brill NV, Leiden, The
Netherlands, 2009.
43. de Meijer, M., Haemers, S., Cobben, W., Militz, H., Surface energy deter-
minations of wood: Comparison methods and wood species. Langmuir, 16,
9352–9359, 2000.
44. Gindl, M., Reiterer, A., Sinn, G., Stanzl-Tschegg, S.E., Effects of surface
ageing on wettability, surface chemistry and adhesion of wood. Holz Roh
Werkst., 62, 273–280, 2004.
45. Odrasková, M., Szalay, Z., Ráhel, J., Zahoranová, A., Cernák, M., Diffuse
coplanar surface barrier discharge assisted deposition of water repellent
films from N2/HMDSO mixtures on wood surface, in: Proceedings 28th
International Conference on Phenomena in Ionized Gases, M. Simek and
P. Sunka (Eds.), IOP Publishing Ltd, pp. 803–806, 2008.
46. Frihart, C.H., Wood adhesion and adhesives, in: Handbook of Wood Chemistry
and Wood Composites, R.M. Rowel (Ed.), CRC Press, Boca Raton, FL, 2005.
47. Riedl, B., Angel, C., Prégent, J., Blanchet, P., Stafford, L., Wood surface
modification by atmospheric-pressure plasma and effect on waterborne
coating adhesion. J. Lignocellulose (Iran), 2, 292–306, 2013.
48. Denes, F.S., Cruz-Barba, L.E., Manolache, S., Plasma treatment of wood,
in: Handbook of Wood Chemistry and Wood Composites, R.M. Rowel (Ed.),
CRC Press, Boca Raton, FL, 2005.
49. Wascher, I.R., Avramidis, G., Kühn, C., Militz, H., Viöl, W., Plywood made
from plasma-treated veneers: Shear strength after shrinkage swelling stress.
Int. J. Adhes. Adhes., 78, 212–215, 2017.
238 Adhesives for Wood and Lignocellulosic Materials

50. Suurnakki, A., Buchert, J., Gronqvist, S., Mikkonen, H., Peltonen, S.,
Viikari, L., Bringing new properties to lignin rich fiber materials. VTT
Symposium, 244, 61–70, 2006.
51. Kudanga, T., Prasetyo, N., Sipila, J., Nyanhongo, G.S., Guebitz, G.M.,
Enzymatic grafting of functional molecules to the lignin mode dibenzodioxo-
cin and lignocellulosic material. Enzyme Microb. Tech., 46, 272–280, 2010.
52. Kudanga, T., Prasetyo, E.N., Sipilä, J., Guebitz, G.M., Nyanhongo, G.S.,
Reactivity of long chain alkylamines to lignin moieties: Implications on
hydrophobicity of lignocellulose materials. J. Biotechnol., 149, 81–87, 2010.
53. Yamaguchi, H., Nagamori, N., Sakata, I., Application of the dehydrogena-
tive polymerization of vanillic acid to bonding of woody fibers. Mokuzai
Gakkaishi, 37, 220–226, 1991.
54. Yamaguchi, H., Maeda, Y., Sakata, I., Application of phenol dehydrogena-
tive polymerization by laccase to bonding among woody fibers. Mokuzai
Gakkaishi, 38, 931–937, 1992.
55. Yamaguchi, H., Maeda, Y., Sakata, I., Bonding among woody fibers by use of
enzymatic phenol dehydrogenative polymerization. Mokuzai Gakkaishi, 40,
185–190, 1994.
56. Roman-Aguirre, M., Marquez-Luceroc, A., Zaragoza-Contreras, E.A.,
Elucidating the graft copolymerization of methyl methacrylate onto
wood-fiber. Carbohydr. Polym., 55, 201–210, 2004.
57. Wang, C. and Piao, C., From hydrophilicity to hydrophobicity: A critical
review—part II: Hydrophobic conversion. Wood Fiber Sci., 43, 41–56, 2011.
58. Bartzoka, V., Brook, M.A., Valentini, D., McDermott, M.R., Surface
Interactions between Proteins and Silicon Polymers: Physical and Covalent
Adhesion. 70th ACS Colloid and Surface Symposium, Potsdam NY, Abstract
147 1996.
59. Brook, M.A., Jiang, J., Heritage, P., Bartzoka, V., Underdown, B., McDermott,
M.R., Silicone–protein interaction at the interface between a functional sili-
cone and a protein/starch microparticle. Langmuir, 13, 23, 6279–6286, 1997.
60. Bartzoka, V., Brook, M.A., McDermott, M.R., Protein–silicone interactions:
How compatible are the two species? Langmuir, 14, 7, 1887–1891, 1998.
61. Bartzoka, V., McDermott, M.R., Brook, M.A., Protein–silicone interactions.
Adv. Mater., 11, 3, 257–259, 1999.
62. Bartzoka, V. and Brook, M.A., Stable Silicone–Protein Emulsions: New
Routes to Topical Delivery of Proteins. Society of Cosmetic Chemists
Conference, Toronto, ON, 2000.
63. Sebe, G. and Brook, M.A., Hydrophobization of wood surfaces—Covalent
grafting of silicone polymers. Wood Sci. Technol., 35, 269–282, 2001.
64. Yu, F., Yang, W., Song, J., Chen, L., Investigation on hydrophobic modifica-
tion of bamboo flour surface by means of atom transfer radical polymeriza-
tion method. Wood Sci. Technol., 48, 289–299, 2014.
65. Gvishi, R., Fast sol–gel technology: From fabrication to applications.
J. Sol-Gel Sci. Technol., 50, 241–253, 2009.
11
The Chemistry of Condensed Tannins

11.1 Introduction
Condensed tannin extracts consist of flavonoid units that have undergone
varying degrees of condensation. They are invariably associated with their
immediate precursors (flavan-3-ols, flavan-3,4-diols), other flavonoid
analogs [1, 2], carbohydrates, and traces of amino and imino acids [3].
Monoflavonoids and nitrogen-containing acids are present in concentra-
tions that are too low to influence the chemical and physical characteris-
tics of the extract as a whole. However, the simple carbohydrates (hexoses,
pentoses, and disaccharides) and complex glucuronates (hydrocolloid
gums) as well as oligomers derived from hydrolyzed hemicelluloses are
often present in sufficient quantity. Equally, carbohydrate chains of various
length [4, 5] are also sometimes linked to flavonoid units in the tannin.
O

O
n
OH
OH
HO HO O
O OH
O O
O OH
OH O O
m
OH
OH O

OH

All these materials are often present in sufficient quantities to decrease


and/or increase viscosity, and excessive variation in their percentages alters

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (239–266) ©  2019
Scrivener Publishing LLC

239
240 Adhesives for Wood and Lignocellulosic Materials

the physical properties of the natural tannin extract independently of the


contribution of the degree of condensation of the tannin.
Monoflavonoids, also known as “phenolic non-tannins” (to which flavo-
noid dimers belong too), represent the most studied group in the commer-
cially important tannin extracts because of their relative simplicity. They
comprise flavan-3,4-diols, flavan-3-ols, dihydroflavonoids (flavonols), fla-
vanones, chalcones, and coumaran-3-ols, thus most of the known classes
of flavonoid analogs [1, 2, 6, 7]. Typical are those of the black mimosa bark
extract (Acacia mearnsii, formerly mollissima, de Wildt) where the four pos-
sible combinations of resorcinol and phloroglucinol (A-rings) with catechol
and pyrogallol (B-rings) coexist, although these monoflavonoids consti-
tute a minor percentage (3–5%) of the total phenolics of the tannin extract
[1]. In mimosa tannin extract, only flavan-3,4-diols and certain flavan-
3-ols (catechin) participate in tannin formation.

OH
OH
OH
8 B
HO 7 O 8 B
OH OH 7 O
A OH
6
4 A
5 OH 6
4
5 OH
profisetinidin prorobinetinidin
Mw = 272.3 Da Mw = 288.3 Da
OH
OH OH
8 B B
HO 7 O HO 7
8
O
OH OH
A A
6 4 4
6
5 OH 5 OH
OH procyanidin OH prodelphinidin
Mw = 288.3 Da Mw = 304.3 Da

In other tannins, epicatechin, delphinidin, and catechin gallate are


present, always in minor percentages but with some exception. The main
exception is cube gambier tannin extract where monoflavonoids, mainly
catechin, can constitute up to 50% of the total extract. In mimosa bark
tannin extract, each of the four combinations of resorcinol and phloro-
glucinol (A-rings) with catechol and pyrogallol (B-rings) are present. In
this tannin, the main polyphenolic pattern is represented by flavonoid
analogs based on roninetinidin, and thus on resorcinol A-ring and pyro-
gallol B-ring. This pattern is reproduced in approximately 70% of the phe-
nolic part of the tannin. The secondary but parallel pattern is based on
fisetinidin and thus on resorcinol A-rings and catechol B-rings. This rep-
resents about 25% of the total polyphenolic bark fraction. Superimposed
on these two predominant patterns are two minor groups of A- and B-ring
The Chemistry of Condensed Tannins 241

combinations. These are based on phloroglucinol (A-ring)–pyrogallol


(B-ring) (gallocatechin/delphinidin) and on phloroglucinol (A-ring)–cat-
echol (B-ring) flavonoids (catechin/epicatechin). These four patterns con-
stitute 65–84% of commercial mimosa bark extract. The remaining parts
of mimosa bark extract are the so-called “non-tannins”, this definition
coming from the leather industry where any polyphenolic oligomer higher
and comprising atrimer is considered a “tannin”. It must be pointed out
that the percentage of non-tannins varies considerably from tannin extract
to tannin extract. For example, pecan nut tannin extract, a predominantly
delphinidin tannin, contains no more than 5% of non-phenolic non-tan-
nins [8,  9]. The percentages are different even for different commercial
tannins [8, 9]. The non-phenolic non-tannins can be subdivided into car-
bohydrates, hydrocolloid gums, and some amino and imino acid fractions
[9, 10]. For example, in commercial mimosa bark extract, the carbohy-
drates 1-pinitol and sucrose predominate, with glucose in a smaller pro-
portion [9, 10]. The hydrocolloid gums contribute between 3% and 8%
to the total weight of commercial mimosa bark extract and are a major
contributor to its viscosity [9, 10]. Imino acids such as L-pipecolic acid,
L-4-hydroxy-trans-pipecolic acid, and L-proline and traces of the amino
acids arginine, alanine, aspartic acid, glutamic acid, and serine have also
been reported [9, 10].
Similar flavonoid A- and B-ring patterns, although slightly different,
have been found for the other major commercial tannin extract, quebracho
(Schinopsis lorentzii and balansae) wood tannin extract, where a predomi-
nance of fisetinidin rather than robinetinidin in the constituent flavonoid
units has been determined [11]. In quebracho extract, there is apparent
absence of quercetrin and myretrecin [6, 9, 11, 12] and much lower propor-
tion or even absence of catechin and gallocatechin (delphinidin) and thus
practically lack of or much lower level of phloroglucinol-like A-rings. This
difference becomes very important from a structural point of view, as it
will be indicated later, as while the quebracho tannins have mainly a linear
structure, mimosa tannin presents a more branched structure, with very
important effects on their viscosity, their stability, and their non-leather
use. Similar couplings of A- and B-ring types occur also in Douglas-fir and
hemlock bark tannins.
However, different A- and B-ring couplings occur in pine and other bark
tannin species [4, 9, 13–20]. In this, only two main patterns occur, pre-
dominantly phloroglucinol A-ring with catechol B-ring and two secondary
patterns in much lower proportion of phloroglucinol A-ring with phenol
B-ring (afzelechin) or of fisetinidin (resorcinol A-ring, catechol-B-ring)
[9, 13–20]. In several procyanidin tannins such as some pine tannins and
242 Adhesives for Wood and Lignocellulosic Materials

exotic African woods, catechin gallate or gallocatechin gallate also occurs


as a constitutive unit of the tannin.
OH
OH OH OH
HO O HO O HO O
OH OH OH
O O O
OH OH OH OH
O O O
OH OH OH
OH OH OH

Roux et al. [6] have shown that condensation of robinetinidin and fise-
tinidin flavonoid units is based on a condensation forming a 4,6 interflavo-
noid link, following an initial 4,8 interflavonoid link formation between a
phloroglucinol A-ring type unit, such as catechin, and a resorcinol A-ring-
type units. For a time, it was thought that the positioning of the phloroglu-
cinolic flavonoid unit was as the “lower” terminal unit of a 4,8 interflavonoid
linkage. While extracts such mimosa show oligomers of this type [9, 21, 22],
the latter finding of the so-called “angular” tannins indicated that the phlo-
roglucinolic unit was not just the lower terminal unit of an oligomer but was
a central unit to which two resorcinol A-ring-type units are linked [9, 23].
OH

HO O
OH

OH
OH

HO 8 9 O
HO 2
OH
7
6
5 10 3
4 OH

OH
O
OH Angular tannin

OH

OH
The Chemistry of Condensed Tannins 243

Moreover, an even later analysis derived from comparative applied


results on the reactivity of aldehydes with different tannins [8, 9, 24]
showed that the phloroglucinol A-ring type units, be it catechin, epicat-
echin, or gallocatechin (delphinidin), was a branching point in the oligo-
mers of some tannin type (mimosa), leading to the concept of such tannins
having been heavily “branched” rather than just “linear”.
OH

HO O
OH

OH
OH

HO 8 9 O
HO 7 2
OH

6
5 10 4 3 OH
OH
OH
O
OH HO O
OH

OH

OH

OH

This was evident as two different rates of reaction with formaldehyde


did not occur but instead viscosity increase graphs as a function of time
are smooth exponential curves indicating that only one reaction site type
is present, with the phloroglucinol unit A-ring reactive sites being blocked.
This explained the particular practical differences in the use of quebra-
cho and mimosa tannins, with quebracho tannin sometimes being subject
to a reaction of partial depolymerization while mimosa is not. Mimosa is
heavily branched while quebracho is not, and thus the latter is more easily
subject to cleavage of the interflavonoid linkage under some drastic con-
ditions [8, 9, 24].
In the case of procyanidin and prodelphinidin-type tannins, which are
composed exclusively of phloroglucinol A-ring type units, the interflavo-
noid link is always 4,8, and thus all these oligomers are “linear”.
244 Adhesives for Wood and Lignocellulosic Materials

OH

HO O
OH

OH
OH
O
HO O
OH

OH
OH
OH
HO O
OH

OH

OH

11.2 Reactions of Condensed Flavonoid Tannins


The reactions with formaldehyde and other aldehydes will not be discussed
here, but other important reactions of tannin will.

11.2.1 Hydrolysis and Acid and Alkaline Condensation


When heated in the presence of strong mineral acids, tannins are subject
to two competitive reactions. One is degradative leading to anthocyani-
dins and catechin formation as illustrated in Scheme 11.1 [6, 9], whereas
the second is condensative as a result of hydrolysis of the p-hydroxybenzyl
ether links of the flavonoid heterocyclic ring. The carbonium ions formed
condense randomly with nucleophilic centers on other tannin units to
form “phlobatannins”, “phlobaphenes”, or “tanner’s red”, insoluble and no
further usable [25].
The Chemistry of Condensed Tannins 245

OH

HO O HO OH
OH

OH OH OH
OH
HO O HO O
OH OH

OH OH
OH OH

OH
HO HO OH
OH
OH

O O OH
Phlobatannin
OH
OH

Scheme 11.1

Even more complex insoluble structures, such as phlobaphenes, occur,


of formula

HO HO OH
OH

OH
HO

HO O O
OH

OH

OH O
HO

OH

OH
246 Adhesives for Wood and Lignocellulosic Materials

A system to stop the condensative reaction to go to phlobaphenes has


been found and used industrially [25], to greatly improve the percentage
extraction yield of tannins. This method consists in using a simple mole-
cule such as urea to block the carbonium ion formed.

OH

HO O HO OH
OH
OH OH
OH OH
HO O OH HO O
OH

OH OH
OH OH

NH2CONH2

H2NCONH
OH

HO OH OH
OH
OH
HO O
OH

OH
OH

Different tannins behave differently as well. For example, procyan-


idin tannins (phloroglucinol A-ring and catechol B-ring) such as pine
bark tannins favor cleavage of the interflavonoid linkage [8, 26], with the
oligomers linearity contributing to such a behavior. Predominantly pro-
delphinidin tannins, such as pecan nut tannin, also show cleavage of the
interflavonoid bond, but due to the effect of the pyrogallol B-ring, they
can also present easy cleavage and opening of the heterocyclic pyran ring
[8, 27, 28], under certain conditions up to markedly limiting interfla-
vonoid link cleavage. Predominantly profisetinidins, such as quebracho
wood tannin, tend to behave as prodelphinidins, again confirming that
interflavonoid link cleavage is facilitated by the “linearity” of the struc-
ture of the oligomers of these tannins. For mimosa tannin, where a cer-
tain degree of branching exists, interflavonoid link cleavage practically
does not occur, with cleavage and opening of the heterocyclic pyran ring
by far being the favorite reaction.
The Chemistry of Condensed Tannins 247

11.2.2 Sulphitation
Sulphitation of tannins is one of the oldest and most useful reactions
in condensed tannins chemistry. It can be useful or damaging if car-
ried out to an excessive extent, according to the final use to which the
tannin extract is destined [9, 29]. Sulphitation allows the preparation
of tannins of lower viscosity and increased solubility and thus easier to
handle.

OH

HO O - HO OH OH
OH SO3- OH

OH OH
SO3Na

Such effects are due to:

(i) The elimination of the heterocyclic ring, which is water


repellent. It must be remembered that tannin extract water
solutions are not true solutions but hydrocolloid suspen-
sions, in which part of the tannin molecule keeps the tan-
nin in solution while other parts tend to push the tannin
out of solution [9, 30].
(ii) The introduction of a sulphonic group and a further
hydroxygroup, both hydrophilic.
(iii) The decrease in polymer rigidity, steric hindrance, and
intermolecular hydrogen bonding obtained by the open-
ing of the heterocyclic ring.
(iv) Hydrolysis of the hydrocolloid gums [9, 31] and of the
interflavonoid bond (for some tannins) [9].
(v) The increase of reactivity towards aldehydes by the trans-
formation of the ether linkage of the heterocyclic ring
into a –OH linked on to the A-rings.

While the above are considerable advantages, excessive sulphitation


can be a distinct disadvantage for some applications. Thus, the introduc-
tion of an excessive proportion of sulphonate groups promote excessive
sensitivity to moisture and water and thus deterioration of resins, plastics
and adhesive bonds that are supposed to be moisture, water, or weather
resistant [9, 32].
248 Adhesives for Wood and Lignocellulosic Materials

11.2.3 Catechinic Acid Rearrangement


Procyanidin- and delphinidin-type tannins, but also other flavonoid tan-
nins, can be subject to a reaction of rearrangement to catechinic acid. This
reaction leads to the deactivation of several reactive sites of the tannin and
thus it is to be avoided. Some manipulation errors in tannin extraction
or in drying generally cause this rearrangement. The appearance of the
catechinic acid derived directly from a flavonoid monomer and the one
derived from a dimer are as follows:

HO OH
HO O OH

OH

OH HO O
O
OH

OH
OH
O
OH

OH
OH

While the catechinic acid rearrangement is easily shown to occur in


model compounds where the reaction is carried out in solution [9, 33], it
is much less evident and more easily avoidable in tannin extracts where
the colloidal nature of the extract limits markedly its occurrence. This is
fortunate as otherwise some fast-reacting tannins such as pine, pecan,
cube gambier, etc. could not be used to produce resins, adhesives, and
other thermosetting plastics as instead they have been successfully used
for [27, 34–38]. It can be easily determined if a tannin extract is affected
by this rearrangement both by its loss of reactivity and by 13C NMR anal-
ysis [39].

11.2.4 Catalytic Tannin Autocondensation


Polyflavonoid tannins have been found to autocondense and harden when
in the presence of particular compounds acting as catalysts. Foremost is
the catalytic effect of small amounts (2–3%) of silica smoke, or nanosilica
or silicates at high pH [40]. This reaction is rapid and is markedly exo-
thermic, a concentrated solution of tannin at 40–50% in water gelling and
The Chemistry of Condensed Tannins 249

hardening at pH 12 and 25°C in 20–30 min. The strong exothermicity of


the reaction leads to this result as the temperature increases several tens of
degrees in a short period [40].

Si(OH)4
Si(OH)4 O
OH
B HO O
HO O OH
OH
OH
OH
OH
OH

OH
HO OH OH Si(OH)4
OH O
HO O
OH OH
OH HO
OH
OH -δ OH
HO OH OH OH
OH

O
OH

OH
OH

Small amounts of boric acid and AlCl3 were found to have the same
effect [40] but are much less exothermic. Interestingly, even the presence
of lignocellulosic material, such as placing the tannin on the wood sur-
face, has a catalytic effect on tannin autocondensation [41–43]. Other
coreactants seem to further favor the reaction [41–43]. This mechanism
of tannin autocondensation appears to depend on the Lewis acid behav-
ior of the reagents. It involves Lewis acid acceptance of electrons from
the ether oxygen of the flavonoid heterocyclic pyran ring with facili-
tation of base-induced pyran ring opening. The reactive C2 site cre-
ated by the heterocyclic pyran ring opening proceeds to condense with
the reactive A-ring of a flavonoid unit on another chain. This denies
to the flavonoid any possibility of rearrangement to catechinic acid or
phlobaphenes. In the SiO2 catalysis, Si has been shown to go through
a coordination state of 5. The portion of Si that has not been able to
complete the reaction due to premature hardening remains attached to
the flavonoids, in this coordination state, in the hardened network. The
Si portion that has completed the reaction and caused the hardening
reverts instead to SiO2 and is detached from the flavonoid. It restarts
the cycle to lead to complete hardening up to when diffusional prob-
lems do not stop the reaction. The reaction is rather exothermic, with
250 Adhesives for Wood and Lignocellulosic Materials

the marked increase in temperature observed contributing to the self-


acceleration of the reaction. The SiO2, boric acid, and AlCl3 catalyses
of tannin autocondensation also reverse the relative ease of cleavage
between interflavonoid bond and heterocyclic ring opening in mainly
procyanidin- and prodelphinidin-type tannins. This reaction has been
used to good effect for no-aldehyde interior-grade tannin adhesives for
wood panels [44]. As the cross-linking obtained is only acceptable for
interior joints, coupling with traditional polycondensation reactions
can be implemented easily [42, 43].

11.2.5 Tannin Complexation of Metals


Tannins readily complex metal ions [45]. This characteristic is at the base
of a number of industrial applications. Thus, this characteristic is used
to capture or precipitate toxic metals in water [46, 47], to isolate a rare
metal such as Germanium from the copper matrix where it is mined,
for paint primers for metal application, and for several other applica-
tions. An old example is the formation of Fe complexes, used to prepare
intensely black/violet inks by formation of ferric tannates. These coordi-
nation complexes are due to the ortho-diphenol hydroxyl groups on the
tannin B-rings.

O O
III
Fe
HO O O OH
O O

OH HO

11.2.6 Tridimensional Structure


While there is an abundant literature in chemical journals on the tridi-
mensional structure of flavonoid monomers, one point in which only scant
literature exists is on the three-dimensional spatial configuration of flavo-
noid oligomers. Only one molecular mechanics study on this subject exists
[48]. This study shows the correlation that exists between the applicability
of these materials and their 3D structure. For example, a tetraflavonoid of
4,8-linked catechins, all 3,4-cis, is in helix configuration, and when looked
along the helix axis, a characteristic structure presenting all the 4 B-rings
pointing outwards appears.
The Chemistry of Condensed Tannins 251

Such a structure, rendering particularly available the hydroxyl groups of


the B-rings obviously facilitates their use and reactions, such as adhesion
to a lignocellulosic substrate, formation of metallic coordination complexes
[45–47], formation of polyurethanes with and without isocyanates [49, 50],
and others where reaction of the B-ring is of interest [51], such as cross-
linking at pH 10 and higher. Conversely, the “spring-like” structure con-
tributes to some of the “resuscitation” behaviors of some plants by holding
together the cellular walls and avoiding cellular walls cracking on drying [52].

11.2.7 Reactivity and Orientation of Electrophilic


Substitutions of Flavonoids
The relative accessibility and reactivity of flavonoid units is of interest for
their use in resins and adhesives. The C8 site on the A-ring is the first one
to react, for example, with an aldehyde, and is the site, when free of high-
est reactivity [6, 8]. The C6 site on the A-ring is also very reactive but less
than the C8 site as this latter presents lower steric hindrance too [6, 8]. The
reactions involve in general only these two sites on the A-ring. The B-ring
is particularly unreactive. A low degree of substitution at the 6 site of the
B-ring can occur. In general, at higher pHs, such as pH 10, the B-ring starts
to react too, contributing to cross-linking as well [51, 53].
1
3 2 3
OH OH
6' 6'
HO 8 O HO O
OH 8 OH
6 6
OH OH
2 OH 1
252 Adhesives for Wood and Lignocellulosic Materials

Thus, for catechins and phloroglucinol A-ring type flavonoids, the reac-
tivity sequence of the sites is C8 > C6 > C6 when these are free. For rob-
inetinidin and fisetinidin and thus for resorcinol A-ring-type flavonoids,
the reactivity sequence is modified to C6 > C8 > C6 due to the greater
accessibility and lower possibility of steric hindrance of the C6 site [6, 9].
The curve of gel time of flavonoid tannins with aldehydes has always the
shape of a bell curve. The longer gel time is at around pH 4 and the fastest
gel times are at lower pHs and higher pHs. The curve reaches an almost
asymptotic plateau of very high reactivity and short gel time at around pH
10 and higher and a fast reactivity too at pHs lower than 1–2 [8, 9]. The
shape of this curve is always the same, but the gel time value is different
for different tannins, being slower for mimosa and quebracho, and much
faster for procyanidin-type tannins (such as pine) [54, 55].

11.2.8 Influence of Tannin Colloidal Behavior on Reactions


Water solutions of 40–50% polyflavonoid tannin extracts appear to be in a
colloidal state as indicated by their zeta-potentials [55]. This is caused by
both the presence of noticeable proportions of hydrocolloid gums (frag-
ments of hemicelluloses) as well as the presence of higher-molecular-mass
tannins. 13C NMR has confirmed that, during chemical treatment of the
tannin extracts, reactions occur in these colloidal solutions that would
not be likely to occur in noncolloidal solutions such as used in model
compound experiments. These reactions center on reactions occurring in
the part of the tannin that is the non-aqueous environment within the
colloidal micelles, away from water, within which reagents can migrate.
An example of this is the role of an organic anhydride by addition of ace-
tic or maleic anhydride to hot, concentrated water solutions of tannin, a
reaction used to increase the reactivity of the tannin [55, 56]. While part
of the anhydride is hydrolyzed to acid in water, part of it does instead react
within the colloidal micelles to give acetylation and maleation of some
flavonoids of the tannin [55, 56] contributing to the marked improvement
of reactivity towards aldehydes of the tannin by allowing an alpha-set
approach [57]. Such reactions appear to be particularly beneficial to the
quebracho and mimosa tannin extracts and have some noticeable positive
effects on the higher-reactivity procyanidin-type tannin extracts, such
as pine bark tannin, but less due to its already much higher reactivity.
However, they may have deleterious effects on higher-reactivity tannin
extracts such as the pecan nut prodelphinidin tannin. The reason for the
latter behavior is the very low level, near absence of colloidal gum in the
extract, and thus very low level of colloidal state, if any. These results are in
The Chemistry of Condensed Tannins 253

line with the established zeta-potentials of the different tannin extracts


measured [55, 56].

11.2.9 New and Unusual Tannin Reactions


Recently, a number of reactions of tannins that could be useful for a
number of different applications have come to light. The first of these is
the reaction of flavonoid tannins with concentrated aqueous ammonia
[58]. Catechin was also used as a model compound and treated under the
same conditions mimosa tannin extract was treated. Solid-state 13C NMR
and matrix-assisted laser desorption/ionization time of flight (MALDI-
TOF) spectroscopy showed that, unlike what was recently thought [59],
amination is not always regioselective and leads to the conversion of one
single –OH group in C4 into a –NH2 group. New reactions have been
evidenced, clearly leading to multiamination of several phenolic hydrox-
ygroups, heterocycle opening, and oligomerization and cross-linking
through the formation of –N= bridges between flavonoid units, as shown
here:
OH OH
HO HO N OH
OH
OH
HO
OH
OH 589Da

The amination reaction of condensed tannins was used, among


others, to totally eliminate synthetic materials in the preparation of
non-isocyanate polyurethanes derived from tannins [60].
Follow-up of the reaction with ammonia was the development of rapid
cross-linking by reaction of tannin extract with diamines and polyamines
[61]. Reaction of a condensed flavonoid tannin, namely, mimosa tannin
extract with a hexamethylene diamine, has been investigated. Catechin
was also used as a flavonoid model compound and treated under similar
conditions. Solid-state CP-MAS 13C NMR and MALDI-TOF mass spec-
troscopy showed that polycondensation compounds leading to resins
were obtained by the reaction of the amines with the phenolic hydroxy
groups of the tannin. Simultaneously, a second reaction leading to the
formation of ionic bonds between the two groups occurred. These new
reactions have been shown to clearly lead to the reaction of several phe-
nolic hydroxyl groups, and flavonoid unit oligomerization, to form hard-
ened resins. MALDI-TOF analysis allowed us to observe the presence of
compounds of the type
254 Adhesives for Wood and Lignocellulosic Materials

HO

OH O OH
HO
HO O NH
NH HO OH
O NH
OH HO NH2
OH
OH
HO H
HO O
NH O
Na
NH2
OH
OH

Clearly indicating how polymerization and cross-linking occurred.


The third very novel reaction is based on the reaction, oligomeriza-
tion, and cross-linking of tannins by triethyl phosphate (TEP) [62] in
the presence or absence of ammonia (this latter being preferable to the
yield). Reaction of condensation and cross-linking of catechin monomer
as a model of condensed (flavonoid) tannin extracts and of mimosa tan-
nin itself, as well as of resorcinol with TEP, have been investigated. Solid-
state CP-MAS 13C NMR, 31P NMR, and MALDI-TOF spectroscopy studies
revealed that reaction occurs mainly on the C3 of the flavonoid heterocycle
ring and on the aromatic C4 and C5 carbons of the flavonoid B-ring,
while TEP does not appear to react on the A-ring. Structures of the type

O O
P
O O

HO O H
O

O
O OH
P
OH O
O

O OH

O
O
P
O OH
O
The Chemistry of Condensed Tannins 255

were obtained and the tannin was cross-linked, with or without being
first reacted with ammonia. The resin so obtained can produce hard
thermoset plastics and films resistant up to temperatures in excess of
400°C.
A difference in the relative proportions of these two reaction sites
for tannin and catechin has been noticed. The main reaction for the
tannin appears to occur on the C3 site of the heterocycle ring while cat-
echin monomer reacts principally on the OH of B-ring. This difference
could be explained by the lower mobility of the tannin, due to its higher
molecular weight and to its colloidal state. The reactions appear to be
dependent on the temperature. The reaction appears to have a tem-
perature of activation below which it does not appear to occur. Thus, it
occurs readily at 185°C but does not at 100°C. This aspect needs further
investigation. FTIR analysis confirmed the results of the MALDI-TOF
and NMR analysis. According to thermogravimetric analysis, materials
obtained from the reaction of tannin with TEP showed high thermal
stability. In this context, the potential of this reaction has been evalu-
ated for the production of new heat-resistant biomaterials and lacquers
[62–64].

11.2.10 Modern Instrumental Methods of Analysis


Analysis of tannin features are done by a number of traditional meth-
ods of all types. However, two instrumental methods have particularly
distinguished themselves in identifying structural features and oligomer
type, size, and distribution to better understand these materials. These
are 13C NMR, both in liquid phase or solid state, and MALDI-TOF mass
spectroscopy. These two techniques are particularly useful not only in
the analysis of tannins but also in the analysis of their reaction prod-
ucts and of the resins derived from them. Below are given the charac-
teristic 13C NMR shifts of a catechin monomer and a range of tannin
extracts from which the structural features of any condensed tannin can
be deduced (Table 11.1) [8].
Equally, MALDI-TOF techniques have allowed the determination
of oligomer type, molecular weight, and distribution of a number of
tannin extracts and products obtained from their reactions [14–20,
65–71]. From a MALDI-TOF spectrum, one can distinguish mono-
mers, dimers, trimers, higher oligomers, and so on. Figure 11.1 shows
MALDI-TOF spectra of a commercial mimosa tannin extract and Table
11.2 shows the more interesting oligomer assignments derived by such
spectra.
256

Table 11.1 Comparative 13C NMR shift assignments and relative band intensities (%) for pure catechin and five types of polyflavonoid
tannin extracts.
Assignment (ppm)

Phloro-
glucinol
inter- Catechinic
C5, C7 C9 C3', C4' C1 C6'* C5', C2'* flavonoid C10 C6 C8 C2 C3 acid C4
(156–157) (155) (145–146) (131) (120–121) (115–117) (110) (101) (96–98) (95–96) (81–82) (67–68) (31–32) (27–28)

Pure 97, 100 58 72, 76 73 97 98, 96 – 68 88 97 92 93 – 80


catechin

Mimosa 53 30 100 44 25 26 51 19 21 31 – – – 21
bark

Quebracho 63 33 100 66 56 99 40, 52 25 20 20 – – – 32


wood

Pine bark 71 71 100 49 39 76 60 37 47 23 – – – –

Pecan nut 66 70 100 62 30 33 66 42 40 29 – – 79 37


pith

Gambier 100 50 100 55 69 106, 95 27 50 60 28 – – 61 61


Adhesives for Wood and Lignocellulosic Materials

Source: Ref. [8].


The Chemistry of Condensed Tannins 257

100 375

90

80 890
70

60 551
% int.

50 1179
40

30 1467

20
488 727 1756
10 988 2045
2333
0
400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400 3600 3800 4000
Mass/Charge
(a)

890
100
906 305 Da
90
80 289 Da
70
60 1179
273 Da
% int.

50 1195
874
40 602 1163

30
922
20 664 727
634 700 750 781 814 858 1147 1211
10 687 759 840 943 988

0
600 650 700 750 800 850 900 950 1000 1050 1100 1150 1200 1250 1300
Mass/Charge
(b)

Figure 11.1 MALDI mass spectrum of (a) natural mimosa tannin extract. (b) Details of
the 600–1300 Da range with indication of the relevant 288-Da repeat unit [66].

11.3 Conclusions
The chemistry and characteristic reactions of condensed flavonoid tan-
nins have been the basis for their extended industrial utilization. It is
on the basis of this chemistry that many heavily or totally biosourced
materials have been developed. Among these are industrialized wood
panel adhesives [31, 34, 37, 72], industrialized wood laminating adhe-
sives [73–75], fire-resistant biosourced rigid and flexible foams [76–87],
foams for acoustic isolation [88], hard plastics [89], grinding disks for
angle grinders [90], automotive brake pads [91], paper impregnating res-
ins [92–94], high-pressure laminates [95], impregnated fiber composites
258 Adhesives for Wood and Lignocellulosic Materials

Table 11.2 MALDI fragmentation peaks for industrial mimosa tannin extract
of Figure 11.1. Note that the predominant repeat units in this tannin is 288 Da,
indicating that this tannin is predominantly a prorobinetinidin [66].
M + Na+ M + Na+ Unit type (MW)
(exp.) (calc.) 274 290 306
Dimers
602 601 – 2 –
Trimers
858 857 2 1 –
874 873 1 2 –
or 2 – 1 Angular tannin
*890 889 1 1 1
or 1 – 2
*906 905 – 2 1 Angular tannin
or 1 – 2 Angular tannin
922 921 – 1 2 a “diangular” structure
Tetramers
1147 1145 2 2 –
or 3 – 1
1163 1161 1 3 –
or 2 1 1
*1179 1177 – 4 –
or 1 2 1
or 2 – 2
1195 1193 – 3 1 Angular tannin
or 1 1 2
1211 1209 – 2 2 Angular tannin
(Continued)
The Chemistry of Condensed Tannins 259

Table 11.2 MALDI fragmentation peaks for industrial mimosa tannin extract
of Figure 11.1. Note that the predominant repeat units in this tannin is 288 Da,
indicating that this tannin is predominantly a prorobinetinidin [66]. (Continued)
M + Na+ M + Na+ Unit type (MW)
(exp.) (calc.) 274 290 306
or 1 – 3 a “diangular” structure
Pentamers
1467 1465
Hexamers
1756 1753
Heptamers
2045 2041
Octamers
2333 2329
*Dominant fragment.

[96, 97], biosourced wood preservatives [98, 99], non-isocyanate poly-


urethane surface finishes and resins [49, 100, 101], and others such as
medical/pharmaceutical applications [102–105] (Pizzi 2008), antioxi-
dant applications [106], and precipitation of heavy metals from wastewa-
ters [46, 47]. There is hope that on the basis of the same chemistry, many
other industrial products of progressively higher added value may also be
developed in the future.

References
1. Drewes, S.E. and Roux, D.G., Condensed tannins. 15. Interrelationships of
flavonoid components in wattle-bark extract. Biochem. J., 87, 167–172, 1963.
2. Roux, D.G. and Paulus, E., Condensed tannins. 8. The isolation and distribu-
tion of interrelated heartwood components of Schinopsis spp. Biochem. J., 78,
785–789, 1961.
3. Saayman, H.M. and Roux, D.G., The origins of tannins and flavonoids in
black-wattle barks and heartwoods, and their associated ‘non-tannin’ com-
ponents. Biochem. J., 97, 794–801, 1965.
260 Adhesives for Wood and Lignocellulosic Materials

4. Drovou, S., Pizzi, A., Lacoste, C., Zhang, J., Abdalla, S., Al-Marzouki, F.M.,
Flavonoid tannins linked to long carbohydrate chains—MALDI TOF anal-
ysis of the tannin extract of the African locust bean. Ind. Crops Prod., 67,
25–32, 2015.
5. Abdalla, S., Pizzi, A., Ayed, N., Chrrier, F., Bahabry, F., Ganash, A., MALDI-
TOF and 13C NMR analysis of Tunisian Zyzyphus jubjuba root bark tannins.
Ind. Crops Prod., 59, 277–281, 2014.
6. Roux, D.G., Ferreira, D., Hundt, H.K.L., Malan, E., Structure, stereochemis-
try, and reactivity of natural condensed tannins as basis for their extended
industrial application. Appl. Polym. Symp., 28, 335–353, 1975.
7. Roux, D.G., Recent advances in the chemistry and chemical utilization of the
natural condensed tannins. Phytochemistry, 11, 1219–1230, 1972.
8. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York,
1994.
9. Pizzi, A., Tannin based adhesives, in: Wood Adhesives Chemistry and
Technology, A. Pizzi (Ed.), Marcel Dekker, New York, 1983.
10. Roux, D.G., Modern Applications of Mimosa Extract, pp. 34–41, Leather
Industries Research Institute, Grahamstown, South Africa, 1965.
11. Mitsunaga, T., Doi, T., Kondo, Y., Abe, I., Color development of proantho-
cyanidins in vanillin–hydrochloric acid reaction. J. Wood Sci., 44, 125–130,
1998.
12. Clark-Lewis, J.W. and Roux, D.G., Natural occurrence of enantiomorphous
leucoanthocyanidian: (+)-mollisacacidin (gleditsin) and quebracho(–)-
leucofisetinidin. J. Chem. Soc., 0, 1402–1406, 1959.
13. Porter, L.J., Extractives of Pinus radiata bark. 2. Procyanidin constituents.
N. Z. J. Sci., 17, 213, 1974.
14. Navarrete, P., Pizzi, A., Pasch, H., Rode, K., Delmotte, L., MALDI-TOF and
13
C NMR characterisation of maritime pine industrial tannin extract. Ind.
Crops Prod., 32, 105–110, 2010.
15. Ucar, M.M., Ucar, G., Pizzi, A., Gonultas, O., Characterisation of Pinus brutia
bark tannin by MALDI-TOF and 13C NMR. Ind. Crops Prod., 49, 679–704,
2013.
16. Abdalla, S., Pizzi, A., Ayed, N., Charrier-El-Bouthoury, F., Charrier, B.,
Bahabry, F., Ganash, A., MALDI-TOF analysis of Aleppo pine (Pinus halep-
pensis) bark tannin. Bioresources, 9, 3396–3406, 2014.
17. Abdalla, S., Pizzi, A., Ayed, N., Charrier, B., Bahabry, F., Ganash, A., MALDI-
TOF and 13C NMR analysis of Tunisian Zyzyphus jubjuba root bark tannins.
Ind. Crops Prod., 59, 277–281, 2014.
18. Saad, H., Charrier-El-Bouthoury, F., Pizzi, A., Rode, K., Charrier, B., Ayed, N.,
Characterization of pomegranate peel tannin extractives. Ind. Crops Prod.,
40, 239–246, 2012.
19. Navarrete, P., Pizzi, A., Pasch, H., Rode, K., Delmotte, L., Characterisation
of two maritime pine tannins as wood adhesives. J. Adh. Sci. Technol., 27,
2462–2479, 2013.
The Chemistry of Condensed Tannins 261

20. Vazquez, G., Pizzi, A., Freire, M.S., Santos, J., Antorrena, G., Gonzalez-
Alvarez, J., MALDI-TOF, HPLC-ESI-TOF and 13C NMR characterisation of
chestnut (Castanea sativa) shell tannins. Wood Sci. Technol., 47, 523–535,
2013.
21. Roux, D.G., Recent advances in the chemistry and chemical utilization of the
natural condensed tannins. Phytochemistry, 11, 1219–1230, 1972.
22. Hundt, H.K.L. and Roux, D.G., Condensed tannins: Determination of the
point of linkage in ‘terminal’(+)-catechin units and degradative bromination
of 4-flavanylflavan-3,4-diols. J. Chem. Soc., Chem. Comm., 0, 696–698, 1978.
23. Botha, J.J., Ferreira, D., Roux, D.G., Condensed tannins. Circular dichroism
method of assessing the absolute configuration at C-4 of 4-arylflavan-3-ols,
and stereochemistry of their formation from flavan-3,4-diols. J. Chem. Soc.
Chem. Comm., 0, 698–700, 1978.
24. Pizzi, A., Natural phenolic adhesives 1: Tannin, in: Handbook of Adhesive
Technology, 2nd edn, A. Pizzi and K.L. Mittal (Eds.), pp. 573–598, Marcel
Dekker, New York, 2003.
25. Sealy-Fisher, V.J. and Pizzi, A., Increased pine tannins extraction and wood
adhesives development by phlobaphenes minimization. Holz Roh Werkstoff,
50, 212–220, 1992.
26. Hemingway, R.M., Laks, P.E., McGraw, G.W., Kreibich, R.E., Wood Adhesives
in 1985: Status and Needs, Forest Products Society, Madison, Wisconsin,
1986.
27. Pizzi, A. and Stephanou, A., Comparative and differential behaviour of pine
vs. pecan nut tannin adhesives for particleboard. Holzforsch. Holzverwert.,
45, 2, 30–33, 1993.
28. McGraw, G.W., Rials, T.G., Steynberg, J.P., Hemingway, R.W., Plant
Polyphenols, R.W. Hemingway and P.E. Laks (Eds.), Plenum Press, New York,
1992.
29. Richtzenhain, H. and Alfredsson, B., Uber Ligninmodellsubstanzen. Chem.
Ber., 89, 378, 1956.
30. Roux, D.G., Wattle Tannin and Mimosa Extract, Leather Industries Research
Institute, Grahamstown, South Africa, 1965.
31. Pizzi, A., Wattle-based adhesives for exterior grade particleboard. For.
Prod. J., 28, 12, 42–47, 1978.
32. Pizzi, A., Sulphited tannins for exterior wood adhesives. Colloid Polym. Sci.,
257, 37–40, 1979.
33. Ohara, S. and Hemingway, R.W., Condensed tannins: The formation of
a  diarylpropanol-catechinic acid dimer from base-catalyzed reactions of
(+)-catechin. J. Wood Chem. Technol., 11, 195–208, 1991.
34. Pizzi, A., Pine tannin adhesives for particleboard. Holz Roh Werkst., 40, 293,
1982.
35. Pizzi, A., Von Leyser, E.P., Valenzuela, J., Clark, J.G., The chemistry and devel-
opment of pine tannin adhesives for exterior particleboard. Holzforschung,
47, 164–172, 1993.
262 Adhesives for Wood and Lignocellulosic Materials

36. Pizzi, A. and Stephanou, A., Fast vs. slow-reacting non-modified tannins
extracts for exterior particleboard adhesives. Holz Roh Werkst., 52, 218–222,
1994.
37. Valenzuela, J., Von Leyser, E., Pizzi, A., Westermeyer, C., Gorrini, B.,
Industrial production of pine tannin-bonded particleboard and MDF. Eur. J.
Wood Prod., 70, 735–740, 2012.
38. Pizzi, A., Valenzuela, J., Westermeyer, C., Low-formaldehyde emission, fast
pressing, pine and pecan tannin adhesives for exterior particleboard. Holz
Roh Werkst., 52, 311–315, 1994.
39. Navarrete, P., Pizzi, A., Bertaud, F., Rigolet, S., Condensed tannin reactiv-
ity inhibition by internal rearrangements: Detection by CP-MAS 13C NMR.
Maderas, 13, 1, 59–68, 2011.
40. Meikleham, N., Pizzi, A., Stephanou, A., Induced accelerated autocondensa-
tion of polyflavonoid tannins for phenolic polycondensates, Part 1: 13C NMR,
29
Si NMR, X-ray and polarimetry studies and mechanism. J. Appl. Polym.
Sci., 54, 1827–1845, 1994.
41. Garcia, R., Pizzi, A., Merlin, A., Ionic polycondensation effects on the radical
autocondensation of polyflavonoid tannins—An ESR study. J. Appl. Polym.
Sci., 65, 2623–2632, 1997.
42. Garcia, R. and Pizzi, A., Polycondensation and autocondensation networks
in polyflavonoid tannins, Part 1: Final networks. J. Appl. Polym. Sci., 70,
1083–1091, 1998.
43. Garcia, R. and Pizzi, A., Polycondensation and autocondensation networks
in polyflavonoid tannins, Part 2: Polycondensation vs. autocondensation.
J. Appl. Polym. Sci., 70, 1093–1110, 1998.
44. Pizzi, A., Meikleham, N., Dombo, B., Roll, W., Autocondensation-based,
zero-emission, tannin adhesives for particleboard. Holz Roh Werkst., 53,
201–204, 1995.
45. Slabbert, N., Complexation of condensed tannins with metal ions, in:
Plant Polyphenols: Biogenesis, Chemical Properties, and Significance, R.W.
Hemingway and P.E. Laks (Eds.), Plenum Press, New York, 1992.
46. Tondi, G., Oo, C.W., Pizzi, A., Trosa, A., Thevenon, M.-F., Metal adsorption
of tannin-based rigid foams. Ind. Crops Prod., 29, 336–340, 2009.
47. Oo, C.W., Kassim, M.J., Pizzi, A., Characterization and performance of
Rhizophora apiculata mangrove polyflavonoid tannins in the adsorption of
copper (II) and lead (II). Ind. Crops Prod., 30, 152–161, 2009.
48. Pizzi, A., Cameron, F.A., Eaton, N.J., The tridimensional structure of poly-
flavonoids by conformational analysis. J. Macromol. Sci. Chem., A23, 4, 515–
538, 1986.
49. Thebault, M., Pizzi, A., Essawy, H., Baroum, A., Van Assche, G., Isocyanate
free condensed tannin-based polyurethanes. Eur. Polym. J., 67, 513–523,
2015.
50. Pizzi, A., Tannin-based polyurethane adhesives. J. Appl. Polym. Sci., 23,
1889–1990, 1979.
The Chemistry of Condensed Tannins 263

51. Pizzi, A., Tannin formaldehyde exterior wood adhesives through flavonoid
B-ring cross-linking. J. Appl. Polym. Sci., 22, 2397–2399, 1978.
52. Pizzi, A. and Cameron, F.A., Flavonoid tannins—Structural wood compo-
nents for draught resistance mechanism of plants. Wood Sci. Technol., 20,
119–124, 1986.
53. Hillis, W.E. and Urbach, G., The reaction of (+)-catechin with formaldehyde.
J. Chem. Technol. Biotechnol., 9, 9, 474–482, 1959.
54. Hillis, W.E. and Urbach, G., Reaction of polyphenols with formaldehyde.
J. Chem. Technol. Biotechnol., 9, 12, 665–673, 1959.
55. Pizzi, A. and Stephanou, A., A 13C NMR study of polyflavonoid tannin adhe-
sives intermediates, Part 2: Colloidal state reactions. J. Appl. Polym. Sci., 51,
2125–2130, 1994.
56. Pizzi, A. and Stephanou, A., Theory and practice of non-fortified tannin
adhesives for particleboard. Holzforsch. Holzverwert., 44, 4, 62–68, 1992.
57. Pizzi, A. and Stephanou, A., On the chemistry, behaviour and cure accelera-
tion of phenol–formaldehyde resins under very alkaline conditions. J. Appl.
Polym. Sci., 49, 2157–2160, 1993.
58. Braghiroli, F., Fierro, V., Pizzi, A., Rode, K., Radke, W., Delmmotte, L.,
Parmentier, J., Celzard, A., Condensation reaction of flavonoid tannins with
ammonia. Ind. Crops Prod., 44, 330–335, 2013.
59. Hashida, K., Makino, R., Ohara, S., Amination of pyrogallol nucleus of
condensed tannins and related polyphenols by ammonia water treatment.
Holzforschung, 63, 319–326, 2009.
60. Thebault, M., Pizzi, A., Santiago-Medina, F.J., Al-Marzouki, F.M., Abdalla, S.,
Isocyanate-free polyurethanes by coreaction of condensed tannins with ami-
nated tannins. J. Renew. Mat., 5, 21–29, 2017.
61. Santiago-Medina, F., Pizzi, A., Basso, M.C., Delmotte, L., Celzard, A.,
Polycondensation resins by flavonoid tannins reaction with amines. Polymers,
9, 2, 37, 1–16, 2017.
62. Basso, M.C., Pizzi, A., Polesel-Maris, J., Delmotte, L., Colin, B., Rogaume, Y.,
MALDI-TOF and 13C NMR analysis of the cross-linking reaction of con-
densed tannins by triethyl phosphate. Ind. Crops Prod., 95, 621–631, 2017.
63. Basso, M.C., Lacoste, C., Pizzi, A., Fredon, E., Delmotte, L., Flexible tannin–
furanic films and lacquers. Ind. Crops Prod., 61, 352–360, 2014.
64. J. Polesel-Maris and I. Jutang, Antiadhesives coatings based on condensed
tannins, patent WO2017/037393 A1, assigned to SEB Development, 2017.
65. Pizzi, A. and Stephanou, A., A comparative 13C NMR study of polyflavonoid
tannin extracts for phenolic polycondensates. J. Appl. Polym. Sci., 50, 2105–
2113, 1993b.
66. Pasch, H., Pizzi, A., Rode, K., MALDI-TOF mass spectrometry of polyflavo-
noid tannins. Polymer, 42, 18, 7531–7539, 2001.
67. Radebe, N., Rode, K., Pizzi, A., Pasch, H., Microstructure elucidation of poly-
flavonoid tannins by MALDI-TOF-CID. J. Appl. Polym. Sci., 127, 1937–1950,
2013.
264 Adhesives for Wood and Lignocellulosic Materials

68. Konai, N., Raidandi, D., Pizzi, A., Meva’a, L., Characterisation of Ficus syco-
morus using ATR-FTMIR, MALDI-TOF MS and 13C NMR methods. Eur. J.
Wood Prod., 75, 807–815, 2017.
69. Ricci, A., Parpinello, G.P., Palma, A.S., Teslic, N., Brilli, C., Pizzi, A.,
Versari,  A., Analytical profiling of food-grade extracts from grape (Vitis
vinifera sp) seeds and skins, green tea (Camellia sinensis) leaves and
Limousin oak (Quercus robur) heartwood using MALDI-TOF-MS, ICP-MS
and spectrophotometric methods. J. Food Comp. Anal., 59, 95–104, 2017.
70. Ricci, A., Parpinello, G., Schwertner, A.P., Teslic, N., Brilli, C., Pizzi, A.,
Versari, A., Quality assessment of food grade plant extracts using MALDI-
TOF-MS, ICP-MS and spectrophotometric methods. J. Food Comp. Anal.,
59, 95–104, 2017.
71. Jahanshahi, S., Pizzi, A., Abdolkhani, A., Doosthoseini, K., Shakeri, A., Lagel,
M.C., Delmotte, L., MALDI-TOF and 13C-NMR and FT-MIR and strength
characterization of glycidyl ether tannin epoxy resins. Ind. Crops Prod., 83,
177–185, 2016.
72. Santiago-Medina, F.J., Foyer, G., Pizzi, A., Caillol, S., Delmotte, L., Lignin-
derived non-toxic aldehydes for ecofriendly tannin adhesives for wood pan-
els. Int. J. Adhes. Adhes., 70, 239–248, 2016.
73. Pizzi, A. and Roux, D.G., The chemistry and development of tannin-based
weather- and boil-proof cold-setting and fast-setting adhesives for wood.
J. Appl. Polym. Sci., 22, 1945–1954, 1978.
74. Pizzi, A., Rossouw, D.dU T., Knuffel, W., Singmin, M., “Honeymoon” pheno-
lic and tannin-based fast setting adhesive systems for exterior grade finger-
joints. Holzforsch. Holzverwert., 32, 6, 140–150, 1980.
75. Pizzi, A. and Cameron, F.A., Fast-set adhesives for glulam. For. Prod. J., 34, 9,
61, 1984.
76. Meikleham, N. and Pizzi, A., Acid and alkali-setting tannin-based rigid
foams. J. Appl. Polym. Sci., 53, 1547–1556, 1994.
77. Tondi, G., Pizzi, A., Olives, R., Natural tannin-based rigid foams as insula-
tion in wood construction. Maderas, 10, 3, 219–227, 2008.
78. Tondi, G. and Pizzi, A., Tannin based rigid foams: Characterisation and
modification, Ind. Crops Prod., 29, 356–363, 2009.
79. Tondi, G., Zhao, W., Pizzi, A., Fierro, V., Celzard, A., Tannin-based rigid
foams: A survey of chemical and physical properties. Bioresource Technol.,
100, 5162–5169, 2009.
80. Basso, M.C., Li, X., Giovando, S., Fierro, V., Pizzi, A., Celzard, A., Green,
formaldehyde-free, foams for thermal insulation. Adv. Mater. Lett., 2, 6, 378–
382, 2011.
81. Basso, M.C., Giovando, S., Pizzi, A., Celzard, A., Fierro, V., Tannin/furanic
foams without blowing agents and formaldehyde. Ind. Crops Prod., 49, 17–22,
2013.
82. Basso, M.C., Pizzi, A., Celzard, A., Influence of formulation on the dynamics
of preparation of tannin based foams. Ind. Crops Prod., 51, 396–400, 2013.
The Chemistry of Condensed Tannins 265

83. Basso, M.C., Pizzi, A., Celzard, A., Dynamic monitoring of tannin foams
preparation: Surfactant effects. Bioresources, 8, 4, 5807–5816, 2013.
84. Lacoste, C., Pizzi, A., Basso, M.C., Laborie, M.-P., Celzard, A., Pinus pinaster
tannin/furanic foams: Part 1: Formulation. Ind. Crops Prod., 52, 450–456,
2014.
85. Lacoste, C., Pizzi, A., Basso, M.C., Laborie, M.-P., Celzard, A., Pinus pinas-
ter tannin/furanic foams: Part 2: Physical properties. Ind. Crops Prod., 61,
531–536, 2014.
86. Basso, M.C., Pizzi, A., Lacoste, C., Delmotte, L., Al-Marzouki, F.A., Abdalle,
S., Celzard, A., Tannin–furanic–polyurethane foams for industrial continu-
ous plant lines. Polymers, 6, 2985–3004, 2014.
87. Basso, M.C., Giovando, S., Pizzi, A., Pasch, H., Pretorius, N., Delmotte, L.,
Flexible-elastic copolymerized polyurethane–tannin foams. J. Appl. Polym.
Sci., 131, 13, 2014.
88. Lacoste, C., Basso, M.C., Pizzi, A., Celzard, A., Ella Bang, E., Gallo, N.,
Charrier, B., Pine (P. pinaster) and quebracho (Schinopsis lorentzii/balansae)
tannin based foams as green acoustic absorbers. Ind. Crops Prod., 67, 70–73,
2015.
89. Li, X., Nicollin, A., Pizzi, A., Zhou, X., Sauget, A., Delmotte, L., Natural
tannin/furanic thermosetting moulding plastics. RSC Adv., 3, 17732–
17740, 2013.
90. Lagel, M.C., Zhang, J., Pizzi, A., Cutting and grinding wheels for angle grind-
ers with a bioresin matrix. Ind. Crops Prod., 67, 264–269, 2015.
91. Lagel, M.C., Hai, L., Pizzi, A., Basso, M.C., Delmotte, L., Abdalla, S., Zahed,
A., Al-Marzouki, F.M., Automotive brake pads made with a bioresin matrix.
Ind. Crops Prod., 85, 3, 372–381, 2016.
92. Pizzi, A., Tannin-based overlays for particleboard. Holzforsch. Holzverwert.,
31, 3, 59–61, 1979.
93. Abdullah, U.H., Pizzi, A., Zhou, X., Rode, K., Delmotte, L., Mansouri, H.R.,
Mimosa tannin resins for impregnated paper overlays. Eur. J. Wood Prod., 71,
153–162, 2013.
94. Abdullah, U.H., Zhou, X., Pizzi, A., Merlin, A., A note on the surface quality
of plywood overlaid with mimosa (Acacia mearnsii) tannin and melamine
urea formaldehyde impregnated paper: Effects of moisture content of
resin-impregnated papers before pressing on the physical properties of
overlaid panels. Int. Wood Prod. J., 4, 4, 253–256, 2013.
95. Abdullah, U.H., Pizzi, A., Zhou, X., High pressure paper laminates from
mimosa tannin resin. Int. Wood Prod. J., 5, 4, 224–227, 2014.
96. Pizzi, A., Kueny, R., Lecoanet, F., Masseteau, B., Carpentier, D., Krebs, A.,
Loiseau, F., Molina, S., Ragoubi, M., High resin content natural matrix-
natural fibre biocomposites. Ind. Crops Prod., 30, 235–240, 2009.
97. Sauget, A., Nicollin, A., Pizzi, A., Fabrication and mechanical analysis of
mimosa tannin and commercial flax fibers biocomposites. J. Adhes. Sci.
Technol., 27, 2204–2218, 2013.
266 Adhesives for Wood and Lignocellulosic Materials

98. Tondi, G., Wieland, S., Lemenager, N., Petutschnigg, A., Pizzi, A., Thevenon,
M.-F., Efficacy of tannin in fixing boron in wood: Fungal and termite resis-
tance. Bioresources, 7, 1, 1238–1252, 2012.
99. Tondi, G., Wieland, S., Wimmer, T., Thevenon, M.F., Pizzi, A., Petutschnigg,
A., Tannin-boron preservatives for wood buildings: Mechanical and fire
properties. Eur. J. Wood Prod., 70, 689–696, 2012.
100. Thebault, M., Pizzi, A., Dumarcay, S., Gerardin, P., Delmotte, L., Fredon, E.,
Polyurethanes from hydrolysable tannins obtained without using isocya-
nates. Ind. Crops Prod., 59, 329–336, 2014.
101. Thebault, M., Pizzi, A., Santiago-Medina, F.J., Al-Marzouki, F.M., Abdalla,
S., Isocyanate-free polyurethanes by coreaction of condensed tannins with
aminated tannins. J. Renew. Mat., 5, 1, 21–29, 2017.
102. Pizzi, A., Tannin: Major sources, properties and applications, Chapter 8,
in: Monomers, Polymers and Composites from Renewable Resources, M.N.
Belgacem and A. Gandini (Eds.), pp. 179–199, Elsevier, Amsterdam, 2008.
103. Krifa, M., El-Mekdad, H., Bentouati, N., Pizzi, A., Sick, E., Chekir-Ghedira,
L., Ronde, P., In vitro and in vivo anti-melanoma effects of Pituranthos tor-
tuosus essential oil via inhibition of FAK and Src activities. J. Cell. Biochem.,
117, 1167–175, 2016.
104. Krifa, M., El-Mekdad, H., Bentouati, N., Pizzi, A., Ghedira, K., Hammami,
M., El-Meshri, S.E., Chekir-Ghedira, L., Immunomodulatory and antican-
cer effects of Pituranthos tortuosus essential oil. Tumor Biol., 36, 5165–5170,
2015.
105. Krifa, M., Mustapha, N., Ghedira, Z., Ghedira, K., Pizzi, A., Chekir-Ghedira,
L., Limoniastrum guyonianum methanol extract induces immunomodula-
tory and anti-inflammatory effects by activating cellular anti-oxidant activ-
ity. Drug Chem. Toxicol., 38, 1, 84–91, 2014.
106. Noferi, M., Masson, E., Merlin, A., Pizzi, A., Deglise, X., Antioxidant char-
acteristics of hydrolysable and polyflavonoid tannins—An ESR kinetic study.
J. Appl. Polym. Sci., 63, 475–482, 1997.
12
Thermosetting Adhesives Based on
Bio-Resources for Lignocellulosic
Composites

12.1 Introduction
Wood adhesives from renewable raw materials have now been a topic of
considerable interest for many years. This interest, already present since the
1940s, became more intense with the world’s first oil crisis in the early 1970s
and subsided again as the cost of oil decreased. At the beginning of the 21st
century, this interest is becoming intense again for a number of reasons.
The foreseen future scarcity of petrochemicals still appears to be reasonably
far into the future. It is a contributing factor but, at this stage, it is not the
main motivating force. The main impulse of today’s renewed interest in bio-
based adhesives is the acute sensitivity of the general public to anything that
has to do with the environment and its protection. It is not even this concern
per se that motivates such an interest. There are rather very strict, for some
synthetic adhesives almost crippling, government regulations that are just
starting to be put into place to allay the environmental concerns of the pub-
lic. All these factors play together and reinforce each other in contributing to
the increasing interest in bio-based adhesives for lignocellulosic materials.
First of all, it is necessary to define what is meant by bio-based wood
adhesives, or adhesives from renewable, natural, non-oil-derived raw mate-
rials. This is necessary because in its broadest meaning, the term might be
considered to include urea–formaldehyde (UF) resins, urea being a non-
oil-derived raw material. This of course is not the case. The term “bio-based
adhesive” has come to be used in a very well-specified and narrow sense
to only include those materials of natural, non-mineral, origin which can
be used as such or after small modifications to reproduce the behavior and
performance of synthetic resins. Thus, only a limited number of materials
can be currently included, at a stretch, in the narrowest sense of this defi-
nition. These are tannins, lignin, carbohydrates, unsaturated oils, proteins,

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (267–291) © 2019
Scrivener Publishing LLC

267
268 Adhesives for Wood and Lignocellulosic Materials

dissolved wood, and wood welding by self-adhesion. The bio-based wood


adhesives approach does not mean, however, to go back to the technology
of natural product adhesives as they existed up to the 1920s and 1930s
before they were supplanted by synthetic adhesives. The bio-based adhe-
sives that we are talking about here are yes derived from natural adhesives,
but using or requiring novel technologies, formulations, and methods. All
these materials and approaches have been and will be further modified in
the future in light of present-day modern chemical knowledge.
Of the classes of bio-based wood adhesives mentioned above, in the case
of tannins and lignins, their interest has been directed primarily at sub-
stituting phenol–formaldehyde (PF) resins, due to the phenolic nature of
these two classes of compounds. In some but not all of these cases, some
formaldehyde is still used, and in the case of lignin, some other additives.
It is then necessary to distinguish between bio-based adhesives in which
a limited amount of synthetic additives are still used, and bio-based wood
adhesives where no synthetic additives are used.
This chapter deals with reviewing the newer technologies that can be
implemented in bio-adhesives, and not the already used, industrial tech-
nologies, that, as good as they can be, are already described in-depth in
other reviews [1, 2].

12.2 Tannin Adhesives


Two different types of natural phenolics can be defined as tannins: hydro-
lyzable tannins and condensed tannins. Hydrolyzable tannins are com-
posed of simple phenols, for example, pyrogallol and/or ellagic acid, plus
monocarbohydrates, in general glucose esterified by gallic and digallic
acids [3]. The low level of phenol substitution they allow, coupled with
low nucleophilicity and their absence of macromolecular structure when
extracted, their lower world production, and higher price have some-
what decreased their chemical and economical interest. Some references
to their utilization, even industrial utilization, do however exist and are
noteworthy.
In 1973, as a consequence of the first oil crisis, Norsechem, a Malaysian
adhesives factory belonging to a Norwegian paint group, producing PF
resins for Southeast Asia plywood manufacturers, was forced to imple-
ment a technological change that was maintained for 3 years in indus-
trial production [4]. This technology consisted in substituting 33% by
weight of phenol with chestnut tannin extract, a hydrolyzable tannin,
in the PF resin during its preparation. Up to 50% substitution in the
Bio-Based Adhesives 269

laboratory was claimed but no glue mix or other information on this was
given [4]. The motive of this was purely economical, as phenol prices
had skyrocketed and the cost of chestnut tannin extract was at that time
much lower. In the only written reference that exists to this development
[4], although results and basic material proportions are reported, resin
formulations are not disclosed. A more recent research work has taken
up again this approach using more modern PF formulations. Phenol–
formaldehyde–chestnut tannin adhesives satisfy the relevant standards
in which phenol:tannin weight ratios 30:70 were obtained and success-
fully used [5].
Conversely, condensed flavonoid tannins constitute in excess of 90% of
the worldwide production of commercial tannins, with production now
being around 220,000 tons/year. These tannins have a greater interest both
chemically and economically to prepare resins and adhesives. They are
widely distributed in nature and are particularly concentrated in the wood
and bark of a variety of trees. These include Acacia (wattle or mimosa bark
extract) species, Schinopsis species (quebracho wood extract), Tsuga (hem-
lock bark extract) species, and Rhus (sumach extract) species. Commercial
tannins are extracted from these species and from the bark of a variety of
Pinus species.

12.2.1 New Technologies for Industrial Tannin Adhesives


Extensive, up-to-date, and in-depth reviews of the technology of tannin
adhesives based on the classical technology of tannin–formaldehyde resins
already exist [3] and are beyond the scope of this review, as these tech-
nologies are commercial, now for several years, and used in a number of
countries. It is sufficient to state here that tannin–formaldehyde adhesives
with very low emission (E0), with fast pressing times, and using unmodi-
fied tannin extracts are well known and are used commercially, and their
technology is commercially and perfectly mastered [3, 6]. The technologies
of interest here are the new ones based either on no addition of aldehydes
or on the use of hardeners that are non-emitting or manifestly nontoxic.
Some research has been promoted to improve formaldehyde emission
due to the pressing need to eliminate or at least decrease formaldehyde
emissions from adhesive-bonded wood panels, although this is possi-
bly unnecessary for tannin adhesives due to their already low emission
(as most phenolic adhesives). Two research approaches have been taken:
(i) using no-emitting hardeners due to the total absence of aldehydes in the
tannin adhesive or because the aldehyde is blocked to be emitted from the
adhesive, and (ii) tannin self-condensation. Methylolated nitroparaffins
270 Adhesives for Wood and Lignocellulosic Materials

such as trishydroxymethyl nitromethane [3, 6], are part of the first class.
They function well and they have considerable advantages for the adhesive
and for the bonded wood panels. In panel products such as particleboard,
medium-density fiberboard (MDF), and plywood, the wood panels’ exte-
rior/marine-grade performances are obtained, coupled with a desirable
and rather marked lengthening in glue-mix pot life. Moreover, this hard-
ener causes a marked reduction in formaldehyde emission from the wood
panels, reducing emission to just the formaldehyde emitted by heating the
wood (and even slightly less). Moreover, trishydroxymethyl nitromethane
is miscible in any proportion with formaldehyde-type hardeners for tan-
nin adhesives, their proportional substitution with it inducing a marked
decrease in formaldehyde emission from the wood panel. All this occurs
without affecting the exterior/marine-grade performance of the wood
panel. Industrial plant trials for MDFs confirm all these properties, the
trial results and conditions having been reported [7–9]. Equally, formula-
tions in which tannins are hardened by furfuryl alcohol just as a glue-mix
hardener or even are prereacted in a reactor with furfuryl alcohol have
been used and extensive tempered hardboard plant trials have been car-
ried out [10, 11]. A cheaper but an equally effective alternative to these
approaches is the use of acetone–formaldehyde resins as hardeners or even
better is the use of hexamine as a tannin hardener.
Recently, a considerable advance in preparing totally biosourced, non-
toxic, environment-friendly tannin adhesives was taken in which a natural
nontoxic aldehyde, vanillin, derived from the pulping of wood, was used as
a hardener of a pine tannin adhesive [12].

12.2.2 Tannin–Hexamethylenetetramine (Hexamine)


Adhesives
Under many wood adhesive application conditions, contrary to what
was thought for many years in the past, hexamine used as a hardener of
a fast-reacting species is not at all a formaldehyde-yielding compound,
yielding extremely low formaldehyde emissions in bonded wood joints
[13]. 13C NMR (nuclear magnetic resonance) evidence has confirmed
[14–16] that the main decomposition (and recomposition) mechanism of
hexamine under such conditions is not directly to formaldehyde. It rather
proceeds through reactive intermediates, hence mainly through the forma-
tion of reactive imines and iminoamino methylene bases (Figure 12.1). 13C
NMR evidence has also confirmed [14–16] that in the presence of chem-
ical species with very reactive nucleophilic sites, such as melamine, resor-
cinol, and condensed flavonoid tannins, hexamine does not decompose
Bio-Based Adhesives 271

to formaldehyde and ammonia. Instead, the very reactive but unstable


intermediate fragments react with the tannin, melamine, etc. to form ami-
nomethylene bridges before any chances to yield formaldehyde. These are
also stable up to 5 h at temperatures as high as 120°C. The intermediate
fragments of the decomposition of hexamine pass first through the forma-
tion of imines followed by their decomposition to imino-methylene bases.
The latter present only one positive charge as the second methylene group
is stabilized by an imine-type bond [14–16] (Figure 12.1). Any species with
a strong real or nominal negative charge under alkaline conditions, be it
a tannin, resorcinol, or another highly reactive phenol, be it melamine or
another highly reactive amine or amide, or be it an organic or inorganic

Φ O OH HO O Φ

-δ -δ
HO H+2N=CH-NR-CH2-NR-CH=N+H2 OH

Ambient temperature and


polymeric ionic
complexes

Φ O O- H+2N=CH-NR-CH2-NR-CH=N+H2 -O O Φ

HO OH

N
CH2 CH2 CH2
N HOCH2OH + HN=CH-NR-CH2-NR-CH=NH HN=CH-N-CH2OH
I
CH2 CH2 CH2OH
N N HOCH2OH
CH2
HN=CH-NR-CH2O-CH2-NR-CH=NH HN=CH-N-CH2+
I
CH2+

HN-CH2+
I and H2C=N-CH2+
CH2+


hardened
Tannin-CH2-NH-CH2-Tannin

Figure 12.1 Schematic representation of the decomposition of hexamine to iminoamino


methylene basis in the presence of a reactive species such as tannin to form (i) ionic polymeric
complexes at ambient temperature and (ii) stable benzylamine covalently bridged network in
hardening, at higher temperature, without producing or releasing any formaldehyde.
272 Adhesives for Wood and Lignocellulosic Materials

anion, is capable of reacting with the intermediate species formed by decom-


position (or recomposition) of hexamine far more readily than formalde-
hyde [14–16]. This explains the capability of wood adhesive formulations
based on hexamine to give bonded panels of extremely low formaldehyde
emissions. If no highly reactive species with strong real or nominal nega-
tive charge is present, then decomposition of hexamine proceeds rapidly to
formaldehyde formation as reported in the previous literature [17].
On this basis, the use of hexamine as a hardener of a tannin, hence a
tannin–hexamine adhesive, is a very environmental-friendly proposition.
Formaldehyde emissions in a great chamber have been proven to be so low
to be limited exclusively to what is generated by the wood itself, hence truly
E0 panels. The panels obtained with tannin–hexamine adhesives, according
to which conditions they are manufactured, can satisfy both interior and
exterior grade standard specification requirements [18]. Steam injection
presses recently have been shown to be better suited to give better results
for exterior grade boards using tannin–hexamine adhesives [19]. Typical
laboratory particle board and full industrial scale plant trials yielded panel
strength results as indicated in Table 12.1 [20].
Comparable results as those reported in Table 12.1 are obtained with
pine tannins or other procyanidins hardened with hexamine [20]. In the
same reference, catalysis of the reaction in the presence of small amounts of
accelerators such as a zinc salt allow even better results or faster press times.

12.2.3 Hardening by Tannin Autocondensation


The autocondensation reactions characteristic of polyflavonoid tannins
have only recently been used to prepare adhesive polycondensates hard-
ening in the absence of aldehydes [21]. This autocondensation reaction is
based on the opening under either alkaline or acidic conditions [19] of the
O1–C2 bond of the flavonoid repeating unit and the subsequent condensa-
tion of the reactive center formed at C2 with the free C6 or C8 sites of a fla-
vonoid unit on another tannin chain [21–26]. Although this reaction may

Table 12.1 Typical results of laboratory and industrial particleboard


bonded with mimosa tannin extract hardened with 6.5% hexamine.
IB 2 h boil tested,
IB dry (MPa) redried (MPa) Density (kg/m3)
Laboratory 0.92 0.27 711
Industrial 0.58 – 680
Bio-Based Adhesives 273

lead to considerable increases in viscosity, gelling does not generally occur.


However, gelling takes place when the reaction occurs (i) in the presence of
small amounts of dissolved silica (silicic acid or silicates) catalyst and some
other catalysts [21–26], and (ii) on a lignocellulosic surface [25].
As in the case of other formaldehyde-based resins, the interaction
energies of tannins with cellulose obtained by molecular mechanics cal-
culations [25] tend to confirm the effect of surface catalysis induced by
cellulose also on the curing and hardening reaction of tannin adhesives.
The considerable energies of interactions obtained can effectively explain
the weakening of the heterocyclic ether bond, leading to accelerated and
easier opening of the pyran ring in a flavonoid unit, as well as the facility
with which hardening by autocondensation can occur. In the case of the
more reactive procyanidins and prodelphinidin-type tannins, such as pine
tannin, cellulose catalysis is more than enough to cause hardening and to
produce boards of strength satisfying the relevant standards for interior
grade panels [26]. The internal bond (IB) strengths of an interior particle-
board prepared using different types of commercial tannins are shown in
Figure 12.2.
Figure 12.3 shows that the slower-reacting tannins can yield an upgraded
IB strength of the board when mixed with small amounts of faster-reacting
tannins. In Figure 12.3, the effect of adding pecan tannin is shown as an
example, but similar upgrades can be obtained by adding pine tannin too.
In the case of the less reactive tannins, however, such as mimosa and que-
bracho, the presence of a dissolved silica or silicate catalyst of some type is

1.0

Pecan
0.8
Pine

0.6
IB (MPa)

0.4 DIN 68763

Mimosa
0.2 Quebracho

0
5 6 7 8 9 10 11 12 13 14
pH

Figure 12.2 Dry IB strength as a function of tannin solution pH of laboratory


particleboards prepared with different commercial tannin extracts without any aldehyde
hardener, using the lignocellulosic substrate-induced tannin autocondensation reaction [26].
274 Adhesives for Wood and Lignocellulosic Materials

0.8

Pine
0.6
IB (MPa) Mimosa Quebracho

0.4 DIN 68763

0.2

0
0 20 40 60 80 100
Pecan Tannin (%)

Figure 12.3 Dry IB strength of laboratory particleboards prepared with mixtures of


different commercial tannin extracts, without any aldehyde hardener, as a function
of percentage of the more reactive pecan nut tannin extract, using the lignocellulosic
substrate-induced tannin autocondensation reaction. Similar results can be obtained for
mimosa and quebracho by just using <3% dissolved silica [26].

the best manner to achieve panel strength as required by the relevant stan-
dards [26]. Autocondensation reactions have been shown to contribute
considerably to the dry strength of wood panels bonded with tannins, but
to be relatively inconsequential in contributing to the bonded panels’ exte-
rior grade properties, which are rather determined by polycondensation
reactions with aldehydes [23, 27, 28]. Combination of tannin autoconden-
sation and reactions with aldehydes and combination of radicals with ionic
reactions have been used to decrease the proportion of aldehyde hardener
and to decrease considerably the already low formaldehyde emissions
yielded by the use of exterior tannin adhesives [23, 27, 28]. A variation on
the same theme of wood adhesives by tannin autocondensation is acid-
catalyzed oxidative condensation, copied altogether from the Lewis
acid-induced tannin autocondensation of the work of other groups [21,
24, 25].
These technologies, tannin autocondensation, and hexamine hardening
are already in industrial use and are under consideration by a further num-
ber of industrial groups worldwide.
Tannin–hexamine resins have been used, at a different level, for fibrous
high-strength composites in which a non-woven natural fiber mat, fibers
such as linen, kenaf, hemp, etc., is impregnated with up to 50–60% of total
weight with tannin–hexamine resin [29]. Both higher-density thin com-
posites and lower-density thicker composites were prepared. Two natural
Bio-Based Adhesives 275

matrix types were used: (i) commercial mimosa flavonoid tannin extract
with 5% hexamine added as hardener, and (ii) a mix of mimosa tannin +
hexamine with glyoxalated organosolv lignin of low molecular weight;
these two resins mixed 50/50 by solids content weight. The results obtained
were good [29]. The composites made with the mix of tannin and lignin
resins as a matrix remained thermoplastic after a first pressing. The flat
sheets prepared after the first pressing were then thermoformed into the
shape wanted.

12.3 Lignin Adhesives


Much has been written about, and much research has been conducted in
the use of lignins for wood panel adhesives. It can safely be said that this
natural raw material has probably been the most intensely researched one
as regards wood adhesives application. Lignins are phenolic materials, they
are abundant and of low cost, but they have lower reactivity towards form-
aldehyde, or other aldehydes, than even phenol. Extensive reviews on a
number of proposed technologies of formulation and application do exist,
and the reader is referred to these in earnest [1, 30–40]. This field is, how-
ever, remarkable for how small has been the industrial success in using
these materials. In general, lignin and lignosulphonates have been mixed
in smaller proportions to synthetic resins, such as PF resins [33–36], and
even UF resins [37], to decrease their cost. Their low reactivity and lower
level of reactive sites, however, conjure that for any percentage of lignin
added, the cost advantage is abundantly lost in the lengthening of the panel
pressing time this causes. The only step forward that has found industrial
application in the last 20 years is to prereact in a reactor lignin with form-
aldehyde to form methylolated lignin, thus to do part of the reaction with
formaldehyde first, and then add this methylolated lignin to PF resins at
the 20% to 30% level [33, 35]. These resins have been used in some North
American plywood mills [33]. Particularly in plywood mills, the pressing
time is not the factor determining the output rate of the factory and so one
can afford to use relatively long press times with good results [33].
None of the many adhesive systems based on pure lignin resins, hence
without synthetic resin addition, have succeeded commercially at an
industrial level, at least not for long periods. Some were tried industrially,
but for one reason or another, too long a pressing time, high corrosiveness
for the equipment, etc., they did not meet with commercial success. Still
notable among these is the Nimz system based on the networking of lignin
in the presence of hydrogen peroxide [1, 2, 39, 40]. Only one system is used
276 Adhesives for Wood and Lignocellulosic Materials

successfully to date, but this is only for high-density hardboard, in several


mills worldwide. This is the Shen system, based on the self-coagulation and
cross-linking of lignin by a strong mineral acid in the presence of some
aluminum salt catalysts [41, 42]. However, attempts to extend this system
to the industrial manufacture of MDF are known to have failed.
Of interest in the MDF field is also the system of adding laccase
enzyme-activated lignin to the fibers or activating the lignin in situ, in the
fibers also by enzyme treatment [43, 44]. The results obtained, however,
yielded boards that did not satisfy the relevant standards, and this at very
long board pressing times. The researchers involved obviated this success-
fully by adding some 1% isocyanate [polymeric 4 4 -diphenyl methane
diisocyanate (PMDI)] to the board [43] and pressing at acceptably short
press times, or by extending the pressing times to ridiculous lengths
(100 s/mm board thickness while industrial press times are of the order
of 3 to 7 s/mm board thickness) [45]. In the former case, an adhesive had
to be used, with the same result obtained by pressing untreated hardboard,
a 100-year-old process, hence just wasting expensive time and enzymes.
The second case instead illustrates even more clearly where the problem
lies and what breakthrough is necessary: enzyme mobilizing lignin works,
but not fast enough. The breakthrough necessary is a new, strong catalyst
of the enzymatic action capable of allowing pressing times of industrial
significance. This has not been found, or even considered, as yet.
A promising new technology based on lignin use for wood adhesives is
relatively recent and uses again pre-methylolated lignin in the presence of
small amounts of a synthetic PF resin and PMDI [46, 47]. The proportion of
pre-methylolated lignin used is 65% of the total adhesive, the balance being
made up of 10–15% PF resin and of 20–25% PMDI. This adhesive presses
at very fast pressing times, well within the fastest range used today industri-
ally, contains a high proportion of lignin, and yields exterior grade boards
[46, 47]. PF resin and methylolated lignin cure accelerators such as triacetin,
other esters, or others, can also be used [46, 47] notwithstanding that just the
presence of the PMDI already gives a considerable acceleration to the curing
rate (Table 12.2). The system is based on cross-linking caused by the simulta-
neous formation of methylene bridges and of urethane bridges, overcoming
with the latter the need for higher cross-linking density that has been one of
the problems that has stopped lignin utilization in the past.
Recently, an adhesive for wood panels based on a 50/50 mix of tannin/
hexamine with pre-glyoxalated lignin, glyoxal being a nontoxic and non-
volatile albeit less reactive aldehyde, has been developed, yielding interest-
ing results that are 100% bio-based [48]. Oxazolidines have also been used
as lignin adhesive hardeners of low formaldehyde emission [49].
Bio-Based Adhesives 277

Table 12.2 Results of laboratory particleboard bonded with methylolated


kraft lignin-based adhesives co-reacted at the glue-mix level with PMDI and a
synthetic PF resin. The proportions of the three materials are by weight.
Press time IB dry IB 2 h boil tested, Density
(s/mm) (MPa) redried (MPa) (kg/m3)
MDI/PF/Kraft Lignin 10 0.85 0.53 687
(methylolated)
22/26/52 by weight
+ ester accelerator
MDI/PF/Kraft Lignin 20 0.68 0.53 696
(methylolated)
11/26/63 by weight
+ ester accelerator
MDI/PF/Kraft Lignin 10 0.53 0.43 680
(methylolated)
11/26/63 by weight
+ ester accelerator

Other than the above, there is not much new in this field, but just liter-
ature rehashing older systems all based on the substitution of some phenol
in PF resins. In general, these papers do not seem to be aware of the slow
pressing time problem, and they do not address it, perpetuating the myth
of PF/lignin adhesives while repeating the same age-old errors. They lead
new researchers in the field to believe they are doing something worth-
while with parameters that do not satisfy the requirements of press rate of
the panel manufacturing industry.
All the above does not mean that there may not be new opportunities in
lignin-based wood adhesives. It simply means that further new, alternative
technologies need to be developed before lignin can be used commercially
and extensively as a wood adhesive. To the knowledge of the writer, such a
quantum leap has not occurred as yet in this field.

12.4 Protein Adhesives


Intense research induced by the sponsorship of the US United Soybean
Board has revised interest in protein adhesives, soya protein primarily,
but also others. These technologies must not be confused with the age-old
protein and bone glues used in carpentry, or blood used as an additive in
278 Adhesives for Wood and Lignocellulosic Materials

plywood glue mixes. Certainly, of the traditional technologies, one stands


out head and shoulders above the others, namely, casein adhesives. These
are still produced and still used industrially to very good effect in some
special plywood and related products. They work as such, they are very
environmentally friendly, and their technology was completely mastered a
long time ago. They are definitely strong candidates for future expansion,
in their actual form or with some further technological improvement.
Soy protein hydrolysate- and soy flour-based adhesives instead are defi-
nitely new. Both addition to traditional synthetic wood adhesives and their
use as panel adhesives after partial hydrolysis and modifications have been
reported, and with acceptable results, it appears [50–53]. These products
are used industrially/commercially but by only one company in the United
States, although the pressing times appear to be long in relation to what is
envisaged with synthetic adhesives in the panel industry today. Aldehyde-
prereacted soy resins [54] or other protein resins, such as wheat gluten for
example [55–57], or protein resins mixed with synthetic adhesives have
been reported with good results [54, 55, 58], especially where the reinforc-
ing synthetic resin is 20–30% MDI and some use of them can be envisaged
in the future too.

12.5 Carbohydrate Adhesives


Carbohydrates in the form of polysaccharides, gums, oligomers, and mono-
meric sugars have been employed in adhesive formulations for many years.
Carbohydrates can be used as wood panel adhesives in three main ways: (i) as
modifiers of existing PF and UF adhesives, (ii) by forming degradation com-
pounds that then can be used as adhesive building blocks, and (iii) directly
as wood adhesives. The second route above leads to furanic resins. Furanic
resins, notwithstanding that their basic building blocks, furfuraldehyde and
furfuryl alcohol, are derived from the acid treatment of the carbohydrates in
waste vegetable material, are considered today as purely synthetic resins [59].
This opinion might need to change. However, both compounds are relatively
expensive and very dark colored, and furanic resins have made their indus-
trial mark in fields where their high cost is not a disadvantage. They can be
used very successfully for panel adhesives, they are used very successfully
in other fields (as foundry core binders), but the relatively higher toxicity of
furfuryl alcohol before it is reacted is a problem that will have to be taken
into consideration if these resins are to be considered for wood products.
The use of carbohydrates directly dissolved in strong alkali as wood
panel adhesives is not a new concept, but is an interesting and a topical
Bio-Based Adhesives 279

one. This technology has been extensively reported [60]. All sorts of agri-
cultural cellulosic materials have been successfully adapted to this technol-
ogy, and the technology and its application have been extensively reported
in the past [60].
Research on the first route has centered particularly on the substitution of
carbohydrates for parts of PF resins. It has been reported that, at laboratory
level, up to 50% to 55% of phenol in a PF resin can be substituted with a vari-
ety of carbohydrates, from glucose to polymeric, tree-derived hemicellulo-
ses [7, 11, 42, 61, 62]. Apparently, reducing sugars could not be used directly
as they are degraded to saccharinic acids under the acid conditions required
in the formulation of the resin. Reducing sugars can be used to successfully
modify PF resins if they are reduced to the corresponding alditols or con-
verted to glycosides. Some carbohydrates appeared to be incorporated into
the resin network mainly through ether bridges [61]. Generally, the resin
is prepared by co-reacting phenol, the carbohydrate in high proportion, a
lower amount of urea, and formaldehyde. Extensive and rather successful
industrial trials of these resins have also been reported [11].
Carbohydrate-based adhesives in which the formulation starts with
the carbohydrate itself have also been reported, but the acid system used
during formulation readily degrades the original carbohydrate to furan
intermediates, which then polymerize. An interesting concept that was
advanced early on in carbohydrate adhesives research was the conversion
of the carbohydrate to furanic products in situ, which then homopolymer-
ize as well as react with the lignin in wood.
Several research groups [61, 63] have recently described the use of liq-
uefied products from cellulosic materials, literally liquefied wood, which
showed good wood adhesive properties. Lignocellulosic and cellulosic mate-
rials were liquefied in the presence of sulfuric acid under normal pressure
using either phenol or ethylene glycol. The cellulosic component in wood
was found to lose its pyranose ring structure when liquefied. The liquefied
product contains phenolic groups when phenol is used for liquefaction. In
the case of ethylene glycol liquefaction, glucosides were observed at the ini-
tial stage of liquefaction and levulinates after complete liquefaction.

12.6 Unsaturated Oil Adhesives


Saturated and unsaturated vegetable oils are now widely available as a bulk
commodity for a variety of purposes and at very acceptable prices. All resin
research to date has focused on oils that contain at least one double bond.
These oils are predominantly a mixture of triglycerides, hence esters, with
280 Adhesives for Wood and Lignocellulosic Materials

a small quantity of free fatty acid; the small proportion of free fatty acid
is dictated both by the plant species and the extraction conditions. As the
number of unsaturations increases, so does the overall molecule reactivity
and its potential for side reactions.
Until fairly recently, only two examples could be found in the literature
where seed oil derivatives were being employed as wood adhesives. Linseed
oil, for example, has been used to prepare a resin that can be used as an adhe-
sive or surface coating material [64, 65]. The chemistry of this resin centers
on an epoxidation of the oil double bonds followed by cross-linking with
a cyclic polycarboxylic acid anhydride to build up molecular weight. The
reaction is started by the addition of a small amount of polycarboxylic acid.
When the epoxidized oil resin was evaluated as a wood adhesive in
composite panels, it could be tightly controlled through the appropriate
selection of triglycerides and polycarboxylic anhydrides. This apparently
enables a wide range of materials with quite different features to be man-
ufactured. The use as wood adhesives is one of the many uses, with the
focus of the development being more on plastic materials. The literature
states that this plastic is well suited for use as a formaldehyde-free binder
for wood fibers and wood particles including fibers and chips from cereal
residues, such as straw and fiber mats.
The literature on this resin [64] claims that cross-linking can be varied
through the addition of specialized catalysts, and several samples were pre-
pared at a range of temperatures (120–180°C) that exhibited high water tol-
erance even at elevated temperatures, but no actual test data were included.
Since the resin of reference [64], research in a number of other countries by
imitators has produced very similar epoxidized oil resins. These are suit-
able for a number of applications, but the writer has tested one or two of
them, finding that for wood adhesives application, these resins have two
major defects: (i) their hot-pressing time is far too slow to be of any interest
in wood panel products, with the exception perhaps of plywood (for which
they have not been tried), and (ii) they are relatively expensive. Unless the
slow hot-pressing problem is overcome, and at a reasonable price, these
resins will remain at the stage of potential interest. There is no doubt that
these resins can be of interest in other fields, but it is symptomatic that
no industrial use for wood panel adhesives has been reported as yet, or
is known to have occurred. Thus, the use of unwoven natural fiber mats
impregnated with epoxidized unsaturated oils and hardened by using
maleic anhydride has potential for applications in which the resin forms a
high proportion of the composite [65].
Bioresins based on soy bean and other oils have been developed also
by other groups, mainly for replacement of polyester resins [66]. These
Bio-Based Adhesives 281

liquid resins were obtained from plant and animal triglycerides by suit-
ably functionalizing the triglyceride with chemical groups (e.g., epoxy,
carboxyl, hydroxyl, vinyl, amine, etc.) that render it polymerizable. The
reference claims that excellent inexpensive composites were made using
natural fibers, such as hemp, straw, flax, and wood in fiber, particle, and
flake form, and that soy-oil-based resins have a strong affinity for natural
fibers and form a good fiber–matrix interface as determined by scanning
electron microscopy of fractured composites. The reference also stated that
these resins can be viewed as candidate replacements for PF, urethane, and
other petroleum-based binders in particleboard, MDF, oriented strand
board (OSB), and other panel types. However, no actual test data have been
supplied, and no industrial use in wood panel adhesives has actually been
reported as yet.
Cashew nut shell liquid (CNSL), mainly composed of cardanol but con-
taining also other compounds (Figure 12.4), is an interesting candidate for
wood-based resins. Its dual nature, phenolic nuclei + unsaturated fatty acid
chain, makes a potential natural raw material for the synthesis of water-
resistant resins and polymers.
Cardanol resins are known from the past, but their use has not been
very diffuse simply because the raw material itself was rather expensive.
The price, however, appears to be more affordable now since the extensive
cashew nut plantations in Mozambique are in production.
The phenol, often resorcinol, group and/or double bonds in the chain
can be directly used to form hardened networks. Alternatively, more suit-
able functional groups such as aldehyde groups and others can be gen-
erated on the alkenyl chain. Generally, modifications of this kind take

OH R1 = H or CH3 Anacardic acid (R1=H, R2=H, R3=COOH)


R1 R3 Cardanol (R1=H, R2=H, R3=H)
R2 = H or OH Cardol (R1=H, R2=OH, R3=H)
R2 Methyl Cardol (R1=CH3, R2=OH, R3=H)
C15H31-n R3 = H or CO OH

C15H31-n = n=0
n=1
n=2
n=3

Structure of Cashew Nut Shell Liquid

Figure 12.4 Chemical composition and type of compounds in CNSL.


282 Adhesives for Wood and Lignocellulosic Materials

several reaction steps, rendering the process too expensive for commercial
exploitation in wood adhesives.
However, the Biocomposite Center in Wales [67] has developed a system
of ozonolysis [67, 68] in industrial methylated spirit [68] through which an
aldehyde function is generated on the alkenyl chain of cardanol. The first
reaction step yields as major product a cardanol hydroperoxide that, fol-
lowing reduction by glucose or by zinc/acetic acid, yields a high proportion
of cardanolaldehyde groups. These cross-link with the aromatic groups of
cardanol itself, and thus a self-condensation of the system, yielding hard-
ened networks (Figure 12.5).
Exploratory laboratory particle board and lap shear bonding yielded the
results shown in Table 12.3. The results reported are good. Nonetheless,
neither the press times used, nor other essential conditions that could help
to evaluate the economical feasibility of these products were reported [67].
It remains to be evaluated if the cost of the ozonolysis allows wood adhe-
sives of a suitably low cost and, again, if the pressing times can match those
of industrial resins. However, the resorcinolic structure of the cardanol
phenol group would appear to indicate that the molecule should be able

HO O3 20°C
in meth spirit

OH
OOH
CHO
Ar +
RO
an hydroperoxide

reduction
(glucose or
z n.acetic acid)

OHC

Phenol/aldehyde
crosslinked network

Figure 12.5 Ozonolysis of CNSL to produce cardanolaldehyde and a hydroperoxide, the


latter further reduced to another aldehyde. The two aldehyde groups and the reactive sites
on the aromatic ring react readily to form a cross-linked network during hardening.
Bio-Based Adhesives 283

Table 12.3 Laboratory results for ozonolysis/reduction of a CNSL bonded


particleboard. Single lap shear bonds cured at 180°C for 3 min and IB results
of laboratory particleboard (percentage resin load of 10%) (press times, board
density, and other conditions were not reported).
Lap shear bond IB dry IB 2h boil
(MPa) (MPa) (MPa)
Phenol–formaldehyde control 5.55 0.69 *
CNSL aldehyde (zinc/acetic acid 6.77 1.05 0.58
reduced)
CNSL aldehyde (glucose reduced) 6.02 * *
* Not reported.

to achieve industrial pressing times. A similar approach has been taken by


using unsaturated sunflower oil ozonolysis to generate the aldehyde groups
coupled with resorcinol and tannin [69]. The results obtained were com-
parable, indicating that more common, widely spread oils can also be used
other than cardanol [69].

12.7 Wood Welding without Adhesives


Welding techniques that are widely used in the plastic and car industries
have recently been applied also to joining wood. A variety of techniques
such as ultrasound, mechanical friction, and others can be used to weld
thermoplastic polymers.
However, the same mechanical techniques at the interface of two solid
wood surfaces in the absence of any thermoplastic material, or any other
binder, yield joints of considerable strength [70, 71]. The equipment used
for the mechanical vibration welding of wood in the absence of an adhesive
is the same type of equipment used for frictional welding of metals. Figure
12.6 shows the characteristic linear vibrational movement of the type of
industrial metal welding machine used as well as the frictional shift and
force applied to the two pieces of wood during welding.
Linear welding of wood can give bonding results satisfying the rel-
evant standards, while orbital welding gives much lower results. Some
of the parameters that influence welding of metals with the same type
of equipment also influence wood welding. Thus, the influence on the
final bond of the vibration welding time, the contact holding time after
the welding vibration had stopped, the welding pressure exerted on the
284 Adhesives for Wood and Lignocellulosic Materials

Applied force

Wood Linear shift

Figure 12.6 Example of the relative movement of two pieces of wood during wood welding
without adhesives.

surfaces, the holding pressure after the welding vibration had stopped,
and the amplitude of the shift imparted to one surface relative to the
other during vibrational welding are of importance. Welding frequen-
cies of 100 Hz to 150 Hz are used. The joint tensile strength depends
on vibration amplitude, showing some good bond strength for a 3-mm
vibrational amplitude. The joint tensile strength depends on welding
pressure, with values of 2 to 2.3 MPa giving the best results. The joint
tensile strength depends on welding time, but less markedly on welding
pressure. In general, combinations of 3 s welding time and 4–5 s holding
time give strong joints exhibiting strength in excess of 10 and sometimes
11 MPa. The relevant European Norm for these types of joints requires
strengths equal to or higher than 10 MPa.
The strong joints obtained are not capable of satisfying specifications
for exterior joints as they show very poor resistance to water. These joints
can then only be considered for interior applications such as for furni-
ture and for interior grade wood joints. Furthermore, the technique at
this stage is only usable for solid wood joints and perhaps joints between
pre-manufactured panels presenting the same type of characteristics as
solid wood, such as OSB. The technique, however, has considerable inter-
est for its low cost and in the implementation of totally environment-
friendly wood joints in joinery and furniture manufacturing. More
recently, a technique to render welded wood joints resistant to water and
exterior exposure, at low cost and still totally natural and environmen-
tally friendly, has been developed [72–75], making its application in exte-
rior structural applications possible.
The mechanism of mechanically induced wood vibration welding has
been shown to be due mostly to the melting and flowing of some amor-
phous, cell-interconnecting polymer material in the structure of wood,
mainly lignin, but also hemicelluloses. This causes partial detachment,
the “ungluing” of long wood cells and wood fibers, and the formation of a
fiber entanglement network in the matrix of molten cell-interconnecting
Table 12.4 Typical parameters in mechanically induced linear vibration wood welding and typical results of tensile strength tests
according to European Norm EN 204 on welded specimens.
Welding time Welding pressure Holding time Holding pressure Water spray Number of Tensile strength
(s) (MPa) (s) (MPa) (g/mm2) specimens tested (MPa)
3 1.3 5 1.3 No 10 9.40 ± 1.2
3 1.3 5 2 No 10 10.45 ± 0.9
3 1.3 5 2 Yes 10 10.37 ± 1.0
4 1.3 5 1.3 No 10 8.78 ± 0.8
4 1.3 5 2 Yes 10 8.47 ± 0.8
Bio-Based Adhesives
285
286 Adhesives for Wood and Lignocellulosic Materials

material that then solidifies. Thus, a wood cell/fiber entanglement network


composite having a molten lignin polymer matrix is formed.
During the welding period, some of the detached wood fibers that are
no longer held by the interconnecting material are pushed out of the joint
as excess fibers. Cross-linking chemical reactions have also been shown
and confirmed to occur (the most likely one of these identified by NMR
appears to be a cross-linking reaction of lignin with carbohydrate-derived
furfural and furfural self-polymerization). These reactions, however, are
relatively minor contributors during the very short welding period. Their
contribution increases after welding has finished, explaining why some
holding time under pressure after the end of welding contributes strongly
to obtaining a good bond. Typical results obtained are shown in Table 12.4.
Even more recently, an approach based on the high-speed rotational
welding of a dowel within a hole in a substrate has also been developed
without any adhesive being used [76–78]. A standard, fixed workshop drill
is sufficient to yield dowel welding presenting the same dry tensile strength
obtained by gluing the same dowel with PVAc. The advantages accrued
are not only the elimination of the synthetic adhesive, with its concurrent
costs and environmental benefits, but also that final strength is reached in
3 s in wood dowel welding, while it takes several hours for PVAc to bond to
the same strength level. These joints too can be made wholly exterior and
waterproof while still totally natural and environment friendly [72, 73].

12.8 Conclusions
In conclusion, bio-based adhesives are alive and well, and research on their
development and application is definitely expanding. Very serious industrial
interest exists in these products, always with an eye on their environmen-
tal acceptability and also on their economic and technical viability. Some
clear industrial applications are already emerging. Almost certainly, further
materials for bio-based adhesives will emerge in the future, as well as further
improvements in the materials presented in this review will occur. There is
equally no doubt that at first perhaps for niche applications, and later on for
more widespread applications, the use of these materials is likely to expand.

References
1. Pizzi, A., Advanced Wood Adhesives Technology, Marcel Dekker, New York,
1994.
Bio-Based Adhesives 287

2. Pizzi, A., Natural phenolic adhesives 2: Lignin, in: Handbook of Adhesive


Technology, 2nd edn, A. Pizzi and K.L. Mittal (Eds.), Marcel Dekker, New
York, 2003.
3. Pizzi, A., Tannin based adhesives, in: Wood Adhesives Chemistry and
Technology, vol. I, A. Pizzi (Ed.), Marcel Dekker, New York, 1983.
4. Kulvik, E., Chestnut wood tannin extract in plywood adhesives. Adhes. Age,
19, 3, 19–21, 1976.
5. Spina, S., Zhou, X., Segovia, C., Pizzi, A., Romagnoli, M., Giovando, S.,
Pasch, H., Rode, K., Delmotte, L., Phenolic resin adhesives based on chestnut
hydrolysable tannins. J. Adhes. Sci. Technol., 27, 2103–2111, 2013.
6. Pizzi, A., Tannin based adhesives. J. Macromol. Sci. Chem., C18, 2, 247–307,
1980.
7. Trosa, A., Développement et application industrielle de résines thermodur-
cissables a base de produits naturels de déchet et leur produits de copolyméri-
sation avec des résines synthétiques pour application aux panneaux composites
de bois. Dissertation, University Henri Poincaré—Nancy 1, France, 1999.
8. Trosa, A., Thermosetting adhesive composition based on condensed tannins
and use thereof in the wood industry, 1999.
9. Trosa, A. and Pizzi, A., A no-aldehyde emission hardener for tannin-based
wood adhesives. Holz Roh Werkst., 59, 266–271, 2001.
10. Trosa, A. and Pizzi, A., Industrial hardboard and other panels binder from
tannin/furfuryl alcohol in absence of formaldehyde. Holz Roh Werkst., 56,
213–214, 1998.
11. Trosa, A. and Pizzi, A., Industrial hardboard and other panels binder from
waste lignocellulosic liquors/phenol–formaldehyde resins. Holz Roh Werkst.,
56, 229–233, 1998.
12. Santiago-Medina, F.J., Foyer, G., Pizzi, A., Calliol, S., Delmotte, L., Lignin-
derived non-toxic aldehydes for ecofriendly tannin adhesives for wood pan-
els. Int. J. Adhes. Adhes., 70, 239–248, 2016.
13. Pizzi, A., Hardening mechanism of tannin adhesives with hexamine. Holz
Roh Werkst., 52, 229, 1994.
14. Pichelin, F., Kamoun, C., Pizzi, A., Hexamine hardener behaviour—Effects
on wood glueing, tannin and other wood adhesives. Holz Roh Werkst., 57,
305–317, 1999.
15. Kamoun, C. and Pizzi, A., Mechanism of hexamine as a non-aldehyde poly-
condensation hardener, Part 1. Holzforsch. Holzverwert., 52, 1, 16–19, 2000.
16. Kamoun, C., Pizzi, A., Zanetti, M., Upgrading of MUF resins by buffering
additives—Part 1: Hexamine sulphate effect and its limits. J. Appl. Polym.
Sci., 90, 203–214, 2003.
17. Walker, J.F., Formaldehyde. Am. Chem. Soc. Monogr. Ser., 159, 1964.
18. A. Pizzi, W. Roll, B. Dombo, Bindemittel auf der Basis von Gerbstoffen.
European Patent EP-B 0 648 807 1998; German Patent DE 44 06 825 A1
1995; US Patent 5,532,330 1996.
19. Pichelin, F., SWOOD unpublished results, 2004.
288 Adhesives for Wood and Lignocellulosic Materials

20. Pizzi, A., Valenzuela, J., Westermeyer, C., Low-formaldehyde emission, fast
pressing, pine and pecan tannin adhesives for exterior particleboard. Holz
Roh Werkst., 52, 311–315, 1994.
21. Meikleham, N., Pizzi, A., Stephanou, A., Induced accelerated autocondensa-
tion of polyflavonoid tannins for phenolic polycondensates, Part 1: 13C NMR,
29
Si NMR, X-ray and polarimetry studies and mechanism. J. Appl. Polym.
Sci., 54, 1827–1845, 1994.
22. Pizzi, A. and Stephanou, A., Comparative and differential behaviour of pine
vs. pecan nut tannin adhesives for particleboard. Holzforsch. Holzverwert.,
45, 2, 30–33, 1993.
23. Pizzi, A. and Stephanou, A., A comparative 13C NMR study of polyflavonoid
tannin extracts for phenolic polycondensates. J. Appl. Polym. Sci., 50, 2105–
211, 1993.
24. Pizzi, A. and Meikleham, N., Induced accelerated autocondensation of poly-
flavonoid tannins for phenolic polycondensates—Part III: CP-MAS 13C NMR
of different tannins and models. J. Appl. Polym. Sci., 55, 1265–1269, 1995.
25. Pizzi, A., Meikleham, N., Stephanou, A., Induced accelerated autoconden-
sation of polyflavonoid tannins for phenolic polycondensates—Part II:
Cellulose effect and application. J. Appl. Polym. Sci., 55, 929–933, 1995.
26. Pizzi, A., Meikleham, N., Dombo, B., Roll, W., Autocondensation-based,
zero-emission, tannin adhesives for particleboard. Holz Roh Werkst., 53,
201–204, 1995.
27. Garcia, R. and Pizzi, A., Polycondensation and autocondensation networks
in polyflavonoid tannins, Part 1: Final networks. J. Appl. Polym. Sci., 70,
1083–1091, 1998.
28. Garcia, R. and Pizzi, A., Polycondensation and autocondensation networks
in polyflavonoid tannins, Part 2: Polycondensation vs. autocondensation.
J. Appl. Polym. Sci., 70, 1093–1110, 1998.
29. Pizzi, A., Kueny, R., Lecoanet, F., Masseteau, B., Carpentier, D., Krebs, A.,
Loiseau, F., Molina, S., Ragoubi, M., High resin content natural matrix–
natural fibre biocomposites. Ind. Crops Prod., 30, 235–240, 2009.
30. Blanchet., P., Cloutier, A., Riedl, B., Particleboard made from hammer milled
black spruce bark residues. Wood Sci. Technol., 34, 11–19, 2000.
31. Lopez-Suevos, F. and Riedl, B., Effects of Pinus pinaster bark extracts content
on the cure properties of tannin-modified adhesives and on bonding of exte-
rior grade MDF. J. Adhes. Sci. Technol., 17, 1507–1522, 2003.
32. Kim, S. and Kim, H.-J., Curing behaviour and viscoelastic properties of pine
and wattle tannin based adhesives studied by dynamic mechanical thermal
analysis and FT-IR-ATR spectroscopy. J. Adhes. Sci. Technol., 17, 1369–1384,
2003.
33. L.R. Calvé, Fast cure and pre-cure resistant cross-linked phenol–formaldehyde
adhesives and methods of making same, Canada Patent 2,042,476, 1999.
34. Shimatani, K., Sono, Y., Sasaya, T., Preparation of moderate-tempera-
ture setting adhesives from softwood kraft lignin. Part 2. Effect of some
Bio-Based Adhesives 289

factors on strength properties and characteristics of lignin-based adhesives.


Holzforschung, 48, 337–342, 1994.
35. Gardner, D. and Sellers, T., Jr., Formulation of a lignin-based plywood adhe-
sive from steam-exploded mixed hardwood lignin. Forest Prod. J., 36, 5,
61–67, 1986.
36. Newman, W.H. and Glasser, W.G., Engineering plastics from lignin-XII.
Synthesis and performance of lignin adhesives with isocyanate and melamine.
Holzforschung, 39, 345–353, 1985.
37. Azarov, V.I., Koverniskii, N.N., Zaitseva, G.V., Dodatek ligniny do klejów UF,
Izvestjia Vysshikh Uchnykh Zavedenii. Lesnai Zhurnal, 5, 81–85, 1985.
38. L. Viikari, A. Hase, P. Quintus-Leina, K. Kataja, S. Tuominen, L. Gadda,
Lignin-based adhesives and a process for the preparation thereof. European
Patent EP 09503029 A1, 1999.
39. Nimz, H.H., Lignin-based adhesives, in: Wood Adhesives Chemistry and
Technology, A. Pizzi (Ed.), pp. 247–288, Marcel Dekker, New York, 1983.
40. Nimz, H.H. and Hitze, G., The application of spent sulfite liquor as an adhe-
sive for particleboards. Cell. Chem. Technol., 14, 371–382, 1980.
41. Shen, K.C., Spent sulphite liquor binder for exterior waferboard. Forest
Prod. J., 27, 5, 32–38, 1977.
42. K.C. Shen, Adhesive Composition, Patent Convention Treaty WO 9837148, 1998.
43. Kharazipour, A., Haars, A., Shekholeslami, M., Hüttermann, A., Enzymge-
bundene holzwerkstoffe auf der basis von lignin und phenoloxidase. Adhäsions,
35, 5, 30–36, 1991.
44. Kharazipour, A., Mai, C., Hüttermann, A., Polyphenoles for compounded
materials. Polym. Degrad. Stabil., 59, 237–243, 1998.
45. Felby, C., Pedersen, L.S., Nielsen, B.R., Enhanced auto adhesion of wood
fibers using phenol oxidases. Holzforschung, 51, 281–286, 1997.
46. Pizzi, A. and Stephanou, A., Rapid curing lignins-based exterior wood adhe-
sives, Part 1: Diisocyanates reaction mechanisms and application to panel
products. Holzforschung, 47, 439–445, 1993.
47. Pizzi, A. and Stephanou, A., Rapid curing lignins-based exterior wood adhe-
sives, Part 2: Acceleration mechanisms and application to panel products.
Holzforschung, 47, 501–506, 1993.
48. Navarrete, A., Mansouri, H.R., Pizzi, A., Tapin-Lingua, S., Benjelloun-Mlayah,
B., Rigolet, S., Synthetic-resin-free wood panel adhesives from low molecular
mass lignin and tannin. J. Adhes. Sci. Technol., 24, 1597–1610, 2010.
49. El Mansouri, N., Yuan, Q., Huang, F., Investigation of curing and thermal
behavior of benzoxazine and lignin mixtures. J. Appl. Polym. Sci., 125, 1773–
1781, 2012.
50. Zhong, Z., Sun, X.S., Wang, D., Ratto, J.A., Wet strength and water resistance
of modified soy protein adhesives and effects of drying treatment. J. Polym.
Environ., 11, 4, 137–144, 2003.
51. Liu, Y. and Li, K., Chemical modification of soy protein for wood adhesives.
Macromol. Rapid Comm., 23, 739–742, 2002.
290 Adhesives for Wood and Lignocellulosic Materials

52. Sun, X. and Bion, K., Shear strength and water resistance of modified soy
protein adhesives. J. Am. Oil Chem. Soc., 76, 977–980, 1999.
53. Hettiarachy, N.S., Kalapotly, U., Myers, D.J., Alkali-modified soy protein with
improved adhesive and hydrophobic properties. J. Am. Oil Chem. Soc., 72,
1461–1464, 1995.
54. Amaral-Labat, G.A., Pizzi, A., Goncalves, A.R., Celzard, A., Rigolet, S.,
Environment-friendly soy flour-based resins without formaldehyde. J. Appl.
Polym. Sci., 108, 624–632, 2008.
55. Lei, H., Pizzi, A., Navarrete, P., Rigolet, S., Redl, A., Wagner, A., Gluten pro-
tein adhesives for wood panels. J. Adhes. Sci. Technol., 24, 1583–1596, 2010.
56. Lage, M.C., Pizzi, A., Redl, A., Phenol–wheat protein–formaldehyde adhe-
sives for wood-based panels. Pro Ligno, 10, 3, 3–17, 2014.
57. Lagel, M.C., Pizzi, A., Redl, A., Al-Marzouki, F.M., Phenol–wheat protein–
formaldehyde adhesives for wood-based panels. Eur. J. Wood Prods., 73, 439–
448, 2015.
58. Wescott, J.M., Frihart, C.R., Lorenz, L., Durable soy-based adhesives,
in: Proceedings Wood Adhesives 2005, Forest Products Society, Madison,
Wisconsin, 2006.
59. Belgacem, M.N. and Gandini, A., Furan-based adhesives, in: Handbook of
Adhesive Technology, 2nd edn, A. Pizzi and K.L. Mittal (Eds.), pp. 615–634,
Marcel Dekker, New York, 2013.
60. Chen, C.-M., State of the art report: Adhesives from renewable resources.
Holzforsch. Holzverwert., 48, 4, 58–60, 1996.
61. Conner, A.H., River, B.H., Lorenz, L.F., Carbohydrate modified phenol–
formaldehyde resins. J. Wood Chem. Technol., 6, 591–596, 1986.
62. Conner, A.H., Lorenz, L.F., River, B.H., Carbohydrate-modified PF resins
formulated at neutral conditions, in: Adhesives from Renewable Resources.
vol. 385, ACS Symposium series, pp. 355–369, 1989.
63. Alma, M.H., Yoshioka, M., Yao, Y., Shiraishi, N., The preparation and flow
properties of HC1 catalyzed phenolated wood and its blends with commer-
cial novolak resin. Holzforschung, 50, 85–90, 1996.
64. Miller, R. and Shonfeld, U., Company Literature, Preform Raumgliederungssysteme
GmBH, Esbacher Weg 15, D-91555 Feuchtwangen, Germany, 2002.
65. Sauget, A. and Pizzi, A., Exploratory results for composites of natural fibre
mats with a natural matrix of epoxidized vegetable oils. J. Adhes. Sci. Technol.,
27, 3–8, 2013.
66. Wool, R.P., Kusefoglu, S.H., Khot, R.N., Zhao, R., Palmese, G., Boyd, A.,
Williams, G., Affordable bio-derived plastics and binders for the compos-
ite industry, in: Proceedings, Second European Panel Products Symposium,
pp. 233–240, Bangor, UK, 1998.
67. Tomkinson, J., Adhesives based on natural resources, in: Wood Adhesion
and Glued Products: Wood Adhesives, M. Dunky, A. Pizzi, M. van Leemput
(Eds.), pp. 46–65, European Commission, Directorate General for Research,
Brussels, 2002.
Bio-Based Adhesives 291

68. Bailey, P.S., Ozonation in Organic Chemistry, Academic Press Inc., New York,
1978.
69. Thebaut, M., Pizzi, A., Fredon, E., Synthesis of resins with ozonized sun-
flower oil and radiata pine tannins. J. Renew. Mat., 1, 4, 242–252, 2013.
70. Gfeller, B., Zanetti, M., Properzi, M., Pizzi, A., Pichelin, F., Lehmann, M.,
Delmotte, L., Wood bonding by vibrational welding. J. Adhes. Sci. Technol.,
17, 1425–1590, 2003.
71. Leban, J.-M., Pizzi, A., Wieland, S., Zanetti, M., Properzi, M., Pichelin, F.,
X-ray microdensitometry analysis of vibration-welded wood. J. Adhes. Sci.
Technol., 18, 673–685, 2004.
72. Mansouri, H.R., Pizzi, A., Leban, J.-M., Delmotte, L., Lindgren, O., Vaziri,
M., Causes of the characteristic improved water resistance in pine wood lin-
ear welding. J. Adhes. Sci. Technol., 25, 1987–1995, 2011.
73. Pizzi, A., Mansouri, H.R., Leban, J.-M., Delmotte, L., Omrani, P., Pichelin, F.,
Enhancing the exterior performance of wood linear and rotational welding.
J. Adhes. Sci. Technol., 25, 2717–2730, 2011.
74. Pizzi, A., Zhou, X., Navarrete, P., Segovia, C., Mansouri, H.R., Placentia-
Pena, M.I., Pichelin, F., Enhancing water resistance of welded dowel wood
joints by acetylated lignin. J. Adhes. Sci. Technol., 27, 3, 252–262, 2013.
75. Amirou, S., Pizzi, A., Delmotte, L., Citric acid as waterproofing additive in
butt joints linear wood welding. Eur. J. Wood Prods., 75, 651–654, 2017.
76. Pizzi, A., Leban, J.-M., Kanazawa, H., Properzi, M., Pichelin, F., Wood dowels
bonding by high speed rotation welding. J. Adhes. Sci. Technol., 18, 1263–
1278, 2004.
77. Kanazawa, F., Pizzi, A., Properzi, M., Delmotte, L., Pichelin, F., Influence
parameters in wood dowels welding by high speed rotation. J. Adhes. Sci.
Technol., 19, 1025–1038, 2005.
78. Ganne-Chedeville, C., Pizzi, A., Thomas, A., Leban, J.-M., Bocquet, J.-F.,
Despres, A., Mansouri, H.R., Parameter interactions in two-block welding
and the wood nail concept in wood dowels welding. J. Adhes. Sci. Technol.,
19, 1157–1174, 2005.
13
Environmental Aspects of Adhesives—
Emission of Formaldehyde

13.1 Introduction
There has been heightened awareness and sensitivity in recent times among
the public the world over about the environmental and health concerns
and the concern on global warming. As a result of this growing apprehen-
sion, adhesives, paints, and coatings have come under increasing scrutiny
since the mid-1970s in respect of emissions in the workplace as well in the
service environments. Formaldehyde emission from wood panel products
is one of such important environmental issues. Increasingly stringent regu-
lations are being enacted specifying new limits for formaldehyde emission
for wood panels from time to time. These are incorporated in specified
standards in Europe, USA, and Japan [1].
Formaldehyde, as an aqueous solution ranging from 37 to 50 wt%, is
the preferred aldehyde for reaction with phenol, urea, or melamine for the
preparation of resin adhesives for wood panel products. Over 30 million
metric tons of formaldehyde is the worldwide consumption of formalde-
hyde for an array of products. These include phenol–formaldehyde (PF)
resins, urea–formaldehyde (UF) resins, melamine–formaldehyde (MF)
resins, polyacetal resins, methylenebis (4-phenyl isocyanate), butanediol,
pentaerythritol, and others.
There are two basic processes to produce formaldehyde from methanol.
They are the silver catalyst process and the metal oxide process.
As regards the health issues connected with formaldehyde, a contro-
versy exists currently regarding the classification of formaldehyde as a
human carcinogen. In 2004, the International Agency for Research on
Cancer (IARC) of the World Health Organization reclassified formalde-
hyde from group 2A (“The agent is probably carcinogenic to humans”) to
group 1 (“The agent is carcinogenic to humans”). By the end of October
2009, despite strong disagreement among participants of the voting body,

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (293–315) © 2019
Scrivener Publishing LLC

293
294 Adhesives for Wood and Lignocellulosic Materials

IARC concluded that there is sufficient evidence in humans to link formal-


dehyde with leukemia. The National Toxicology Program in its 14th Report
of Carcinogens published on November 3, 2016, classified formaldehyde as
a “known human carcinogen”.
Formaldehyde (HCHO) is released from wood panel products such
as particleboard, hardwood plywood, and medium-density fiberboard
(MDF) made by employing UF resins and those made with PF resins are
softwood plywood, oriented strandboard, etc. [2–5].
Formaldehyde gas is also emitted from motor vehicles, combustion of
many products, and heating wood during drying.

13.2 Scientific Analysis of the Problem


Traditional wood adhesives are thermosetting resins, namely, phenolic
resins and amino resins. UF and MF resins belong to the family of amino
resins. Both phenolic and amino resins are based on formaldehyde as one
of the major reactants. Thus, 95% of the total wood adhesives used for
panels is composed of formaldehyde-based resins. Out of this, UF resins
account for nearly 85% of the total; melamine, 10%; and phenolics, 5%.
The formaldehyde emission from these two class of adhesives depends
on two factors:

(a) Residual formaldehyde present in the resin


(b) Formaldehyde that is evolved due to the susceptibility of
the cured resin to hydrolytic attack

Figures 13.1 and 13.2 depict the respective degree of susceptibility to


hydrolytic attack of the chemical bonds connecting the structural units
existing in UF resins, MF resins [6], and PF resins:
It can be seen from the above figures that the methylene (-CH2-) and
methylene ether (-CH2-O-CH2-) bridges are both prone to hydrolytic
attack. MF resins are more resistant to hydrolysis due to the stabilization of
the C–N bond by the heterocyclic structure of melamine and to the repel-
lancy to water of its triazine ring. In the case of phenolic resins, the C–C
bonds are highly stable and not subject to hydrolytic attack.
It should be mentioned that wood itself generates traceable amount
of formaldehyde when exposed to certain conditions to which common
wood panel products are exposed [7–9]. This so-called “native” formalde-
hyde has been shown to be transient and rapidly decreases to levels below
those set by the standards [10].
Emission of Formaldehyde 295

C CH2
H2N N N
H H
Highly susceptible
1a
UREA FORMALDEHYDE
RESIN

C CH2 CH2
H2N N O N
H H

Most susceptible
1b

UREA FORMALDEHYDE
RESIN

Figure 13.1 Susceptible bonds in UF.

NH2

C
N N

C C CH2
CH2
N N
NH2 H N
H
1c

MELAMINE-FORMALDEHYDE Stabilization of the C-N bond


RESIN due to the heterocyclic
structure of melamine

OH
H2
C

C-C bond
very strong
1d
PHENOL-FORMALDEHYDE
RESIN

Figure 13.2 Susceptible bonds in MF and PF.


296 Adhesives for Wood and Lignocellulosic Materials

Thus, unbound formaldehyde from UF polymers and copolymers


remaining after the pressing process and the hydrolysis of the cured resin
during service conditions have been found to be the cause of formaldehyde
emission.

13.3 Factors Affecting the Amount of Formaldehyde


Emission
The amount of formaldehyde liberated from UF-bonded wood panels,
for example, depends on a number of factors, which are endogenic and
exogenic. Endogenic factors are related to the resin characteristics and also
the conditions employed for the production of the wood panels. Exogenic
factors are the ambient conditions, namely, relative humidity and tem-
perature, under which the boards are tested [11–15]. Another important
exogenic factor is the aging of the boards. It has been reported that aging
decreases the emission of formaldehyde significantly [13,14].
Other exogenic conditions to which formaldehyde-based products
may be subjected to could be heating or vibration such as grinding,
sanding, cutting, sawing, etc. These factors can lead to increased levels of
formaldehyde in the air. In a closed environment or room, the presence
of formaldehyde-containing materials may give rise to air contamina-
tion, which depends on [9]:

• material properties typical for a given type of wood panel


(or furniture made thereof) and a given set of production
conditions
• room temperature
• relative humidity
• loading factor (= ratio of emitting surface area to room
volume)
• air exchange rate
• air velocity at the emitting surface and potentially of
• sinks (e.g. reactive surfaces) continuously removing
formaldehyde.

For finished panels, formaldehyde release is highest in newly manufac-


tured pressed wood products and decreases over time. Experiments have
demonstrated that in a stable environment (temperature and humidity),
formaldehyde release does decrease over time and the low initial values
Emission of Formaldehyde 297

typical of particleboard and MDF panels will decrease by at least 50%


within a few weeks of manufacture.

13.4 Exposure
• Formaldehyde is normally present in both indoor and out-
door air at low levels, usually less than 0.03 parts of formal-
dehyde per million parts of air.
• When formaldehyde is present in the air at levels exceeding
0.1 ppm, some individuals may experience adverse effects
such as watery eyes; burning sensations in the eyes, nose,
and throat; coughing; wheezing; nausea; and skin irritation.
Some people are very sensitive to formaldehyde, whereas
others have no reaction to the same level of exposure.

In the UK, formaldehyde has been assigned a Maximum Exposure Limit


of two parts per million (ppm). In the United States, the Occupational
Safety and Health Administration (OSHA) has set a permissible exposure
level of 0.75 ppm. In Sweden and Germany, the maximum permissible
indoor level is 0.1 ppm. Many other nations have similar exposure level
regulations in place.

13.5 Safe Level of Formaldehyde Exposure


A number of countries have adopted regulations for indoor air quality. In
regard to environment regulations, Japan maintains the most stringent
HCHO emission in the world with the F**** standard. The US Department
of Housing and Urban Development has specified limits for formaldehyde
emitting products with underlayment and decking materials limited to
emissions of 0.2 ppm and panelling and other products limited to 0.3 ppm.
The Europeans have adopted the E1 regulation, which limits formaldehyde
emissions from products to a level that produces a maximum indoor air
concentration of 0.1 ppm. The Japanese currently have the most stringent
regulations under their “Sick House Legislation”. Thus, indoor environ-
ment products must conform to the F**** classification, which restricts
emissions to 0.03 ppm. This is equivalent to background levels and is effec-
tively a zero emission limit. Internationally, the generally accepted emis-
sion limit is 0.1 ppm.
298 Adhesives for Wood and Lignocellulosic Materials

13.6 Evolution of Formaldehyde Emission Standards


The organizations involved in formulating the formaldehyde emission
standards are given below:

13.6.1 US HUD Manufactured Housing Standard


This standard specifies a 0.20 ppm emission limit for (non-structural)
plywood using the ASTM E1333 method [16].

13.6.2 California Air Resources Board (CARB) Air Toxic


Control Measure for Composite Wood Products
In the United States, formaldehyde is regulated by a number of agencies.
Perhaps the most important is the CARB. According to the studies by
CARB, the major sources of exposure to formaldehyde are from the com-
posite wood products containing UF resins.
CARB approved on April 26, 2007, an airborne toxic control measure
(ATCM) with a view to reducing formaldehyde emissions from compos-
ite wood products, including hardwood plywood, particleboard, MDF,
thin MDF (thickness ≤ 8 mm), and also furniture and other finished
products made with composite wood products. Thus, all the wood prod-
ucts above come under the scrutiny of the ATCM. The law was passed
accordingly, limiting the amount of formaldehyde emissions from com-
posite wood products and restricting formaldehyde emission to very
low limits. The ATCM that controls formaldehyde emissions from com-
posite wood products was approved on April 18, 2008, by the Office of
Administrative Law.
Accordingly, CARB promulgated the ATCMs in two phases. Phase 1
emission standards were to take effect on either January 1, 2009, or July
1, 2009, depending on the composite wood product type. Likewise, the
more stringent Phase 2 emission standards were to take effect between
January 1, 2010, and July 1, 2012 [17–20]. The details are given in
Table 13.1.
It has to be mentioned that all standards are “caps,” meaning that they
cannot be exceeded. Compliance with emission standards must be veri-
fied using a CARB-approved third-party certifier. Manufacturers who
use “ultra-low-emitting” or “no-added formaldehyde” resins can apply to
CARB for approval to test their products less frequently or for an exemp-
tion to third-party certification requirements.
Emission of Formaldehyde 299

Table 13.1 Phase 1 and Phase 2 CARB emission standards.


Phase 1 emission standards:
By January 2009 Hardwood Plywood-Veneer Core = 0.08 ppm
Particleboard = 0.18 ppm
Medium-Density Fiberboard = 0.21 ppm
By July 2009 Hardwood Plywood Composite Core = 0.08 ppm
Phase 2 emission standards:
By January 2010 Hardwood Plywood-Veneer Core = 0.05 ppm
Particleboard = 0.09 ppm
Medium-Density Fiberboard = 0.11 ppm
By January 2012 Hardwood Plywood Composite Core = 0.05 ppm

13.7 CARB Green Adhesive Formaldehyde Emission


Standards
As given in Table 13.1, CARB Phase 1 requires adhesives to have formal-
dehyde emissions that are equal to or less than 0.08 ppm. These standards
well exceed the formaldehyde emission standards set forth by the U.S.
Department of Labor OSHA and the U.S. American National Standards
Institute (ANSI), as well as European E1 emission standards.
According to CARB Phase 2, the adhesive formaldehyde emission stan-
dards in California will fall to a more stringent 0.05 ppm, which exceeds
both E1 and E0 formaldehyde emission standards for green adhesives.

13.8 Japanese JIS/JAS Formaldehyde Adhesive


Emission Standards [21–23]
In response to the so-called Sick Building Syndrome, a public health con-
cern due to poor indoor air quality, the Japanese Industrial Standards
(JIS) Committee amended the Japanese Building Standard Code in 2002.
Accordingly, new stringent standards had to be adopted, which would
restrict the levels of permissible formaldehyde emission. A number of wood
panel products employed in buildings such as plywood, wood flooring,
structural panels, laminated veneer lumber, adhesives, paints, and many
300 Adhesives for Wood and Lignocellulosic Materials

others have been included to conform to the new standards. The Japanese
developed their own rating system in their JIS and Japanese Agricultural
Standards (JAS) to apply to all the formaldehyde emitting building materi-
als mentioned above. Thus, with respect to environment regulations, Japan
maintains the most stringent HCHO emission control in the world with
the F**** requirement.
The Japanese JIS/JAS emission standards are classified as F*, F**, F***, and
F****. Because the emission standards are measured in mg/m2h, it makes it
difficult to compare ratings to those given by the California Air Resource
Board (CARB) Phase 1 and Phase 2 standards and the European E1 and E2
Formaldehyde Emission Standards, which are measured in parts per million
(ppm) formaldehyde emission ratings, which are commonly used worldwide.

(a) Japanese JIS/JAS F*—Adhesives with this emission level


have a formaldehyde emission rate that is greater than
0.12 mg/m2h. This is considered a high level of formal-
dehyde emissions and is not permitted for use in Japan.
Engineered hardwood, bamboo, or laminate floors with
an F* rating are known as Type 1 formaldehyde-emitting
building materials and would not be considered “green”.
(b) Japanese JIS/JAS F**—With this Japanese JIS/JAS stan-
dard, the formaldehyde emission rate of an adhesive
cannot exceed 0.12 mg/m2h. These are known as Type 2
formaldehyde-emitting building materials and they have
restricted areas of use.
(c) Japanese JIS/JAS F***—Formaldehyde emissions must
not exceed 0.005 mg/m2h with the F*** emission rating.
These materials are called Type 3 formaldehyde-emitting
building materials and, like F**, also have limited approved
areas of use.
(d) Japanese JIS/JAS F****—Any adhesive with the F****
mark can proudly claim a formaldehyde emission rate
below 0.005 mg/m2h. These adhesives are approved for use
in Japan without restrictions.

To sum it up, composite wood products that contain adhesives that meet
or exceed Japanese JIS/JAS F**** Emission Standards are the most pre-
ferred of the Japanese ratings when it comes to choosing wood products
made with green adhesives. Under this, panels meeting the most stringent
formaldehyde requirements (F****) are required, using test method JIS A
1460, to have average emission levels below 0.30 mg/l.
Emission of Formaldehyde 301

13.9 European Formaldehyde Emission Standards


[24–33]
E1 and E0 are the European formaldehyde emission standards. E1 emis-
sion standards are well known in the flooring industry and have been used
for years. Wood flooring adhesives that meet E1 formaldehyde standards
have less than 0.75 formaldehyde parts per million (ppm).
E0 is an updated version of E1 with much more stringent standards
requiring formaldehyde emissions to be equal to or less than 0.07 ppm.
Therefore, composite wood products that meet E0 standards would be
considered a safer, greener choice than those that meet E1 standards. Thus,
E1 standards are equivalent to those set forth by the U.S. Department of
Labor OSHA that are deemed by manufacturers as “eco–friendly”.
Both the California Air Resource Board Phase 2 CARB Formaldehyde
Emission Standards and the Japanese Emission Standards JIS/JAS F**** are
more stringent and would therefore be preferred over those that only meet
E1 or E0 standards.

13.10 Standardization and Test Methods


“Test Standards” may be divided into three types of methods.

13.10.1 Reference Methods


Formaldehyde release into the air from wood panel products has been
measured under reality-simulated conditions in a test chamber developed
at the Fraunhofer WKI, Germany. This has become the main reference
method and adopted in the European standard. The reference method is
expensive and time-consuming.
EN717-1 is a reference standard that specifies methods for 40-m3, 1-m3,
and 0.25-m3 chambers [28].
Simple test methods were developed concurrently with the reference
methods that are less expensive, and the formaldehyde emission from
wood panel products could be evaluated within a short period of time. The
results obtained from these test methods can be correlated with the results
derived from test chamber procedure (the reference method). The simpli-
fied test procedures are meant to facilitate the manufacturers to monitor
continuously the formaldehyde emissions from the produced wood-based
materials. The examples of these methods are certification methods and
quality control methods that are described below.
302 Adhesives for Wood and Lignocellulosic Materials

13.10.2 Certification Methods


There are a number of certification methods that are used for certifying
product for sale. These methods are easier and cheaper to set up and run,
but give reasonable correlations with the reference methods. The two
methods commonly used are EN120 [24], the toluene extraction method,
and JISA5908 [21], the Japanese desiccator method.

13.10.3 Quality Control Methods


For in-house use, a number of quick methods have been developed. These
methods allow production to be monitored regularly but are not normally
used for certification of a product. In this category, there are various flask
methods using a plastic bottle to simulate the test chamber. There are
also a number of specialized devices such as the DMC (Dynamic Micro
Chamber) and FLEC (Field and Laboratory Emission Cell), and these have
the advantage of rapid analysis.

13.11 Different Standards and Test Methods


Different standards and test methods evolved in advanced countries are
given in Table 13.2.

Table 13.2 Standards and test methods.


Test method Standard designation
1. Chamber EN 717-1, ASTM-E 1333, ASTM D
6007, JIS A 1901, JIS A 1911, ISO
12460-1, ISO 12460-2
2. Gas analysis EN 717-2, ISO 12460-3
3. Flask method EN 717 3, AWPA method
4. Desiccator method ASTM D 5582, JIS A 1460, JAS 235,
JAS 233, ISO 12640-4
5. Perforator method EN 120, ISO 12460-6
6. Ficid and Laboratory Emission Cell FLEC
Emission of Formaldehyde 303

13.11.1 Reference Method


13.11.1.1 Chamber Methods
Basic principle: Steady-state conditions (ECA REPORT 2)
Principle
Test pieces of known surface area are placed in a chamber, in which the
temperature, relative humidity, air velocity, and exchange rate are con-
trolled at defined values. Formaldehyde emitted from the test pieces
mixes with the air in the chamber. The air in the chamber is sampled
periodically. The formaldehyde concentration is determined by draw-
ing air from the chamber through gas washing bottles containing water,
which absorbs the formaldehyde. The formaldehyde concentration in the
water is determined. The concentration of formaldehyde in the cham-
ber atmosphere is calculated from the concentration in the water in the
gas washing bottles and the volume of the sampled air. It is expressed
in milligrams per cubic meter (mg/m3). Sampling is periodically contin-
ued until the formaldehyde concentration in the chamber has reached a
steady state.
The chamber method is one that allows the measurement of steady-state
concentrations of formaldehyde emitted from wood-based materials in a
reproducible way. Chambers of very different sizes have been used to study
steady-state concentrations of formaldehyde. The chamber sizes fall essen-
tially into two classes:

• Small-scale chambers with volumes ranging from about


0.02 to 1 m3, and
• Large-scale chambers of the “walk-in” type with volumes
ranging from about 12 m3 to about 80 m3.

13.11.1.2 ASTM E 1333 [16]


This test method is based on a chamber of 22 m3 (800 ft3) minimum size.
In this method, the formaldehyde concentration in air and emission
rate from wood products containing formaldehyde are measured in a
large chamber under specific test conditions of temperature and relative
humidity. The objective of the method is to simulate the performance of
the formaldehyde-containing products in actual service. This method
employs a single set of environmental conditions but different product
loading ratios to assess formaldehyde concentrations in air and emission
rates from certain wood products.
304 Adhesives for Wood and Lignocellulosic Materials

The procedure consists of the following:

(a) Conditioning of specimens prior to testing.


(b) Exposing the surface area of the specimens in the test
chamber.
(c) Controlling the test chamber temperature and relative
humidity.
(d) Changing the number of air changes per hour.
(e) Providing air circulation within the chamber.
(f) At the end of a 16- to 20-h period in the test chamber, the
air is sampled and the concentration of formaldehyde in
air and the emission rate are determined.

The quantity of formaldehyde in the air sample taken from the chamber
is determined by an adaptation of the National Institute for Occupational
Safety and Health (NIOSH) chromotropic acid test procedure (Figure 13.3).
If another analytical procedure is used to determine the quantity of
formaldehyde in the air sample, that procedure has to give results that are
equivalent to or of greater accuracy and precision than the adapted chro-
motropic acid procedure. Detailed procedures based on acetylacetone,
pararosaniline (see Test Method), 2,4-dinitrophenylhydrazine (DNPH)
(see Test Method D5197), and 3-methyl-2-benzothiazoline (MBTH) (see
Test Method D5014) have been found to give results equivalent to or of
greater accuracy and precision than the chromotropic acid test. The test
report shall note the analytical procedure employed.

13.11.1.3 ASTM D6007-02(2008) Standard Test Method


for Determining Formaldehyde Concentration in Air
from Wood Products Using a Small-Scale Chamber [34]
Since the ASTM E-1333 requires a large chamber and larger samples,
ASTM D 60007-02 (2008) was introduced. This test method allows testing

OH OH O3S OH HO SO3
O=

+ C H++ O

O3S SO3 H H [O]

Chromotropic acid Formaldehyde


SO3 SO3

Figure 13.3 Reaction between chromotropic acid and formaldehyde.


Emission of Formaldehyde 305

of smaller samples and reduces the time required for testing. Accordingly,
this test method requires the use of a chamber of 0.02 to 1 m3 in volume
to evaluate the formaldehyde concentration in air using the following con-
trolled conditions: product loading, temperature, relative humidity, air
exchange rate, and air circulation within the chamber.
The results may be correlated to values obtained from Test Method E 1333.
The following relationship should be kept in mind with respect to
comparing different standards for formaldehyde emission:
At 23°C and 1 013 hPa, the following relationship exists for
formaldehyde:

1 ppm (parts per million) = 1.24 mg/m3


1 mg/m³ = 0.81 ppm (parts per million)

13.11.1.4 ISO 12460-1 and Part 2: 2007 [35]


The ISO 12460-1:2007 is a reference method and specifies a 1-m3 chamber
method for the determination of the formaldehyde emission from wood-
based panels under defined conditions, relating to typical conditions in
real life. Conversely, ISO 12460-2 2007 is a derived method.

13.11.1.5 Japanese Small Chamber Method JIS A1901 [22]


To determine the emission rate of formaldehyde, a sample is tested in
accordance with the provisions of JIS A 1901. JIS A 1901 is a standard
on “Determination of the emission of volatile organic compounds and
aldehydes for building products—Small chamber method”. A 20-l capacity
chamber is widely used in Japan. The test conditions are given below:

Test temperature: 28°C ± 1°C


Relative humidity: 50% ± 5%
Ventilation rate: 0.5 1/h ± 0.05 1/h
Loading factor: 2.2 m2/m3 as a standard

13.11.2 Derived Methods


13.11.2.1 Gas Analysis according to EN 717-2 [29]
Measuring principle
The gas analysis method according to DIN EN 717-2 is a quick measuring
method in which the formaldehyde emissions of wood-based panels are
306 Adhesives for Wood and Lignocellulosic Materials

measured for 4 h at 60°C. The formaldehyde emission is indicated in mg/


m²h. The edges of the samples are sealed prior to testing. The filtered and
predried airflow in the measuring chamber is 60 l/h. The air with the form-
aldehyde emitted by the test sample is fed from the measuring chamber
into wash bottles with water in which the formaldehyde is absorbed for a
time period of 1 h each (Figure 13.4). Afterwards, the formaldehyde con-
tent is spectrophotometrically determined by means of the acetylacetone
method at a wavelength of 412 nm.

13.11.2.2 Flask Method


This European Standard specifies a method, known as the flask method,
for determination of formaldehyde release from uncoated wood-based
panels. The principle of the flask method for measuring the formaldehyde
release has been published by the Wilhelm-Klauditz-Institute [36].

(a) Principle
Formaldehyde release is determined by suspending test pieces
of known mass over water in a closed container, maintained

Figure 13.4 Emission according to gas analysis.


Emission of Formaldehyde 307

at a constant temperature (Figure 13.5). Formaldehyde


released from the test pieces during a defined period of time
is absorbed by the water. The formaldehyde content of the
water is determined photometrically by the acetylacetone
method, and the result is expressed in milligrams of formal-
dehyde per kilogram of dry board.
(b) Apparatus
(a) Polypropylene or polyethylene flask container of 500 ml
volume with tightly fitting lid of the same material.
(b) Metal test piece holder or rubber band and hook. Metal
parts shall be of stainless steel.
(c) Determination of formaldehyde concentration of the aque-
ous solution

The determination of formaldehyde is based on the Hantzsch reaction


[37] in which aqueous formaldehyde reacts with ammonium ions and

3
125

4
≈40

76

1 500 ml polyethylene bottle with bottle top


2 Hook out of stainless steel
3 Elastic rubber band
4 Surface of water
Dimensions in milimetres

Figure 13.5 Test apparatus for the flask method.


308 Adhesives for Wood and Lignocellulosic Materials

O=
O=
2CH3-C-CH2-C-CH3 + HCHO + NH4+
Acetylacetone Formaldehyde ammonium ion

OH

O=
C C
H3C CH3
N
Diacetyldihydrolutidine

Figure 13.6 Reaction between acetylacetone and formaldehyde.

acetylacetone to yield diacetyldihydrolutidine (DDL) (Figure 13.6). DDL has


an absorption maximum at 412 nm. The reaction is specific to formaldehyde.

13.11.2.3 Desiccator Method


(a) ASTM D5582-00(2006) standard test method for deter-
mining formaldehyde levels from wood products using
a desiccator [38]
This test method incorporates a desiccator, with the desiccant
removed, having a 250-mm (10-in.) inside diameter and a vol-
ume of approximately 10.5 l (641 in.) with the desiccator lid in
place (Figure 13.7). Specified number, size, and edge sealing
of wood specimens are to be placed in the desiccator. A petri
dish containing water is placed in the desiccator and closed.
Formaldehyde is absorbed by the water in the petri dish.
Samples of 25 ml from the petri dish are analyzed for formal-
dehyde at the end of a 2-h period in the closed desiccator.
Wood products typically evaluated by this test method
are made with UF adhesives and include particleboard,
hardwood plywood, and MDF. This test method is used for
product quality control and is a small bench test method
that correlates with the large-scale acceptance test, the Test
Method E 1333 for determining formaldehyde levels from
wood products.
(b) Japanese desiccator method [23]
The test pieces are conditioned to constant mass in an atmo-
sphere with a mean relative humidity of 65% ± 5% and a
temperature of 20°C ± 2°C.
Emission of Formaldehyde 309

Figure 13.7 Japanese desiccator method.

Method
JIS A 1460 (2001) for building boards (to be updated 2005)
JIS A 5905 (2003) for fiberboards—refer to JIS A 1460 method
JIS A 5908 (2003) for particleboards—refer to JIS A 1460 method
JAS 240 (2003) for flooring
JAS 233 (2003) for plywood
(SOP-KEMI-050)

The details of the test method are summarized below:

(a) Desiccator material: Glass


(b) Desiccator volume: 9–11 L
(c) Sample size: Total area of 1800 cm2 including butt ends
and double surface.
~10 pieces, 150 ± 1 mm × 50 ± 1 mm
(d) Sealing: No
310 Adhesives for Wood and Lignocellulosic Materials

(e) Conditioning: ≥1 day in plastic at 20°C ± 1°C (after cut-


ting of test pieces) for JAS. One week open at 20°C ± 2°C,
65% ± 5% RH for JIS methods.
(f) Test duration: 24 h
(g) Temperature: 20°C ± 0.5°C
(h) Quantity of water for absorption of formaldehyde: 300 ±
1 ml water in a petri dish
(i) Air exchanges: Static measurement
(j) Results in mg/l

13.11.2.4 Criteria of Acceptance for Different Grades are Given


in the following Table:

Emission classes Average (mg/l) Maximum (mg/l)


F**** ≤0.3 ≤0.4
F*** ≤0.5 ≤0.7
F** ≤1.5 ≤2.1
F* ≤5 ≤7

A desiccator method has also been more recently approved as an ISO


standard test method [39].

13.11.2.5 The Perforator Method (EN 120) [24]


Principle
The total extractable formaldehyde content of MDF or OSB is measured
using the perforator method. For this test, approximately 500 g of board,
cut into small pieces, is boiled in toluene in a special set of laboratory
glassware (Figure 13.8). The toluene absorbs the formaldehyde contained
in the board pieces. It, in turn, is washed with water and the amount of
formaldehyde originally contained in the board sample but now in solu-
tion in the water is determined by chemical analysis using a colorimet-
ric method. The results are expressed as milligrams of formaldehyde in
100 grams of board (mg/100 g).
To compensate for the influence of different moisture contents on the
perforator value in the range 3% ≥ u ≤ 9%, the measured perforator values
Emission of Formaldehyde 311

Figure 13.8 The perforator method.

have to be adjusted to a moisture content of 6.5% by a formula published


by Jann and Deppe (DIBt 100) [40] as indicated in EN 312

F = −0.133u + 1.86

where F is the perforator correction factor and u is the percentage moisture


content.
The perforator method is the most common laboratory method
employed worldwide, especially in manufacturing and quality control of
wood panel products. The test is very sensitive to the moisture content in
312 Adhesives for Wood and Lignocellulosic Materials

the material and in the toluene during the extraction of the free formalde-
hyde [41]. Storage conditions and the species of wood can also affect the
perforator value [42]. The test is only suitable for uncoated PB and MDF
and cannot be used to characterize emission or aging of the material [43].
It has also been found to be unreliable to measure emission from very low
emitting panels [44].

References
1. Young, S., Formaldehyde Emissions—Understanding the Standards, 2004.
http://www.eximcorp.co.in/knowledge-documents/formaldehyde_emissions_
understanding_the_standards_timber_test_new_zealand.pdf.
2. Godish, T., Residential formaldehyde contamination: Sources and levels.
Comments. Toxicology, 2, 3, 115–134, 1988.
3. Otson, R. and Fellin, P., Gaseous Pollutants: Characterization and Cycling,
J.O. Nriagu (Ed.), pp. 335–421, John Wiley and Sons, New York, 1992.
4. Kelly, T.J., Smith, D.L., Satola, J., Emission rates of formaldehyde from
materials and consumer products found in California homes. Environ. Sci.
Technol., 33, 81–88, 1999.
5. Zelenius, O., Standards and regulations concerning the formaldehyde emis-
sions from wood panels. 13th International Conference: Standardization,
Prototypes and Quality: Means of Balkan Countries Collaboration, Brasov,
Romania, November 3–4, 2016.
6. Dunky, M., Urea–formaldehyde (UF) adhesive resins for wood. Int. J. Adhes.
Adhes., 18, 95–107, 1998.
7. Frihart, C.R., Wescott, J.M., Birkeland, M.J., Gonner, K.M., Formaldehyde
Emissions from ULEF- and NAF-Bonded Commercial Hardwood Plywood as
Influenced by Temperature and Relative Humidity. Proceedings, International
Convention Society of Wood Science and Technology and United Nations
Economic Commission for Europe—Timber Committee, Geneva, Switzerland,
October 11–14, 2010.
8. Schäfer, M. and Roffael, E., On the formaldehyde release of wood. Holz Roh
Werkst., 58, 259–264, 2000.
9. Roffael, E., Volatile organic compounds and formaldehyde in nature, wood
and wood based panels. Holz Roh Werkst., 64, 144–149, 2006.
10. Birkeland, M.J., Lorenz, L., Wescott, J.M., Frihart, C.R., Determination of
native (wood derived) formaldehyde by the desiccator method in particle-
boards generated during panel production. Holzforschung, 64, 429–433,
2010.
11. Peterson, H., Reuther, W., Eisele, W., Wittmann, O., Zur Formaldehydabspaltung
bei der Spanplattenerzeugung mit Harnstoff-Formaldehyd-Bindemitteln: 1.
Mitteilung. Holz Roh Werkst., 30, 429–436, 1972.
Emission of Formaldehyde 313

12. Peterson, H., Reuther, W., Eisele, W., Wittmann, O., Zur Formaldehydabspaltung
bei der Spanplattenerzeugung mit Harnstoff-Formaldehyd-Bindemitteln: 2.
Mitteilung: der Einfluss der Festharzmenge, Preßzeit, und Preßtemperatur.
Holz Roh Werkst., 31, 463–469, 1973.
13. Sundin, B. and Roffael, E., Einfluß der Alterung auf die Formaldehyd-
emissionen von UF-Spanplatten niedrigen Formaldehydabgabepotentials.
Holz-Zent., Bl, 115, 704, 1989.
14. Sundin, B., Estimation of formaldehyde emission from composite wood prod-
ucts. PhD Dissertation, Czech University of Life Sciences, Prague, 1989.
15. Sundin, B., Risholm-Sundman, M., Edenholh, K., Emission of formaldehyde
and other volatile organic compounds (VOC) from sawdust and lumber,
different wood-based panels and other building materials: a comparative
study, in: Proceedings, 26th International Particleboard/Composite Materials
Symposium, Washington State University, Pullman, USA, 1992.
16. ASTM, Standard test method for determining formaldehyde concentrations
in air and emission rates from wood products using a large chamber. Method
ASTM E 1333–96, p. 12, American Society for Testing and Materials, West
Conshohocken, PA, 2002b.
17. California Air Resources Board CARB, Identification of Formaldehyde as
a Toxic Air Contaminant. Part A. Exposure Assessment, p. 103, Technical
Support Document, Stationary Source Division, Sacramento, CA, 1992.
18. California Air Resources Board CARB, Composite Wood Survey Package,
2003. Accessible at: http://www.arb.ca.gov/toxics/compwood/Survey2002.pdf.
19. California Air Resources Board CARB. Composite Wood Products, ATCM
(homepage), California Air Resources Board, Sacramento, CA, 2007,
Accessible at http://www.arb.ca.gov/toxics/compwood/compwood.htm.
20. California Air Resources Board CARB, California Air Resources Board,
2008. http://www.arb.ca.gov/regact/2007/compwood07/fro–final.pdf.
21. Japanese Industrial Standard JIS A 5908, JIS standard specification for parti-
cleboard, Japanese Standards Association, Tokyo, Japan, 1994.
22. Japanese Industrial Standard JIS A 1901, Determination of the emission
of volatile organic compounds and aldehydes for building products–small
chamber method, Japanese Standard, January 2003, Japanese Standards
Association, Tokyo, Japan, 2003.
23. Japanese Industrial Standard JIS A 1460. Building boards, Determination of
formaldehyde emission–desiccator method, March 2001, Japanese Standards
Association, Tokyo, Japan, 2001.
24. European Norm EN 120, Wood-based panels—Determination of form-
aldehyde content–extraction method called perforator method, European
Standard, European Committee for Standardization, Bruxelles, Belgium,
1993, September 1993.
25. European Norm EN 312, Particleboard—Specifications, European Standard,
European Committee for Standardization, Bruxelles, Belgium, September
1993, 2003.
314 Adhesives for Wood and Lignocellulosic Materials

26. European Norm EN 322, Wood-based panels—Determination of moisture


content, European Committee for Standardization, Bruxelles, Belgium, 1993.
27. European Norm EN 622–1, Fibreboards. Specifications, General require-
ments, European Committee for Standardization, Bruxelles, Belgium, 2003.
28. European Norm EN 717–1, Wood-based panels—Determination of form-
aldehyde release—Part 1: Formaldehyde emission by the chamber method,
European Standard, European Committee for Standardization, Bruxelles,
Belgium, 2004.
29. European Norm EN 717–2, Wood-based panels—Determination of form-
aldehyde release—Part 2: Formaldehyde release by the gas analysis method,
European Standard, European Committee for Standardization, Bruxelles,
Belgium, 1994.
30. European Norm EN 717–3, Wood-based panels—Determination of formal-
dehyde release—Part 3: Formaldehyde release by the flask method, European
Standard, European Committee for Standardization, Bruxelles, Belgium,
1996.
31. European Norm EN 13986, Wood-based panels for use in construction.
Characteristics, evaluation of conformity and marking, European Committee
for Standardization, Bruxelles, Belgium, 2002.
32. European Norm EN 13419-1, Building products. Determination of the
emission of volatile organic compounds. Emission test chamber method,
European Committee for Standardization, Bruxelles, Belgium, 1999.
33. European Collaborative Action, Indoor air quality and its impact on man
ECA REPORT 2 COST PROJECT 613. Directorate-General for Research
and Innovation, European Commission, Brussel, 1996.
34. ASTM, Standard test method for determining formaldehyde concentration in
air from wood products using a small-scale chamber. Method D 6007–02, p. 8,
American Society for Testing and Materials, West Conshohocken, PA, 2002.
35. ISO 12460-1 and-2, Wood-based panels—Determination of formaldehyde
release—Part 1: Formaldehyde emission by the 1-cubic-metre chamber method,
European Committee for Standardization, Bruxelles, Belgium, 2007.
36. Roffael, E., Messung der Formaldehydabgabe—Praxisnahe methode zur
Emittlung der Formaldehydabgabe harnstoffgebundener Spanplatten fur das
Bauwesen. Holz-Zent., 111, 1403–1404, 1975.
37. Hantzsch, A., Condensationprodukte aus Aldehydammoniak und Ketonartigen
Verbindungen. Chem. Ber., 14, 2, 1637–1638, 1881.
38. ASTM D5582–00, Standard Test Method for Determining Formaldehyde
Levels from Wood Products Using a Desiccator, American Society for Testing
and Materials, West Conshohocken, PA, 2006.
39. ISO/CD 12460–4, Wood-based panels—Determination of formaldehyde
release—Desiccator method, The European Committee for Standardization,
Bruxelles, Belgium, 2007.
40. DIBt 100, Richtlinie über die Klassifizierung und Überwachung von
Holzwerkstoffplatten bezüglich der Formaldehydabgabe (DIBt Richtlinie 100),
Emission of Formaldehyde 315

DIN-German Istitute for Standardization, Beuth Verlag, Berlin, Germany,


1994.
41. Marutzky, R., Release of formaldehyde by wood products, in: Wood Adhesives
Chemistry and Technology, vol. 2, A. Pizzi (Ed.), pp. 307–387, Marcel Dekker,
New York, 1989.
42. Romeis, M., Influence of time and storage on perforators. WKI Rep No 19.
Fraunhofer Institute for Wood Research, Braunschweig, 1988.
43. Flentge, A. and Meyer, B., Extruded particleboards—formaldehyde release—
ageing, WKI Short Rep 7/90, Wilhelm Klauditz Institut, Braunschweig,
Germany 1990.
44. Roffael, E., Johnsson, B., Engström, B., On the measurement of formaldehyde
release from low-emission wood-based panels using the perforator method.
Wood Sci. Technol., 44, 369–377, 2010.
14
Rheology and Viscoelasticity of Adhesives

14.1 Rheology of Adhesives


Rheology is the science that deals with the deformation and flow of flu-
ids. This is a very important subject for adhesives and adhesion since, for
instance, wood adhesion involves flow, spreading, transfer, and penetra-
tion of adhesives into the wood adherends. Furthermore, the consistency
of the adhesives is an important property to be controlled during the man-
ufacture of adhesives. In addition, the flow properties directly or indirectly
control the reactivity of the adhesives during the curing process. Wood
adhesives consist not only of pure oligomers but also of additives such as
fillers, extenders, surface active agents, etc. These formulated adhesives
exhibit non-Newtonian flow characteristics. Hence, rheological methods
are used to test the flow properties and the reactivity of adhesives during
their routine production as well as during the development of new adhe-
sive systems. The physical property of importance for the study of rheology
is the viscosity. Hence, the theory of viscosity has to be understood first.

14.2 Viscosity—Theory
Viscosity is a measure of “resistance to flow” [1]. Figure 14.1 illustrates this.
Two plates with the area A contain a thin layer of fluid. The bottom
plate does not move, while the top plate is forced to move because of the
force F [N]. The shear stress is defined as

F
[N/ m 2 ]
A

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (317–345) © 2019
Scrivener Publishing LLC

317
318 Adhesives for Wood and Lignocellulosic Materials

F
A
dux

d
y

Figure 14.1 Viscous flow between two parallel plates.

The upper plate is moving at a velocity dux [m/s]. The shear rate is
defined as the velocity gradient

dux 1
[s ]
dy

If the shear stress and the shear rate are known, the dynamic viscosity
can be calculated as

[P a. S]

1 centipoise [cP] = 0.001 pascal second [Pa·s] or 1 cP = 1 mPa·s (one


millipascal-second). or 1 pascal second is equal to 10 poise.
The dynamic viscosity has the dimensions: ML−1 T−1
If M = 1 g, L= 1 cm, and T = 1 s, then ML−1 T−1 = 1 g cm−1 s−1 = 1 Poise
(g cm−1 s−1)
1 Poise = 100 centipoises.

14.3 Capillary Viscometry


The viscosity of even very dilute polymer solutions is considerably higher
than that of solvents. The increase in viscosity depends on several factors
such as temperature, solvent–polymer interaction, polymer concentration,
and the size of polymer molecules. The dependence of viscosity on poly-
mer size permits one to estimate an average molecular weight from the solu-
tion viscosity. The average molecular weight that is measured is called the
viscosity average molecular weight M v to distinguish it from M n and  M W.
Rheology and Viscoelasticity of Adhesives 319

The relationship between the increase in viscosity and the molecular weight


is given by the Mark–Howink–Sakurada equation:

[η] = K Ma

K and a are empirical constants for a given polymer–solvent and tem-


perature combinations and a increases in the range 0.5 with the degree
of expansion of the molecular coils from their unperturbed dimensions.
Under “theta” conditions, the molecular coils have unperturbed dimen-
sions a = 0.5. The Mark–Howink constants K and a are evaluated from [η]
and molecular weight (M) of calibration samples with narrow molecular
weight distribution. Having determined the constants K and a, the molec-
ular weight of polymers can be determined from the viscosity data. The
viscosity method of molecular weight determination requires very little
investment in apparatus and can be carried out quite rapidly by flow time
measurements with a simple glass viscometer (capillary viscometer) set up
in a constant temperature bath (Figure 14.2) [2].
Capillary viscometers basically are used for Newtonian, incompress-
ible, and wall-adherent liquids. The flow inside the capillary is assumed
to have the following idealized conditions: laminar, incompressible, and
stationary. Moreover, the flow influence at the entry and exit of the cap-
illary is neglected and the viscosity of fluids is assumed to be pressure
independent.

Upper
mark

Lower
mark

Figure 14.2 Ostwald’s viscometer.


320 Adhesives for Wood and Lignocellulosic Materials

Capillary viscometers are working based on the Hagen–Poiseuille Law


as their physical basis:

v πR 4ΔP
t 8L

in which V, t, R, ΔP, L, and η are the volume, the time, the radius of the
capillary, the pressure drop, the length of the capillary, and the dynamic
viscosity of fluids, respectively. The volume flow measurement through the
capillary at a given differential pressure is the fundamental measurement
criteria for capillary viscometers. In other words, the viscosity is deter-
mined by measuring the time required for a defined volume of liquid to
flow through a capillary tube due to the hydrostatic pressure of the liquid
column itself.
Two marks (the upper mark and lower mark) before and after a ball-
shaped extension are reference points that enable measurement of the time
of flow of the liquid through the capillary.
ΔP in the above equation = hρg, where h is the average height of the
height of the liquid during measurement, ρ is the density of the polymer
solution, and g is the acceleration due to gravity. Since all the quantities
except η are known, the viscosity η can be calculated.

14.4 Rotational Viscometers


Rotational viscometers function on the principle that the force required
to turn an object in a fluid is a measure of the viscosity of that fluid [3].
A rotating spindle (cylinder or disk) is submerged in the sample fluid.
The liquid offers resistance to the rotational motion of the spindle. The
force (torque) required to overcome this resistance against the rotation
of the spindle at a constant speed is measured. Thus, the rotational vis-
cometers work by measuring the torque on a vertical shaft that rotates
a spindle. The spindle is in the test sample and its rotation is impeded
in proportion to the viscosity of the sample. With precise calibration,
these instruments measure viscosity. Rotational viscometers fall into
two main types:

1. Spring type
2. Servo system
Rheology and Viscoelasticity of Adhesives 321

14.4.1 Spring Type


This system uses a spring to capture the torque exerted by the liquid on the
spindle. These systems use sets of calibrated springs.
A spring viscometer system is based on a pivot and spring concept. A
pivot/spring assembly rotates on the shaft. The spindle is attached to this
assembly. As the spindle rotates, the drag of the fluid against the spindle
causes the spring to deflect. This deflection is proportional to the torque
caused by the viscosity of the sample. This relationship to viscosity is calcu-
lated automatically [4]. The spring system has an advantage in that it offers
measurements of very high accuracy [5].

14.4.2 Servo Systems


The servo system is the second system for the determination of the torque.
A precision servo motor is used to drive the shaft. The spindle is attached
directly to the shaft instead of a pivot, as in the case of the spring system.
The current needed to drive the shaft rotating at the test speed is first deter-
mined. When the spindle is submerged in the liquid and is allowed to rotate
therein, the drag of the liquid exerted on the spindle makes the system use
more current to keep the spindle running at the set speed. This current is
an indirect measure of the torque and it is proportional to the viscosity.
Viscosity is taken as to be directly proportional to the total amount of cur-
rent produced. The servo system is reported to be more advantageous than
the spring system.

14.5 Cone-and-Plate Viscometer


A cone-and-plate viscometer is an instrument for determining the abso-
lute viscosity of liquids in small sample volumes. Its cone-and-plate
geometry is such that complete rheological properties can be precisely
determined [5].

Principle of operation
A cone-and-plate viscometer is a precise torque meter that is driven at dis-
crete rotational speeds. The sample of the liquid is introduced in the space
between the cone and the plate. The torque measuring system consists of
a calibrated spring (e.g., beryllium–copper) that connects the drive mech-
anism to a rotating cone and senses the resistance to rotation caused by
liquid between the cone and a stationary flat plate. The resistance to the
322 Adhesives for Wood and Lignocellulosic Materials

rotation of the cone produces a torque that is proportional to the shear


stress in the fluid. With a specific diameter, angle of cone, torque, and the
angular velocity the shear stresses, shear rates and viscosity can be deter-
mined by using appropriate equations [4, 6].
The major advantage of this measuring system is the constant shear rate
throughout the sample. Also, the sample size is small that results in less
heat generated at high shear rates. Another advantage is the ease of clean-
ing the system.

14.6 Parallel Plate Viscometer


The parallel plate viscometer consists of two identical, coaxial and parallel
circular plates. The shear rate and shear stress and hence the viscosity are
calculated from the radius of a circle with the center of the plate, radius of
the plate, the gap between two plates, torque, and the angular velocity by
employing the appropriate equations [5, 7].
Parallel plate systems are referred to by the diameter of the upper plate.
For instance, a PP40 is a 40-mm-diameter plate. The lower plate is either
larger than or the same size as the upper plate.
The major application of this measuring system is to study the curing
behavior of different wood adhesives. Some work on the curing of phenol–
formaldehyde (PF) resins has been reported [8].

14.7 Concentric Cylinder Viscometer


The concentric cylinder viscometer consists of two concentric cylinders
with different radii. The rheological behavior of samples filled in the gap
between the two cylinders is determined. If the gap is small enough com-
pared to the diameters of the cylinders, shear rate is approximately con-
stant across the sample in the gap. The shear rate, shear stress, and viscosity
can be determined from the radius of the outer cylinder, the radius of the
inner cylinder, the angular velocity, effective immersed length of the sam-
ple, and the torque from the appropriate equations [5].

14.8 Ford Cup Viscosity


The Ford viscosity cup (Figure 14.3) is a simple device for the determina-
tion of time of flow of a known volume of liquid through an orifice located
Rheology and Viscoelasticity of Adhesives 323

Figure 14.3 Ford cup.

at the bottom. This device is employed to check continuously the time of


flow of the resin as the reaction proceeds in the reaction vessel. This time of
flow indicates the extent of polymerization of the resin in the kettle. Under
ideal conditions, this rate of flow would be proportional to the kinematic
viscosity (expressed in stokes and centistokes) that is dependent on the
specific gravity of the draining liquid. However, the conditions in a simple
flow cup are seldom ideal for making true measurements of viscosity.
The Ford cup is used mainly to determine the time of flow of a resin
during manufacture and to determine the completion of the reaction as
well as maintain the consistency of the quality of the resin. The tempera-
ture of the cup and the liquid is maintained, as ambient temperature makes
a significant difference to viscosity and thus flow rate.
The original Ford cup was based on imperial (US) measurement of the
aperture. Many other types of flow cups are being used, depending on the
industry or region, such as Din Cup 4  mm; standard DIN 53211 (can-
celled) ISO Cup 2, 3, 4, 5, 6, and 8 mm; standard ISO 2431 AFNOR Cup
2, 5, 4, 6, and 8 mm; and standard NF T30-014 ASTM Cup 1, 2, 3, 4, and 5
standard ASTM D1200.

14.9 Gardner–Holt Tubes


Gardner bubble viscometer tubes are used to quickly determine kinematic
viscosity of known liquids such as resins and varnishes. The bubble vis-
cometer tubes are also described as Gardner–Holt tubes. Repeated read-
ings may be taken easily once the temperature has been controlled. The
principle of operation is based on the time required for an air bubble to
rise in the adhesive from one end to the other end of each tube, which is
directly proportional to the viscosity of the liquid—the faster the bubble
324 Adhesives for Wood and Lignocellulosic Materials

rises, the lower the viscosity. A complete set of Gardner–Holt tubes covers
in general ranges from 0.05 to 1000 stokes.

14.10 Newtonian and Non-Newtonian Fluids


Two types of fluids exist: Newtonian fluids and non-Newtonian fluids.
They are governed by the power law

n
dux
K [N/ m 2 ]
dy

The factor K [Pa sn] is the consistency index or power law coefficient
and n is the power law exponent. For Newtonian flow, n is equal to 1 and
K equals the viscosity. For shear-thinning fluids, n is less than 1 while it is
larger than 1 for shear-thickening fluids.
Newtonian fluids have a constant viscosity independent of shear rate.
Hence, for Newtonian fluids, changing the force (stress) applied to the fluid
will not change their viscosity. The viscosity remains constant as the force
applied changes.
Non-Newtonian fluids instead have viscosities that change according
to the magnitude of the force (stress) that is applied upon the fluid. The
viscosity changes as the force applied changes.

14.10.1 Types of Non-Newtonian Fluid Behavior


The viscosity of non-Newtonian fluids is dependent on shear rate or shear
rate history. The behavior of different types of non-Newtonian fluids is
given in Figure 14.4.
In a Newtonian fluid, the relation between shear stress and the shear
rate is linear, passing through the origin, the constant of proportionality
being the coefficient of viscosity.
For a Bingham Plastic fluid (see Figure 14.4), stress can be applied, but
it will not flow until a certain value, the yield stress, is reached. Beyond
this point, the flow rate increases steadily with increasing shear stress. The
physical reason for this behavior is that the liquid contains particles (such
as clay) or large molecules (such as polymers) that have some kind of inter-
action, creating a weak solid structure, formerly known as a false body,
and a certain amount of stress is required to break this structure. Once
the structure has been broken, the particles move with the liquid under
Rheology and Viscoelasticity of Adhesives 325

Bingham Plastic

Pseudoplastic Fluid

Shear Stress, t
Dilatant Fluid

Newtonian Fluid

Shear Rate, k

Figure 14.4 Non-Newtoninan fluids.

viscous forces. If the stress is removed, the particles associate again. It is


that causes the thixotropic behavior of many adhesives.

(a) Pseudoplastic fluids


They are shear-thinning fluids. The viscosity of these fluids
decreases as the shear rate increases.
(b) Dilatant fluids
Called shear thickening fluids in which the apparent viscosity
increases as the shear rate increases.
(c) Rheopectic fluids
Rheopecty or rheopexy is the rare property of some non-
Newtonian fluids to show a time-dependent increase in vis-
cosity (time-dependent viscosity); the longer the fluid under-
goes shearing force, the higher its viscosity. Rheopectic fluids,
such as some lubricants, thicken or solidify when shaken.
(d) Thixotropic fluids
Thixotropy is a time-dependent shear-thinning property. Certain
gels or fluids that are thick, or viscous, under static conditions
will flow (become thin, less viscous) over time when shaken,
agitated, sheared, or otherwise stressed (time-dependent vis-
cosity). Constant shear stress can be applied by shaking or
mixing.

14.11 Viscoelasticity of Adhesives


Materials respond in specific manners when they are subjected to a force
(stress). An ideal elastic solid follows Hooke’s law, which states that the
326 Adhesives for Wood and Lignocellulosic Materials

stress/strain ratio is constant and is called the Young’s modulus. In con-


trast, an ideal viscous liquid follows Newton’s law, which states that the
stress is proportional to the rate of strain. But there are a number of mate-
rials encountered in normal life that are not ideal. They exhibit both elas-
tic and viscous responses to an applied force. They are called viscoelastic
materials. Polymeric adhesives for instance exhibit such unique viscoelas-
tic responses when subjected to a stress.
Dynamic mechanical analysis (DMA) is the most common technique
for determining viscoelastic properties and their dependence on time and
temperature.

14.11.1 Phenomenological Models for Viscoelastic Materials


Models to describe the behavior of viscoelastic materials are constructed
based on coupled or uncoupled spring and dashpot elements. The spring
represents an ideal elastic material while the dashpot portrays an ideal
Newtonian liquid. Attempts have been made to account for the response
of real materials in terms of suitable combinations of spring and dashspot.
Thus, a viscoelastic material can both store the energy corresponding
to the elastic component and dissipate energy corresponding to the vis-
cous component. In a viscoelastic material, therefore, part of the energy
is stored while the remaining is dissipated through viscous flow. Some of
these models are described below.

14.11.1.1 Maxwell Element (Elastic Deformation + Flow)


A Maxwell element (spring and dashpot in series) (Figure 14.5) symbol-
izes a material that can respond elastically to stress but can also undergo
viscous flow. The two contributions to the strain are additive in this model:
ε  = εelas + εvisc
The force under constant stretch decays exponentially with time in the
Maxwell model. The relaxation time is given by

Kv/Ke

where Ke is the linear spring constant (ratio of force and displacement, in


N/m) and Kv is the linear dashpot constant (ratio of force and velocity, in
Ns/m). Thus, the Maxwell element predicts stress relaxation experiments
but fails to describe a creep experiment. Other viscoelastic models may be
envisioned, and it must be recognized that neither a Maxwell element nor
Rheology and Viscoelasticity of Adhesives 327

I.O

F
X-axis: Time s; Y-axix: σ(t)/σ(0)

Figure 14.5 Maxwell element.

γ∞

F
X-axix: Time (s) and Y-axis: Y-axis: Strain (γ)

Figure 14.6 Voigt element.

any complex combination of springs and dashpots succeeds in portraying


the overall response of a viscoelastic materials.

14.11.1.2 Voigt Element (Spring and Dashpot in Parallel)


A Voigt element symbolizes a material whose elastic response is not
instantaneous but retarded by a viscous resistance. The Voigt model cap-
tures the creep behavior of viscoelastic solids (Figure 14.6).

14.11.1.3 Maxwell–Voigt Mixed Model (Figure 14.7)


G , the storage modulus, represents the elastic deformation of a material
and is a measure of the hardness of the adhesive at a given temperature.
A typical hot melt adhesive has a G at 25°C, varying from 107 to 108 Pa. A
good PSA has a value of 104 to 105 at room temperature.
328 Adhesives for Wood and Lignocellulosic Materials

X-axis: Time; Y-axis: Strain

Figure 14.7 Maxwell–Voigt mixed model.

G , the loss modulus, is associated with energy absorption mechanisms


and correlates to the viscous deformations in a material; i.e., it represents
in a sense the flexibility of an adhesive.

14.12 Dynamic Mechanical Analysis


Adhesives for wood such as urea–formaldehyde (UF), melamine–formal-
dehyde (MF), PF, epoxies, PU, etc. are essentially thermosetting in nature.
They basically exhibit viscoelastic properties. DMA is the most common
technique employed for studying such viscoelastic behavior and its depen-
dence on time and temperature.
DMA and torsional braid analysis (TBA) have shown potential for
detecting thermosetting events [9–12]. For DMA and TBA, glass fiber
braids are typically impregnated with liquid PF, and the cure phenomenon
is detected from changes in storage and loss moduli. Palmese and Gillham
demonstrated that thermoset gelation and vitrification events are accom-
panied by a damping peak as well as a rapid increase in storage modulus
[13]. Likewise, vitrification has been detected in several cure studies of
neat PF resins [12].
In the case of thermosetting resins, as the reaction proceeds from the
liquid state to a gel, there is an increase in molecular weight and viscos-
ity with time initially followed by an accelerated (or abrupt) increase in
viscosity as shown in Figure 14.8. Macroscopically, the thermoset can be
characterized by an increase in its viscosity η approaching infinite vis-
cosity, implying that all the chains become linked to one another at the
gel point through a cross-linked network structure of infinite molecular
Rheology and Viscoelasticity of Adhesives 329

0 Conversion 100%

η0 Ge

Newtonian Network at Hookean


liquid the gel point solid

Figure 14.8 Macroscopic development of rheological and mechanical properties during


network formation.

weight. Figure 14.8 illustrates the macroscopic progress from uncured to


fully cured thermoset [14].
The uncured thermoset, often a mixture of monomers and oligomers, is a
Newtonian liquid. As cure progresses, the viscosity increases with increasing
molecular weight, which can be monitored by rheological measurements. As
the viscosity approaches infinity at the gel point, steady shear measurements
reach their limits. However, oscillatory or dynamic rheology and dynamic
mechanical measurements can characterize material in the gelation region.
Reaction continues beyond the gel point to complete the network forma-
tion. Macroscopically, physical properties such as modulus build to levels
characteristic of a fully developed network. It should be noted that while
DMA may be applied to uncured materials below their gel point, the sam-
ples require a support such as metal shim, glass fabric, or wire mesh [15].
DMA is ideally suited to characterize supported samples from pre-
gelation to the completion of cure, as well as fully cured thermosets.
DMA measures the deformation of a polymeric material as a response
to a small stress applied in a time-varying periodic, e.g., sinusoidal fashion.
The strain response to the sinusoidal stress depends on the property of
the polymeric material. For elastic solids subjected to infinitesimal strain,
Hooke’s law states a linear stress–strain relationship. Hence, under sinusoi-
dal loading of elastic solids, the strain is in-phase with the stress [10, 16].
330 Adhesives for Wood and Lignocellulosic Materials

For a viscous liquid on the other hand, Newton’s law describes a linear
relationship between stress and rate of strain under infinitesimal strain
rate. This implies that viscous liquids respond to a sinusoidal stress with
a strain that is 90° out of phase with the stress.
Thus in the case of an ideal (100%) elastic polymer, the strain will be
in-phase with the stress wave. On the other hand, for an ideal (100%) vis-
cous material, the strain curve will be 90° out of phase (i.e., a cos wave)
with the input sinusoidal stress. This is shown in Figure 14.9.
The dynamic storage modulus (E ), the loss modulus (E ), and the
mechanical damping or internal friction (tan δ = E /E ) are obtained by
this technique. Tan δ gives the ratio of the amount of energy dissipated as
heat to the energy stored during the deformation, and is often the parame-
ter chosen to relate dynamic data to molecular or structural motion in the
sample (Figures 14.9 and 14.10).

δ = 0° δ = 90°

Shear
stress

Shear
strain

100% elastic 100% viscous

Figure 14.9 Response of elastic and viscous materials to sinusoidal stresses.

γ0 strain σ0
stress
Stress σ(t)
Strain γ(t)

0.0 0.0

–γ0 δ –σ0

0 90 180 270 360


ωt (degrees)

Figure 14.10 Strain response to sinusoidal stress.


Rheology and Viscoelasticity of Adhesives 331

In a DMA experiment, two independent properties, storage (E ) and


loss moduli (E ), are measured to provide a complete description of visco-
elastic properties of a polymeric adhesive at a given temperature.
The relationship between the storage and loss modulus is shown below:

E* = E + iE (14.1)

tan δ = E /E (14.2)

Thus, the ratio of loss and storage moduli (Equation 14.2) defines
another useful parameter in DMTA, called tan δ.
E , the storage modulus represents the elastic deformation of a material
and is a measure of the hardness of the adhesive at a given temperature.
A typical hot melt adhesive has a G at 25oC, varying from 107 to 108 Pa.
A good PSA has a value of 104 to 105 at room temperature. E , the loss
modulus, is associated with energy absorption mechanisms and correlates
to the viscous deformations in a material; i.e., it represents in a sense the
flexibility of an adhesive [17].
The loss tangent (tan δ), can be used to define the gel point. The latter
occurs at the point where E´ crosses E and where tan δ equals 1, i.e., as the
crossover point of the storage modulus and the loss modulus (E = E ). The
crossover point is generally accepted as the gel point of gel temperature of
the cure reaction [18].

E”(t) E*(t)

δ(t)

E’(t)

Figure 14.11 Graphical representation of storage and loss moduli (Argand diagram).

E” ~ energy loss in internal


motion

E’ ~ elastic response

Figure 14.12 Significance of elastic response and energy loss.


332 Adhesives for Wood and Lignocellulosic Materials

The physical significance of E and E is illustrated diagrammatically in


Figures 14.11 and 14.12.
That is, if there is no energy loss E , then the ball would have reached the
original position from which it is dropped.
Thus, in a dynamic mechanical test, an oscillating strain (sinusoidal
or other waveform) is applied to a sample and the resulting stress devel-
oped in the sample is measured. The output signals are analyzed, and,
using established mathematical methods, the rheological parameters are
computed.

14.13 TTT and CHT Diagrams


Temperature–time–transformation (TTT) and continuous heating trans-
formation (CHT) cure diagrams provide a complete cure characterization
of thermosetting systems in that they define the time and temperature
required for achieving specific viscoelastic properties. TTT and CHT
diagrams were developed for the first time by Gilham for epoxy resins
[19, 20]. These diagrams define zones of curing, vitrification, and degrada-
tion of resins such as epoxies in their pure state and not in contact with any
resin-interacting substrate. Gilham studied and modeled the times to gela-
tion and to vitrification for the isothermal cure of an amine‐cured epoxy,
measuring on macroscopic and molecular levels by dynamic mechanical
spectrometry (TBA and Rheometrics dynamic spectrometer), infrared
spectroscopy, and gel fraction experiments. The relationships between the
extents of conversion at gelation and at vitrification and the isothermal
cure temperature formed the basis of a theoretical model of the TTT cure
diagram (Figure 14.13), in which the times to gelation and to vitrifica-
tion during isothermal cure versus temperature are predicted. The model
demonstrates that the “S” shape of the vitrification curve depends on the
reaction kinetics, as well as on the physical parameters of the system, i.e.,
the glass transition temperatures of the uncured resin (Tg0), the fully cured
resin (Tg∞), and the gel (gelTg).
In the case of the CHT diagram, Wisanrakkit and Gilham [20] found
that the overall reaction kinetics of a high‐Tg tetrafunctional aromatic
diamine/difunctional epoxy system, which can satisfactorily describe the
rate of the reaction in both kinetically and diffusion‐controlled regimes,
can be determined from isothermal conversion/Tg data by differential scan-
ning calorimetry. They used the mathematical expression of the kinetics,
together with the unique one‐to‐one relationship between Tg and chemical
conversion, to calculate the material’s Tg vs. time under heating at constant
Rheology and Viscoelasticity of Adhesives 333

Degradation

τgo
Devitrification
Ce
ll
ed
ru
bb
er

Celled glass

Culation curve
τcure(°C)

Vitrification curve
Liquid

τgo

Non grilled class

In time

Figure 14.13 Details of a TTT diagram of an epoxy resin adhesive on a non-interacting


glass fiber substrate.

rates. For a heating scan from below Tg0 (the glass transition temperature
of the unreacted material), initial devitrification corresponds to the reac-
tion temperature (Tcure) first passing through the Tg of the reacting mate-
rial; vitrification corresponds to Tg becoming equal to the increasing Tcure
after initial devitrification; and finally, upper devitrification corresponds
to Tg eventually falling below the rising Tcure. The results of the calcula-
tion correlated well with the available experimental data of the dynamic
mechanical behavior of the resin during temperature scans at constant
rates obtained by the TBA.
Although these works were a considerable innovation, they could not be
transferred to resins such as wood adhesives containing water as a solvent
and especially in contact with a lignocellulosic substrate interacting with
both the water and the resin. The diagrams had then to be extensively mod-
ified for the case of water-carried wood adhesives in contact with wood
to take into account the fundamental influence of these two parameters,
water and the absorbent substrate. Furthermore, TTT and CHT diagrams
334 Adhesives for Wood and Lignocellulosic Materials

were difficult and very time-consuming to build with the approach of


Gilham. It is the use of thermomechanical analysis (TMA) techniques that
instead allowed for the first time the determination of these diagrams in an
easier and faster way, just on wood adhesive resins in contact with wood
[9, 21–28].
The use of the equation f = km/αE, correlating number of degrees of
freedom m of polymer segments between cross-linking nodes in polycon-
densation networks to the energy of interaction polymer segment/polymer
segment, both within the same polymer and at different polymer interfaces,
through measures of deflection f in bending by dynamic TMA, yielded
a number of consequences of interest in the field of polycondensation-
hardened networks and of their process of hardening and, among others,
opened the way to an easier determination of TTT and CHT diagrams for
wood adhesives on lignocellulosic substrates [9, 21].
Thus, lignocellulosic substrates such as wood were found to have a
marked modifying influence on a well‐defined region of CHT diagrams
during hardening of PF and UF adhesive resins. This was ascribed to more
complex resin phase transitions due to resin/substrate interactions peculiar
to these substrates. The chemical and physical mechanisms of the interac-
tions of the resin and substrate causing such CHT diagram modifications
were presented and discussed. The Di Benedetto equation [29] describing
the glass transition temperature Tg of the system as a function of the resin
degree of conversion p was then slightly modified to take into account the
modified TTT and CHT diagram [21]. The modified diagrams can be used
to good effect to describe the behavior of polycondensation resins when
used as wood adhesives during their curing directly into the wood joint.
Lignocellulosic substrates such as wood were found to have a marked
modifying influence on both lower-temperature and higher-temperature
zones of TTT and CHT diagrams during hardening of phenol–resorcinol–
formaldehyde (PRF) and melamime–urea–formaldehyde (MUF) polycon-
densates. Although the modifying influence of the substrate on the higher-
temperature zone of CHT diagrams presented the same trend of what
was found first for PF and UF polycondensates, marked differences from
what was reported in the literature were recorded for TTT diagrams of all
these polycondensates as well for the lower-temperature zones of the CHT
diagrams on lignocellulosic substrates [25]. The chemical and physical
mechanisms of the interactions of the resins, the substrate, and the water
carrier causing such marked variations were presented and discussed.
Although in the higher-temperature zones both substrate and water car-
rier play an important role, in the lower-temperature zone, the presence of
water appears to be the dominant factor causing the observed variations.
Rheology and Viscoelasticity of Adhesives 335

The  generalized modified CHT and TTT diagrams characteristic of the


behavior of these water‐borne polycondensates on lignocellulosic sub-
strates can be used to describe the behavior and complex changes of phase
the formaldehyde‐based polycondensation resins undergo when used as
wood adhesives during their curing directly in the wood joint. The results
also show that diagrams obtained with pure resins cannot be used to pre-
dict the behavior of the polycondensate when this is markedly modified by
the presence of interacting solvents and substrates.
Furthermore, the modifying influence of some of the most import-
ant manufacturing parameters of the resins on a CHT diagram were also
explored with some clear trends being shown [27]. In the case of melamine–
urea–formaldehyde (MUF) resins for wood adhesives, the molar ratio
(M + U):F appeared to be the dominant parameter influencing the relative
position of gel and vitrification curves of the CHT diagram in relation to
each other. The ratio of melamine to urea did not appear to have any effect
on the relative position of the curves, lacking any clear trend, at least at the
higher (M + U) molar ratio of 1:1.9 used for this series of resins. Thus, the
influence of resin manufacturing parameters on CHT curing diagrams was
studied in combination with the modifications introduced by the substrate.
TTT and CHT curing diagrams were also built for tannin‐based adhe-
sives by TMA by following the in situ hardening directly in a wood joint
[30]. The curve trends observed were similar to those previously observed
for synthetic polycondensation resins on lignocellulosic substrates. Of the
parameters that most influence the relative position of vitrification and
gel curves on the diagrams (i.e., where the influence has been quantified),
chief among them is the reactivity of the tannin with formaldehyde and any
factor influencing it: thus, the inherent higher reactivity of the A‐ring of
the tannin (such as in procyanidins versus prorobinetinidins) and the pH
of the tannin solution. The percentage formaldehyde hardener has some
influence in CHT diagrams, especially for the slower‐reacting tannins, but
practically no influence in TTT diagrams within the 4–10% formaldehyde
range used. As in the case of synthetic polycondensation adhesive resins,
regression equations relating the internal bond strength of a wood par-
ticleboard, prepared under controlled conditions, with the inverse of the
minimum deflection, obtained by constant heating rate TMA of a wood
joint during resin cure, have been obtained for the two types of tannins
of lower reactivity (profisetinidins and prorobinetinidins) but not for the
faster‐reacting procyanidin and prodelphinidin tannins.
Different trends to those reported for TTT and CHT diagrams of
epoxy resins reported in the literature occur in the higher- and lower-
temperature zones of the diagrams of water-borne formaldehyde-based
336 Adhesives for Wood and Lignocellulosic Materials

resins hardening on wood. CHT and TTT diagrams have been reported for
PF, UF, MUF, PRF, and tannin–formaldehyde thermosetting resins [24, 25].
The experimental TTT diagrams shown in Figure 14.14 show, however,
quite a different trend from the CHT diagram for the same resins (Figure
14.18) and for the TTT diagrams reported in the literature for epoxies on
glass fiber (Figure 14.13). To start to understand the trend shown in Figure
14.14, it is first necessary to observe what happens to the modulus of the
wood substrate alone (without a resin being present) when examined
under the same conditions of a wood joint during bonding. No significant
degradation occurs up to a temperature of 180°C as shown by the relative
stability of the value of the elastic modulus as a function of time. Some
slight degradation starts to occur at 200°C, but after some initial degrada-
tion, the elastic modulus again settles to a steady value as a function of time
and at a value rather comparable to the steady value obtained at lower tem-
peratures. Evident degradation starts to be noticeable in the 220–240°C
range, and this becomes even more noticeable at higher temperatures. The
effect of substrate degradation on the TTT diagram in Figure 14.14 can
then only start to influence the trends in gel and vitrification curves at tem-
peratures higher than 200°C, and it is for this reason that the region of the
curves higher than 200°C are indicated by segmented lines in Figure 14.13.
At a temperature <200°C, the trends observed are only due to the resin.

Substrate degradation Degradation

Solvent remobilisation from substrate


τ g∞
Devitrification Resin to τg∞
No
n
geGe
llelle Gelled glass
d
ru d ru
Tcure (°C)

Pseudo gel bb bb
curve er er

Gelation curve Vitrification curve

Liquid
To time– ∞
τgo-O°C
Due to solvent
Non gelled glass

In time

Figure 14.14 Generalized TTT diagram of water-carried PF resin adhesives on an


interacting lignocellulosic substrate (wood).
Rheology and Viscoelasticity of Adhesives 337

In this range of temperature, the eventual turning to longer time and stable
temperature of the vitrification curve, characteristic of the TTT diagrams of
epoxy resins, becomes also evident for the TTT diagrams of the water-borne
PF and other formaldehyde resins on lignocellulosic substrates, indicating
that diffusion hindrance at a higher degree of conversion also becomes, for
these resins, the determinant parameter defining reaction rate. What how-
ever differs from previous diagrams is that the trend of all the curves, namely,
gel curve, initial pseudogel (entanglement) curve, and start and end of vit-
rification curve, is the same. In epoxy resins, TTT diagrams of the trend of
the gelation curve are completely different from what was reported here. The
result shown in Figure 14.14 is however rather logical because if diffusion
problems alter the trend of the vitrification curve, then the same diffusional
problem should also alter the gel and pseudogel curves. This is indeed what
the experimental results in Figure 14.14 indicate. It may well be that in water-
borne resins, the effect is more noticeable than in epoxy resins. This is the
reason why it is possible to observe it for PF, UF, PRF, and MUF resins. With
the data available and with the limitation imposed by the start of wood sub-
strate degradation at higher temperatures, it is not really possible to say if the
gel curve and the vitrification curve run asymptotically towards the same
value of temperature at time = ∞, although the indications are that this is
quite likely to be the case. What is also evident in the trend of the two curves
is the turn to the right, hence the inverse trend of their asymptotic tendency
towards Tg∞. This turn cannot be ascribed to substrate degradation because
for very reactive resins, such as PRFs, such a turn already occurs at a tem-
perature lower than 150°C, hence much lower than the temperature at which
substrate degradation becomes significant. This inverse trend can only be
attributed to movements of water coming from the substrate towards the
resin layer as the curve trend indicates an easing of the diffusional problem
already proven to occur at such a high degree of conversion [24, 25].
Two other aspects of the TTT diagrams in Figures 14.13, 14.14, and
14.15 must be discussed, these being the trend of the curves at tempera-
tures higher than 200°C and the trend of the devitrification (or resin deg-
radation) curve. The segmented line trend and experimental points of all
the curves at temperatures higher than 200°C are clearly only an effect
caused by the ever more severe degradation of the substrate: degradation
of the substrate infers a greater mobility of the polymer network consti-
tuting the substrate, hence the continuation of the curves as shown in
their segmented part. That this is the case is also supported by the vir-
tual negative times yielded by the TMA equipment when the temperature
becomes extreme, as well as by the trend of the resin’s higher degradation
curve, which tends to intersect the vitrification curve at about 200–220°C
338 Adhesives for Wood and Lignocellulosic Materials

Degradation

Devitrification curve
No Ge
n lle gelled glass
ge
lle d
Tcure (°C)

d ru
ru bb
bb er
er Vitrification curve
Pseudo gel curve

Gelation curve

Liquid To time–∞
τgo-O°C
Due to solvent
Non gelled glass (frozen sovent)

In time

Figure 14.15 Generalized CHT diagram of water-carried PF resin adhesives on an


interacting lignocellulosic substrate (wood).

or higher, this being a clear indication that one is measuring the changes
in the reference system, the substrate itself, and that these are at this stage
so much more important than the small changes occurring in the resin to
be able to dominate the whole complex system, which is the bonded joint.
The CHT and TTT diagrams pertaining to water-borne formaldehyde-
based polycondensation resins on a lignocellulosic substrate should then
appear in their entirety as shown in Figures 14.14 (TTT) and 14.15 (CHT).

14.14 Experimental Results


Before describing some more recent results in DMA, it will be worth
describing the considerable series of experiments that were done in deter-
mining by TMA by studying the curing behavior and viscoelastic prop-
erties of PF, MUF, UF, and tannin adhesives and their correlation with
the board properties obtained. The method of following the progressive
curing, hardening, and degradation of the adhesives directly on a wood
joint was developed by the research group of Pizzi in the late 1990s and
early 2000s [31–35]. To obtain the TMA samples, the system consisted in
placing the adhesive in its viscous liquid state between two layers of beech
Rheology and Viscoelasticity of Adhesives 339

wood veneer of 0.6-mm thickness as shown in Figure 14.16, and a three-


point bending mode was applied. The adhesive thickness was 0.2 mm, but
the total tested sample dimensions were slightly different from those in
Figure 14.16 at 21 × 6 × 1.4 (mm). The specimens were then tested by
TMA in three points bending on a span of 18 mm exercising a force cycle
of 0.1/0.5 N on the specimens with each force cycle of 12 s (6 s/6 s).
PF, MUF, and UF resin adhesives were tested in this manner at a con-
stant temperature rate of increase of 10°C/min up to the hardening of
the adhesive glue mixes and the TMA results related to the IB strength of
boards made with the same resins [31]. Equally, commercial pine (Pinus
radiata) and mimosa (Acacia mearnsii) bark tannin extract solutions in
water were tested by TMA at different temperatures and at different per-
centage paraformaldehyde hardener addition both under isothermal con-
ditions and at a constant temperature rate of increase of 10°C/min up to
the hardening of the tannin resin glue mix, and the results were compared
with those obtained with the ABES equipment [36]. Equally in the case of
PF and UF adhesives, the effect of humidity on viscoelastic response and
cross-linked and entanglement networking of formaldehyde-based wood
adhesives were tested by the same TMA technique in bending [31]. The
principle of this TMA (and DMA) approach is that as the adhesive is still
a viscous liquid, only the lower beech veneer lamina resents the bending

1.0

141
0.8
Modulus/max. modulus

0.6
119
173
0.4 181
193

0.2

0.0
0 50 100 150 200
Temperature (°C)

Figure 14.16 TMA curve of the hardening of a PF resin in situ in a beech wood joint.
Increase of MOE of the joint as a function of temperature at 10°C/min constant heating
rate (O); first derivative (Δ) (from Ref. [37]).
340 Adhesives for Wood and Lignocellulosic Materials

stress imparted to the sandwich, but as the resin starts progressively to


thicken and to cure, the stress is equally progressively felt and transmit-
ted to both the glue line and to the upper beech lamina of the sandwich.
It is still thus possible not only to follow the progressive variation of the
adhesive to liquid to solid, but also to rapidly compare different adhesives,
or also to compare how the viscoelasticity of an adhesive behaves under
different curing conditions [34].
TMA on wood joints bonded with PF adhesives, as in Figure 14.16, has
shown that, frequently, the joint increase in modulus does not proceed in
a single step but in two steps, yielding an increase of the modulus first
derivative curve presenting two major peaks rather than the single peak
obtained for mathematically smoothed modulus increase curves [37]. This
behavior has been found to be due to the initial growth of the polycon-
densation polymer leading first to linear polymers of critical length for the
formation of entanglement networks. The reaching of this critical length is
greatly facilitated by the marked increase in concentration of the PF poly-
mer due to the loss of water on absorbent substrates such as wood coupled
to the linear increase of the average length of the polymer due to the ini-
tial phase of the polycondensation reaction. The combination of these two
effects lowers markedly the level of the critical length needed for entangle-
ment. Two modulus steps and two first derivative major peaks then occur,
with the first due to the formation of linear PF oligomer entanglement
networks, and the second one due to the formation of the final covalent
cross-linked networks. The faster the reaction of phenolic monomers with
formaldehyde, or the higher the reactivity of a PF resin, the earlier and at
lower temperature the entanglement network occurs, and the higher is its
modulus value in relation to the joint modulus obtained with the final,
covalently cross-linked resin.
Suitable sample geometry for the DMA measurement of liquid adhe-
sives can be the “wire mesh geometry” as shown in Figure 14.17 with the
following details:

• Support for liquid or paste materials.


• Note that the wire mesh is used on a 45° bias; otherwise, the
experiment would be dominated by the high stiffness wire.
When used in the 45° bias, the properties of the resin will
dominate the response.
• Controlled thickness, sensitive measure of Tg.
• Can extract thermoset G , G via composite analysis.
Rheology and Viscoelasticity of Adhesives 341

Stationary
clamp
Wire mesh sample
45° bias

Movable
clamp

Figure 14.17 Wire mesh geometry.

Lei and Frazier [38] used filter paper since this was proven to be suit-
able as the substrate for the preparation of DMA testing specimens to
predict the curing behavior of PF resin adhesives for its stability during
the curing temperature span. With this method, the curing behav-
ior of PF resin was monitored by DMA in tensile-torsion mode. With
the strain curves, the onset of curing temperature of PF resin could be
determined clearly. The curing degree of PF resin could be calculated
by the integral area in strain curves. The method to combine storage
modulus (G ), tan δ, and strain curves together could explain the curing
behavior of PF resin more comprehensively than the commonly used
method using only G and tan δ curves. The DMA test results of PF resin
with different viscosity and with accelerator implied the reliability of
this novelty method.
The viscoelastic properties of the blends of MF resin and poly(vinyl
acetate) were studied by DMTA by Kim and Kim [39]. The DMTA ther-
mogram of MF resin showed that the storage modulus (E) increased as
the temperature was further increased as a result of the cross-linking
induced by the curing reaction of the resin. E of MF resin increased both
as a function of increasing temperature and with increasing heating rate.
Kim and Kim [40] followed on the early work of other groups [9, 21, 30]
on the curing behavior and viscoelastic properties of tannin-based adhe-
sives. They studied the curing behavior and viscoelastic properties of two
types of tannin-based adhesives, wattle and pine, with three hardeners by
DMTA:

(a) Paraformaldehyde
(b) Hexamethylenetetramine
(c) TN [tris(hydroxylmethyl)nitromethane]
342 Adhesives for Wood and Lignocellulosic Materials

28 mm
5 mm 0.2 mm tannin-based adhesive
0.6 mm sliced beecd veneer
1.4 mm

Figure 14.18 Sample configuration for the DMTA (three-point bending mode) test.

Paraformaldehyde was shown to be more reactive with wattle tannin


than the other hardeners, while hexamethylenetetramine was the most
reactive with the pine tannin.
The storage modulus (E ), loss modulus (E ), and loss factor (tan δ) of
each adhesive system were obtained by DMTA. With increasing tempera-
ture, as the tannin-based adhesives hardened, the storage modulus (E )
increased in all adhesive systems.
Kim and Kim [40] imitated the same testing system first developed
much earlier for formaldehyde resins by another research group [9, 21,
26, 30–34, 36, 37, 41], as shown by Figure 14.18. As the adhesives were in
the viscous liquid state, to obtain the DMTA samples, the tannin-based
adhesives were placed between two layers of beech wood veneer of 0.6-mm
thickness as shown in Figure 14.18, and a three-point bending mode was
applied. The adhesive thickness was 0.2 mm, and the total tested sample
dimensions were 28 × 5 × 1.4 (mm). During the DMTA experiments, the
static force was kept at 20% of the dynamic force and the frequency was
maintained at 1 Hz with a strain of 0.05.

References
1. Barnes, H.A., Hutton, J.F., Walters, K., An Introduction to Rheology, Elsevier,
Amsterdam, The Netherlands, 1989.
2. Chanda, M., Advanced Polymer Chemistry, Marcel Dekker, New York, 2000.
3. Anonymous, How does a rotational viscometer work? http://ttwud.org/
how-does-a-rotational-viscometer-work/ 2000.
4. Rheosys, A basic introduction to viscometers and viscosity, http://www.rh[9]
eosys.com/intro.html 2011.
5. Yang, X., Organic fillers in phenol–formaldehyde wood adhesives, PhD
Dissertation, Virginia Polytechnic Institute and State University, Blacksburg,
VA, 2014.
6. Gupta, R.K., Polymer and Composite Rheology, CRC Press, Boca Raton, FL,
2000.
7. Klimets, M., Torque and its calculation, http://earthphysicsteaching.home
stead.com/Torque_Lab.html 2002.
Rheology and Viscoelasticity of Adhesives 343

8. Halasz, L., Vorster, O.C., Pizzi, A., Guasi, K., A rheology study of the gelling
of phenol–formaldehyde polycondensates. J. Appl. Polym. Sci., 80, 898–902,
2001.
9. Pizzi, A., On the correlation of some theoretical and experimental param-
eters in polycondensation cross-linked networks. J. Appl. Polym. Sci., 63,
603–617, 1997.
10. Laborie, M.-P., Investigation of the wood/phenol–formaldehyde adhesive
interphase morphology, PhD Dissertation, Virginia Polytechnic Institute
and State University, Blacksburg, VA, 2002.
11. Steiner, P.R. and Warren, S.R., Rheology of wood–adhesive cure by torsional
braid analysis. Holzforschung, 35, 273–278, 1981.
12. Kim, M.G., Nieh, W.L.-S., Meacham, R.M., Study of phenol–formaldehyde
resol resins by dynamic mechanical analysis. Ind. Eng. Chem. Res., 30, 798–
803, 1991.
13. Palmese, G.R. and Gilham, J.K., Time–temperature-transformation (TTT)
cure diagrams relationship between Tg and the temperature and time of cure
for a polyamic acid/polyimide system. J. Appl. Polym. Sci., 34, 1925–1939,
1987.
14. Winter, H.H., Baumgaertel, M., Sockey, P.R., A parsimonious model for
viscoelastic liquids and solids, in: Techniques in Rheological Measurement,
A.A. Collyer (Ed.), pp. 123–160, Chapman and Hall, London, 1997.
15. Prime, B.R., Dynamic Mechanical Analysis of Thermosetting Materials,
2013. https://ruc.udc.es/dspace/bitstream/handle/2183/11489/CC-80%20
art%2013.pdf?sequence=1&isAllowed=y.
16. Ward, I.M., Mechanical Properties of Solid Polymers, 2nd edn., John Wiley &
Sons, New York, 1979.
17. Franck, A.J., Adhesives Rheology, Brochure Rheometrics, AFERA Congress,
Chester, UK, 1992.
18. Kendall, E.W., Benton, L.D., Trethewey, B.R., Jr., Thermal analysis of the
polymerization of phenol formaldehyde resins. Proc. TAPPI Plast. Laminates
Symp., 1, 117–132, 2000. https://imisrise.tappi.org/TAPPI/Products/DIL/
DIL0408.aspx.
19. Enns, J.B. and Gillham, J.K., Time–temperature–transformation (TTT) cure
diagram: Modeling the cure behavior of thermosets. J. Appl. Polym. Sci., 28,
2831–2844, 1983.
20. Wisanrakkit, G. and Gillham, J.K., Continuous heating transformation
(CHT) cure diagram of an aromatic amine/epoxy system at constant heating
rates. J. Appl. Polym. Sci., 41, 1985–1995, 1990.
21. Pizzi, A., On the correlation of some theoretical and experimental parame-
ters in polycondensation cross-linked networks, Part 2: Interfacial energy and
adhesion on cellulose substrates. J. Appl. Polym. Sci., 65, 1843–1847, 1997.
22. Pizzi, A., Extension of simple polycondensation gelation theories to simple
radical and mixed polycondensation/radical gelation. J. Appl. Polym. Sci., 71,
517–521, 1999.
344 Adhesives for Wood and Lignocellulosic Materials

23. Pizzi, A., On the correlation equations of liquid and solid 13C NMR, TMA,
Tg and network strength in polycondensation resins. J. Appl. Polym. Sci., 71,
1703–1709, 1999.
24. Pizzi, A., Lu, X., Garcia, R., Lignocellulosic substrates influence on TTT and
CHT curing diagrams of polycondensation resins. J. Appl. Polym. Sci., 71,
915–925, 1999.
25. Pizzi, A., Zhao, C., Kamoun, C., Heinrich, H., TTT and CHT curing dia-
grams of water-borne polycondensation resins on lignocellulosic substrates.
J. Appl. Polym. Sci., 80, 2128–2139, 2001.
26. Pizzi, A., Garcia, R., Deglise, X., Thermomechanical analysis of entanglement
networks—Correlation of some calculated and experimental parameters.
J. Appl. Polym. Sci., 67, 1673–1678, 1998.
27. Properzi, M., Pizzi, A., Uzielli, L., Influence of preparation parameters on
the variation of CHT curing diagrams of MUF polycondensation resins.
J. Appl. Polym. Sci., 81, 2821–2825, 2001.
28. Garnier, S. and Pizzi, A., CHT and TTT curing diagrams of polyflavonoid
tannin resins. J. Appl. Polym. Sci., 81, 3220–3230, 2001.
29. Pascault, J.P. and Williams, R.J.J., Glass transition temperature versus
conversion relationships for thermosetting polymers. J. Polym. Sci.,
Part B, Polym. Phys., 28, 85–95, 1990. https://doi.org/10.1002/polb.1990.
090280107.
30. Garnier, S., Huang, Z., Pizzi, A., Commercial tannin adhesives-bonded
particleboard IB forecasting by TMA bending. Holz Roh Werkst., 59, 46,
2001.
31. Kamoun, C., Pizzi, A., Garcia, R., The effect of humidity on cross-linked
and entanglement networking of formaldehyde-based wood adhesives.
Holz Roh Werkst., 56, 235–243, 1998.
32. Laigle, Y., Kamoun, C., Pizzi, A., Particleboard, I.B., forecast by TMA
bending in UF adhesives curing. Holz Roh Werkst., 56, 154, 1998.
33. Zhao, C., Garnier, S., Pizzi, A., Particleboard dry and wet IB forecasting
by gel time and dry TMA bending in PF wood adhesives. Holz Roh Werkst.,
56, 402, 1998.
34. Zhao, C., Pizzi, A., Garnier, S., Fast advancement and hardening acceler-
ation of low condensation alkaline PF resins by esters and copolymerized
urea. J. Appl. Polym. Sci., 74, 359–378, 1999.
35. Kamoun, C. and Pizzi, A., Particleboard IB forecast by TMA bending in
MUF adhesives curing. Holz Roh Werkst., 58, 288–289, 2000.
36. Lecourt, M., Humphrey, P., Pizzi, A., Comparison of TMA and ABES as
forecasting systems of wood bonding effectiveness. Holz Roh Werkst., 61,
75–76, 2003.
37. Garcia, R. and Pizzi, A., Cross-linked and entanglement networks in ther-
momechanical analysis of polycondensation resins. J. Appl. Polym. Sci., 70,
1111–1116, 1998.
Rheology and Viscoelasticity of Adhesives 345

38. Lei, H. and Frazier, C.E., A dynamic mechanical analysis method for pre-
dicting the curing behaviour of phenol–formaldehyde resin adhesive.
J. Adhes. Sci. Technol., 29, 981–990, 2015.
39. Kim, S. and Kim, H.-J., Thermal stability and viscoelastic properties of
MF/PVAc hybrid resins on the adhesion for engineered flooring in under
heating system; ONDOL. Thermochimica Acta, 444, 134–140, 2006.
40. Kim, S. and Kim, H.-J., Curing behaviour and viscoelastic properties of
pine and wattle tannin-based adhesives studied by dynamic mechanical
thermal analysis and FT-IR-ATR spectroscopy. J. Adhes. Sci. Technol., 17,
1369–1383, 2003.
41. Zhao, C., Pizzi, A., Kuhn, A., Garnier, S., Fast advancement and hardening
acceleration of low condensation alkaline PF resins by esters and copoly-
merized urea. Part 2: Esters during resin reaction and effect of guanidine
salts. J. Appl. Polym. Sci., 77, 249–259, 2000.
15
Hot Melt Adhesives

15.1 Introduction
Hot melt adhesives set (i.e., transformation of liquid adhesives into solid)
by cooling. Other types of conversion of a liquid adhesive to solid are by
solvent evaporation and by chemical reactions (cross-linking).
Hot melt adhesives basically consists of a polymer, a tackifier, and a
number of functional additives such as antioxidants. These components
are mixed together to produce a hot melt adhesive. The hot melt adhesives
are applied on the substrates in the molten state, and the adhesive bond
formation takes place by rapid solidification of the melt.

15.2 Polymers Commonly Used for Hot Melt Adhesives


While there are many polymers that are used in adhesive applications,
the bulk is served by four types of polymers: styrenic block copolymers
(SBCs), ethylene vinyl acetate copolymers (EVAs), linear low-density
polyethylene, and atactic poly alpha olefins (APAO), reactive PU, and
polyamide.

15.2.1 Ethylene Vinyl Acetate Copolymers


Ethylene copolymers, of which EVA is particularly predominant, are the
most commonly used in HMA formulations [1]. These copolymers are
characterized by their polar (ester) functionality that provides excellent
adhesion to more polar substrates such as paper, wood and metal.
EVA is a copolymer made from two different monomer types: ethylene
and vinyl acetate. The chemical structure of EVA is given in Figure 15.1.
The polymerization of EVA corresponds closely to the polymerization
of low-density poly(ethylene).

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (347–365) © 2019
Scrivener Publishing LLC

347
348 Adhesives for Wood and Lignocellulosic Materials

CH2-CH2 CH2-CH
n
O

C=O

CH3

Figure 15.1 Ethylene vinyl acetate copolymer.

Ethylene and vinyl acetate are copolymerized by free radical polym-


erization. The options are solution polymerization, suspension polymer-
ization, emulsion polymerization, and bulk polymerization. But solution
polymerization is preferred [2]. A method of either continuous type or
batch type may be employed. Methanol is generally used as the solvent.
Commonly used catalysts are 2,2 azobisisobutyronitrle, or organic perox-
ides. 2,2 -Azobis(4-methoxy-2,4-dimethyl valeronitrile) is the most pre-
ferred catalyst [3]. To stop the polymerization reaction, a polymerization
inhibitor is added to the reaction mixture.
The properties of EVA copolymer can determined by three structural
attributes, namely, the weight percent of vinyl acetate, molecular weight,
and distribution and molecular branching. Vinyl acetate content has two
basic influences on the properties of EVA copolymers.

(1) To disrupt the crystalline regions due to the polyethylene


segments of the copolymer [4].
(2) High vinyl acetate content tends to make the polymer
softer and more rubbery, while low vinyl acetate content
tends to make the polymer harder and more crystalline. If
vinyl acetate exceeds 45%, the crystallinity of the EVA will
be lost [5].

The composition of a hot melt adhesive usually includes a thermoplas-


tic polymer backbone and a diluent system. The diluent system in a hot
melt adhesive can include materials such as wax, tackifier, plasticizer, and
antioxidant.
By adding these materials accordingly, the properties of hot melt adhe-
sives can be modified and customized to a specific end usage.
Generally, hot melt adhesives based on EVA resins contain a vinyl acetate
concentration of 18–40%. A higher vinyl acetate content provides better
adhesion to polar substrates such as aluminum and steel, while the lower
vinyl acetate copolymers are often used for bonding low-energy surfaces.
Hot Melt Adhesives 349

The materials with high vinyl acetate concentration exhibit reduced crys-
tallinity and increased polarity. At about 50% vinyl acetate content, all
crystallinity is lost.
The properties required for specific end uses of EVA hot melt adhe-
sives depend on the degree of crystallinity of EVA copolymers. Hence,
the degree of crystallinity is responsible for versatility of these products
[4]. The common grades for EVA copolymer have vinyl acetate contents
ranging from 1% to 50% by weight, depending on the application [6].
Acetoxy groups present in the EVA confer polar characteristics on the
hot melt adhesives. As the vinyl acetate content increases, the polarity of
the EVA also increases. This is the second effect of vinyl acetate group
in the EVA-based hot melt adhesives. EVA-based hot melt adhesives are
therefore widely used in various substrates such as paper, wood, plastics,
rubbers, and metals [5].
The domain structure of EVA copolymers consists of stiff and partially
crystalline polyethylene blocks, and flexible, soft and polar amorphous
vinyl acetate blocks [7]. Thus, EVA polymers, like the SBCs, also exhibit
a two-phase morphology. The more crystalline polyethylene segments
segregate upon cooling from the melt from the amorphous domains
formed by the ester-containing segments. The ultimate crystallinity of the
polymer is also affected by branching formed during their free-radical-
catalyzed polymerization. The crystalline polyethylene domains deter-
mine the ultimate melting point of the polymer and contribute to the
hardness and creep resistance of the adhesive. Because they can cocrys-
tallize with the polyethylene regions, waxes are especially effective in
EVAs in enhancing the attributes (e.g., melting point) of these domains.
The amorphous regions are responsible for specific adhesion, tack, and
flexibility.
An example of a general-purpose EVA-based adhesive is shown in
Table 15.1.

Table 15.1 General-purpose EVA-based adhesive.


Components Parts by weight
Ethylene vinyl acetate (Elvax 220, DuPont) 30
Tackifier, polyterpene resin 50
Plasticizer, microcrystalline wax 20
Antioxidant 0.5–1.0
350 Adhesives for Wood and Lignocellulosic Materials

15.2.2 Styrenic Block Copolymers


SBCs are important components in a wide range of hot melt adhesives. As
the name implies, SBC is a copolymer formed when the two monomers,
namely, styrene and butadiene, cluster together and form “blocks” of
repeating units. SBCs are the largest volume and one of the lowest-priced
commercial thermoplastic elastomers. The styrene end blocks are hard
polymers that add strength to an HMA, but do not contribute to the tack
and adhesion properties. On the other hand, the polyisoprene (SIS), poly-
butadiene (SBS), ethylene–butene copolymer (SEBS), and others are the
midblocks. The midblock combined with tackifiers and plasticizers pro-
duce the adhesive properties.
Thus, three segments, styrene–polyolefin–styrene, are needed to attain
the unique properties of this class of polymers. The unique properties of
SBCs are the result of the two-phase morphology they are transformed
into in the solid state. Polystyrene segments are immiscible with poly-
isoprene and polybutadiene and are connected by the rubbery midblock
segments as shown in Figure 15.2 to form what is called thermoplastic
elastomer. The idea of thermoplastic elastomer comes from the concept
of reversible cross-link. Thus, the polystyrene clusters act as cross-links
for the polybutadiene blocks in the solid state. The cross-links however
break up when the SBS is heated, so it can be processed and recycled like
a non-cross-linked polymer.
However, it has to be kept in mind that the double bonds are present
in the midblock structure and are susceptible to thermal and oxidative
degradation. By selective hydrogenation of the midblock, the copolymers
become substantially more stable. These so-called SEBS block copolymers
have much improved thermal and weathering resistance [8].

polystyrene
domain

elastomeric
phase

Figure 15.2 Two-phase morphology of SBC.


Hot Melt Adhesives 351

This physical cross-link system imparts strength and flexibility to the


thermoplastic elastomer. The midblocks contribute flexibility to the adhe-
sive. It is also the region compatible with tackifiers and provides tack and
adhesion.

15.3 Polyureathane Reactive Hot Melt Adhesives


Reactive urethanes are the newest and fastest-growing category of hot melt
adhesives meant for high-impact applications. The conventional hot melt
adhesives are not chemically reactive and therefore do not develop sufficient
strength, heat, and chemical resistance required in certain applications.
In order to fulfill these properties, the hot melt adhesives should be
capable of undergoing cross-linking reaction, which will result in the for-
mation of durable bonds. Consequently, such reactive hot melt adhesives
can acquire characteristics significant enough to be used in a number of
performance-driven applications.
Reactive polyurethane hot melt adhesives fall into this category. An
isocyanate terminated polyurethane is first produced by the step-growth
polymerization of a polyol and excess of polyisocyanate to control the
molecular weight and to ensure the availability of moisture curable isocya-
nate end groups in the final product.
The physical properties that are key to performance are the melt vis-
cosity, rheology, and melt stability. The atmospheric moisture as well as
moisture present in the substrates can diffuse into the surface and the bulk
of the adhesive and can cause chemical changes to occur [9] [see reactions
of isocyanates with water in Section 7.3 (“Reactions of isocyanates”)]. The
essential reactions are as follows:

1. Reaction of isocyanate with moisture to form amine and


carbon dioxide.
2. Chain extension by reaction of isocyanate with amine
formed to produce substituted urea.
3. Branching reactions of isocyanate with urethane to produce
allophanate and with urea to produce biuret.

The above sequence of reactions leads to cross-linking. This in turn


confers superior properties on the bonded assembly.
A wide range of hot melt PU adhesives is possible given a wide range of
raw materials that are currently available, namely, polyether polyols such
352 Adhesives for Wood and Lignocellulosic Materials

as poly(propylene glycol) in a range of molecular weights, a wide range of


polyester polyols, and a variety of isocyanates.
Because of the cross-linked structure in the bond line, the reactive PU hot
melt adhesives develop good temperature and chemical resistance, a high
degree of degree of stiffness, as well as toughness. The toughness of polyure-
thane is attributed to the flexibility of the long-chain polyols called “soft seg-
ments”. The stiffness of the polyurethane is attributed to the isocyanates with
relatively short chains and high Tg. This domain is called “hard segments”.
Applied as a drop or a thin bead, high-viscosity reactive urethanes set
quickly, becoming structurally rigid in minutes. The secondary moisture
cure continues for up to 24 h.
Since reactive urethane adhesives are heat resistant, they can be used in
heat-sensitive end-use applications.

15.4 Silane Reactive Hot Melt Adhesives


The silane reactive hot melt adhesive composition comprises one or more
silane-modified polymers. The silane-modified polymer has an organic
backbone, bearing one or more terminal or pendant silane or alkoxylated
silane groups. The silane groups are hydrolyzed by water to silanol groups,
which in turn can condense with one another to form polysiloxane or with
reactive species on the adherent surfaces. The silane-modified polymer
may be prepared with one or more variety of polymer backbones such as
polyurethane, polyether, polyester, polyolefin, polycaprolactone, polyacry-
late, polybutadiene, polycarbonate, polyacetal, polyester amide, polythio-
ether, polyolefin, and the like.

15.5 Polyamide Hot Melt Adhesives


Polyamides are condensation polymers, produced by reacting dibasic acids
with diamines. Examples of dibasic acids and diamines employed for the
production of polyamides are given below:

Dibasic Acids:
Dimer acid (dimerized fatty acids)
Dodecanedioic acid
Sebacic acid
Azelaic acid
Adipic acid
Hot Melt Adhesives 353

Diamines:
Ethylene diamine
Hexamethylene diamine
Diethylene triamine
Triethylene tetramine

The thermoplastic polyamide resins are amides prepared from dimer


acids and low-molecular-weight aliphatic amines. Dimer acids are the
Diels Alder adducts of an unsaturated aliphatic monocarboxylic acid such
as linolenic or linoleic acid (Figure 15.3).
The amines employed for the preparation of the amides are low-
molecular-weight aliphatic amines containing two terminal -NH2 groups
such as ethylenediamine or hexamethylenediamine.
Thermoplastic polyamides derived from dimer acid and an aliphatic
diamine are shown in Figure 15.4. Polyamides are used as hot melt

(CH2)7-COOH

CH
HC CH (CH2)7 COOH
+
CH CH2 CH=CH (CH2)4 CH3
HC
CH Linoleic acid

(CH2)5 CH3

Isomerized linoleic acid

(CH2)7-COOH

CH
HC CH (CH2)7 COOH

CH CH2 CH=CH (CH2)4 CH3


HC
CH3

(CH2)5 CH3

Dimer acid

Figure 15.3 Formation of dimer acid.


354 Adhesives for Wood and Lignocellulosic Materials

(CH2)7-COOH
+ 2 H2N-CH2-CH2-NH2
CH
Ethylenediamine
HC CH (CH2)7 COOH

CH CH2 CH=CH (CH2)4 CH3


HC
CH3

(CH2)5 CH3

Dimer acid
(CH2)7-CO NH-CH2-CH2-NH

CH
HC CH (CH2)7 CO NH-CH2-CH2-NH

CH CH2 CH=CH (CH2)4 CH3


HC
CH3

(CH2)5 CH3

Dimer acid polyamide

Figure 15.4 Dimer acid polyamide.

adhesives, particularly for applications that require good adhesion and


high tensile strength of adhesive bonds.
The thermoplastic polyamide adhesives are available commercially in
three general molecular weight classes. The lowest molecular weight (for
inexpensive equipment), the intermediate molecular weight (for equip-
ment suitable for handling higher melt viscosities), and high molecular
weight (for high-strength applications) [10].
Polyamides are excellent adhesives and can be used under severe envi-
ronmental conditions such as exposure to solvents, gas, or oil. They are
also suitable for applications up to 160°C. They possess superior thermal
stability, good chemical resistance, and strength. They are tough polymers
and are excellent for electrical potting applications. Polyamide hot melts
offer good adhesion to a wide range of substrates including metal, wood,
vinyl, ABS, and treated polypropylene and polyethylene.
Hot Melt Adhesives 355

15.6 Amorphous Polyolefin (APO/APAO)


Hot Melt Adhesives
The preparation of a high-performance and low-cost hot melt adhesive
based on atactic polypropylene and a minor proportion of different poly-
mers such as isotactic polypropylene and polyethylene has been reported
[11]. The resulting hot melt adhesives have excellent properties for bond-
ing materials, particularly paper. They have light color, making them suit-
able for applications where clarity is important.
Atactic (essentially non-crystalline) polypropylene is formed during the
stereospecific polymerization of propylene using the Ziegler–Natta cata-
lyst system. The polymerization takes place in the presence of a catalyst
comprising a coordination complex of a transition metal halide with an
organometallic compound. Atactic polypropylene usually represents from
about 5% to 15% by weight of the polymerization product, the remainder
being isotactic (essentially crystalline) poly propylene. The solid atactic
polypropylene suitable for the composition of the adhesive has a molecular
weight of 15,000 to 60,000 and more particularly from 16,000 to 20,000. It
is soluble in boiling pentane, hexane, heptanes, and other hydrocarbons.
They have excellent chemical resistance to polar solvents, acids, bases,
esters, and alcohols, but only moderate heat resistance and poor chemical
resistance against nonpolar solvents like alkanes, ethers, and oils. When
compared with EVA and polyamide hot melt adhesives, polyolefins have
extended open times for positioning of parts. They also have lower melt
viscosity and slower set times compared to EVAs. Some APOs can be used
without any additives, but often they are compounded with tackifiers,
waxes, and plasticizers (mineral oil, polybutene oil) [8].
Zhou (2006) reported novel polyolefin hot melt adhesives based on the
blend of amorphous/atactic polypropylene (or copolymers) and crystalline/
isotactic polypropylene as well other additives [11]. The newly formulated
polyolefin-based adhesive made from the PP blends was successfully used as
a multifunctional hot melt adhesive for construction.
Polyolefin hot melts are a unique combination of base resins and tackifi-
ers. This hot melt technology provides superior adhesion to polypropylene,
a good barrier against moisture and water vapor, and excellent chemical
resistance against polar solvents and solutions including acids, bases, and
alcohols.
356 Adhesives for Wood and Lignocellulosic Materials

15.7 Tackifiers
Tackifiers are chemical compounds used in formulating adhesives to
increase the tack, the stickiness of the surface of the adhesive. They have a
crucial role in providing initial adhesion. They are usually low-molecular-
weight compounds with high glass transition temperature.
Tackifiers function as polymer modifiers. Hence, they should pos-
sess good compatibility with the main polymer for good performance.
Compatibility among two or more materials is governed by their molecular
weights, molecular weight distributions, and solubility parameters (com-
positions). Low molecular weights, narrow molecular weight distributions,
and close solubility parameters result in good tackifier–polymer compati-
bility. The Hildebrand solubility parameter gives an estimate of the degree
of interaction between tackifier and polymer. Materials with similar values
are likely to be miscible and have favorable viscoelastic properties [12].

15.7.1 Aromatic Hydrocarbon Resins


Hydrocarbon resins are amorphous thermoplastic polymers produced by
polymerization of unsaturated hydrocarbons. The feedstock are various
by-products of naphtha crackers, resins from coal tar commonly called
coumarone-indene resin. These resins have typically a low molecular
weight ranging from about 400 to 5000 g/mol. The three main types are C5
aliphatic, C9 aromatic, and DCPD cycloaliphatic resins. These resins are
produced by catalytic solution polymerization of aromatic crude streams
containing indene as the principal polymerizable monomer along with
varying minor percentage of styrene, methyl indenes, methyl styrenes,
coumarone, and dicyclopentadiene. Aromatic resins can be obtained in
various softening points from about 10°C to over 150°C and in colors rang-
ing from pale straw to dark amber. The iodine number generally taken
as a measure of reactivity or unsaturation is relatively low in these resins
ranging from 30 to 60 [13].

15.7.2 Aliphatic Hydrocarbon Resins


Aliphatic hydrocarbon tackifier resins are thermoplastic resins that pro-
mote adhesion and tack in hot melt adhesives. Hydrocarbon tackifier
resins are selected on the basis of the softening point, molecular weight,
molecular weight distribution, and composition of the tackifier. Aliphatic
hydrocarbon resins (C5 resins) are made from C-5 piperylene (cis- and
Hot Melt Adhesives 357

trans-piperylene) and its derivatives. The most important ones are cis/
trans 1,3-pentadienes, 2-methyl-2-butene, cyclopentene, cyclopenta-
diene, and dicyclopentadiene (see below). The liquid C5 feedstock can
be polymerized to a hard resin using a Lewis acid catalyst and carefully
selecting temperature and pressure to obtain the desired softening point
and molecular weight. The structure of the polymerized resin is difficult
to characterize since the various isomers of the feedstock will combine
unpredictably [13]. Hydrogenation is used after a resin is polymerized
to decrease its color and improve its stability towards heat, oxygen, and
ultraviolet light.

15.7.3 Mixed Aliphatic and Aromatic Resins


C5 aliphatic and C9 aromatic resins can be modified by mixing the two feed
streams together at a carefully chosen ratio to produce hybrid polymers.
This ratio determines the aliphatic/aromatic balance of the resin, which is
an essential determinant of the resin’s compatibility [14]. The polymeriza-
tion is carried out using either AlCl3 or BF3 as required. The MMAP* and
DACP** cloud points are the method used to characterize C5/C9 resins.
As the cloud points decrease, the compatibility with polar and aromatic
polymers increases

[*MMAP is 1:2 mixture of methylcyclohexane and aniline


**DACP is 1:1 mixture of xylene and di-acetone alcohol
(4-hydroxy-4-methyl-2-pentanone)

Cloud point determination is easy to do and involves weighing a stan-


dard amount of resin that is dissolved in the solvent at high temperature.
When homogeneous, it is allowed to cool, with mixing. The temperature
at which the resin begins to form a separate phase is defined as the cloud
point.]
A study explored the use of dynamic mechanical analysis to qualitatively
show the effects of resin aromaticity, as measured by the modified mixed
aniline point (MMAP), on adhesive properties [15]. Another presentation
described the use of MMAP and a diacetone alcohol cloud point (DACP)
as predictors of tackifier utility in different elastomer systems [16].
These resins are the result of research efforts by various manufactur-
ers to fine-tune the compatibility and tackifying characteristics of heir C-5
resins to produce products with outstanding performance of the hot melt
adhesives [13].
358 Adhesives for Wood and Lignocellulosic Materials

15.7.4 Terpene Resins


Terpene-based resins are important tackifiers, besides rosin esters, derived
from renewable resources. The terpenes used in the tackifier industry are
obtained from turpentine, which can be either gum turpentine or crude
sulfate turpentine (CST) obtained from tall oil during the Kraft pulping of
pine wood. The two major terpenes obtained from turpentine are α-pinene
and β-pinene (Figure 15.5). Another very commonly used terpene for tack-
ifiers that is not derived from turpentine is d-limonene, which is obtained
from citrus sources, i.e., orange peels.
Terpene resins are primarily synthesized by a cationic polymerization
process where a suitable solvent and AlCl3 or other Lewis acid catalysts are
employed. The three major classes of terpene resins in the tackifier indus-
try are polyterpene resins, phenol-modified polyterpene resins (terpene–
phenol resins), and styrene-modified polyterpene resins (styrenated ter-
pene resins) (Figure 15.6). The actual monomers are usually one or a
combination of the terpenes, α-pinene, β-pinene, and dipentene or lim-
onene [17].
Terpene resins are light in color and can be obtained in softening points
10°C to 140°C and molecular weights from about 300 to almost 2000.
Generally speaking, at the equivalent softening points, β-pinene resins
have the highest molecular weights [18].

α−Pinene β−Pinene Dipentene

Figure 15.5 Terpenes for producing tackifiers.

Styrene
Polyterpene Styrenated Terpene Resin

α−Pinene
Phenol

Terpene-Phenol Resin

Figure 15.6 Three major classes of terpene resins.


Hot Melt Adhesives 359

15.7.5 Terpene–Phenol Resins


The terpene-phenolic tackifier resins are produced by an alkylating reac-
tion of a compound containing at least two phenolic groups with a terpene
in the presence of a Friedel Crafts catalyst. The phenolic compound should
have two phenolic groups and should have at least one ortho or para posi-
tion free for alkylation by a terpene.
A method for making a phenol–terpene resin by reacting phenol or alkyl
phenol and a terpene in the presence of a Friedel Crafts catalyst and an aro-
matic, naphthenic, or paraffinic hydrocarbon solvent has been reported by
Gonzenbach [19].
Wang et al. [20] describe the preparation of terpene–phenol–
aldehyde resin, a copolymer product obtained by cationic copolymeriza-
tion of monoterpene with phenol and aldehyde, as a new modified pheno-
lic aldehyde resin. It is a transparent, light yellow material that is soluble
in most solvents and is compatible with rubber, rosin, and others. The ter-
penes usually used for the synthesis of terpene–phenol–aldehyde resin are
α-pinene, β-pinene, limonene, carene, camphene, and others. The phenols
used are phenol, 3;4-dimethylphenol, 4-methylphenol, and others. The
aldehyde used is mainly formaldehyde. There are two methods for syn-
thesizing the resin. With the first method, phenol is condensed with an
aldehyde in the presence of an acid catalyst and is then alkylated with the
terpene in the presence of Lewis acid or protonic acid. With the second
method, phenol is first polymerized with the terpene in the presence of an
acid catalyst and is then condensed with the formaldehyde [20].

15.7.6 Rosin and Rosin Derivatives


The use of wood resins in connection with the construction of early sailing
ships led to the general term naval stores for rosin, turpentine, and tall
oil, which are the most important group of materials today with regard to
commercial production. These naval stores are obtained either by tapping
living pine trees (gum resin, oleo resin), by solvent extraction of stumps
of coniferous trees (wood rosin), or as a by-product of pine and softwood
kraft pulping (tall oil rosin) [21].
Tall oil rosin, the third type of rosin, is obtained from the tall oil, which
is a by-product of the paper industry. During the pulping operation, wood
chips are cooked under heat and pressure in the alkaline medium. The pulp
is removed by filtration and washing, and the liquid portion is concen-
trated, resulting in a precipitate that is removed and acidified to give crude
360 Adhesives for Wood and Lignocellulosic Materials

tall oil. Crude tall oil consists mainly of fatty acids and rosin acids that are
separated by distillation.
While the production of oleo resins is generally stagnant, the produc-
tion of tall oil and sulfate turpentine in connection with the kraft pulping
of pines and other coniferous woods remains the dominant source of naval
stores [21].
Rosin from any of the above sources is a mixture of organic acids called
rosin acids. Minor components consist of rosin esters and anhydrides,
unsaponifiable matter, and fatty acids.
Rosin acids can be divided into two types. These are the abietic acid type
and pimaric acid type (Figure 15.7). Unmodified rosin consists primarily
of abietic acid type. The figure shows some of the more common members
of these two types.
It can be seen that the difference in the two types is the substitution
pattern at the C13 position. Abietic acid types have an isopropyl group at
the C13 and pimaric acid types have instead a methyl and a vinyl group.
Rosin by itself is generally unsuitable for use in modern adhesive sys-
tems because it is prone to oxidation and further reactions as well as
crystallization.

Abietic acid type

COOH COOH COOH


COOH
Abietic acid Neo-abietic acid Dehydroabietic acid Dihydroabietic acid

Pimeric acid type

COOH
COOH
Palustric acid Isopimeric cid

Figure 15.7 Rosin acids.


Hot Melt Adhesives 361

Most rosins for adhesive use are modified to give rosin derivatives
through disproportionation, polymerization, hydrogenation, and esterifi-
cation reactions.
Disproportionation takes place when rosin is heated to about 270°C.
The reaction may be accelerated by use of various catalysts such as plati-
num or iodine. The result is a hydrogen exchange between the rosin acids.
Accordingly, two abietic acids are converted into dihydroabietic acid and
dehydroabietic acid.
Polymerization is carried out by treating a solution of rosin in an
inert solvent with an acid catalyst such as boron trifluoride or sulfuric
acid. Removal of solvent and the catalyst results in a product composed
largely of rosin dimer and having a higher softening point than unmod-
ified rosin. Hydrogenation using Raney nickel or a similar catalyst can
be tailored to give products that are only partially saturated such as
dihydroabietic acid or completely saturated such as tetrahydroabietic
acid [13].
Rosin acids react with a variety of alcohols to form esters. Because
of steric hindrance, however, the esterification requires higher tempera-
ture, but the resulting esters are quite stable and resistant to hydroly-
sis. Usually, polyhydric alcohols are used for esterification in order to
give products with higher softening points and higher molecular weight.
Ethylene glycol, glycerol, and pentaerythritol are the most common alco-
hols for the purpose. Rosin esters are compatible with a broad range of
polymers, including high and low vinyl acetate EVA, acrylics, SBR, SIS,
and SBS. This characteristic enables formulation flexibility in various hot
melt, water-based, and solvent-based adhesive applications. Solid rosin
esters can be formulated into SBC- or EVA-based adhesives for packag-
ing, PSA, and woodworking applications. In addition to these end uses,
liquid rosin esters can also be used in water-based acrylic for construc-
tion adhesives [12].

15.8 Antioxidants [22]


Adhesives and their raw materials react like other organic materials with
oxygen in a process called “autoxidation”. Autoxidation is initiated by heat,
light, mechanical stress, catalyst residues, or impurities.
Free radicals are formed and react in the presence of oxygen to form
peroxy radicals, which further react with other hydrogen donors present,
leading to hydroperoxides (ROOH).
362 Adhesives for Wood and Lignocellulosic Materials

All these reactions ultimately lead to a change in chemical composition


and such polymer properties as molecular weight, which has immediate
impact on polymer degradation affecting the service life of an adhesive.
Some results of adhesive degradation are as follows:

• discoloration
• viscosity changes
• char formation
• cracking and
• loss of adhesion

Antioxidants are used in a variety of adhesive formulations to protect


against degradation caused by reaction with atmospheric oxygen. The
introduction and type of antioxidant will depend on the nature of the base
polymer, the processing parameters, and the end-use application. The anti-
oxidant or stabilizer maintains viscosity, color, and physical properties and
prevents thermal degradation.
Excessive oxidation generally results in undesirable changes in the
adhesive’s mechanical, aesthetic, or bonding properties. Oxidation can
occur at all stages of an adhesive’s life from synthesis to final end use. It is
usually recognized at high processing temperatures such as during mixing,
compounding, or extrusion (in the case of hot melt adhesives). However,
oxidation can also occur at relatively low temperatures including ambient
storage and also on exposure to UV light.

15.8.1 Oxidation-Sensitive Components in Hot Melts


Adhesive components especially susceptible to oxidation include base syn-
thetic polymers such as ethylene vinyl acetate and styrene block copoly-
mers, polyolefins, polyamides, and polyurethane. Hydrocarbon additives,
such as tackifiers and waxes, are also vulnerable to oxidation and can actu-
ally contribute to the oxidation of the base polymer. Metallic and other
impurities in the adhesive can accelerate the oxidation process. Depending
on the aging environment, most adhesives can benefit from antioxidants.

15.8.2 Antioxidants Used in Hot Melts


Hindered phenols, amines, phosphites, and thioester are commonly used
stabilizers for hot melt adhesives. The chemical types of common antioxi-
dants that are most often used in hot melt adhesive applications are shown
in Table 15.2 [22]. The hot melt adhesive composition includes from about
Hot Melt Adhesives 363

Table 15.2 Common antioxidants by chemical type with major resin applications.
Types Common resin applications Comments
Amine Rubber, some pigmented Arylamines tend to
polymers, and discolor and cause
polyurethane polyols staining
Phenolic Polyolefins, styrenics, and Phenolics are generally
most engineering resins stain resistant
and include
simple phenolics
(BHT), various
polyphenolics, and
bisphenolics
Organo-phosphite Polyolefins, styrenics, and Phosphites can improve
most engineering resins color stability and
property retention,
but can be corrosive
if hydrolyzed
Thioester Polyolefins and styrenics The major disadvantage
with thioesters is
their odor, which is
transferred to the
host polymer

0.1% to about 1.0% by weight antioxidant. There are numerous antioxi-


dants and blends of antioxidants that have proven beneficial for various
types of adhesives.

15.9 Plasticizers
Apart from the basic polymers and the tackifiers, plasticizers are the other
most common additive in hot melt adhesive formulations. The purpose of
adding the plasticizers is to impart flexibility and toughness to the product.
When the base resin is excessively stiff, it is often blended with an elasto-
meric to obtain a tough material with improved energy dissipation and
reduced glass transition temperature. These materials act opposite to the
tackifiers in that they decrease Tg. The plasticizer must be completely solu-
ble in the base resin and be sufficiently nonvolatile. The addition of a plas-
ticizer promotes wetting and reduces the melt viscosity of the formulation.
364 Adhesives for Wood and Lignocellulosic Materials

15.10 Mineral Oil and Wax


Mineral oil and wax are commonly employed as diluents [23]. When added
to the matrix polymer, wax and oils can negatively affect the adhesion
properties because of the shrinkage of the adhesive. Waxes are often used
in hot melt formulations to lower surface tension and decrease melt vis-
cosity. Certain waxes such as microcrystalline waxes also reinforce the hot
melt by forming crystallites that resist deformation under load. These are
used in formulations that require a relatively high degree of creep strength.
They are performance-enhancing additives.

References
1. Fink, J.K., Ethylene vinyl acetate copolymers, in: Handbook of Engineering
and Speciality Thermoplastics: Polyolefins and Styrenics, Scrivener Publishing,
New York, 2010.
2. T. Kawahara and T. Tuboi, Method for producing saponified ethylene vinyl
acetate copolymer. US Patent 7,041,731, 2006.
3. T. Kawahara and M. Takai, Method for manufacturing ethylene-vinyl acetate
copolymer and apparatus for manufacturing the same. US Patent 6,831,139,
2004.
4. Henderson, A.M., Ethylene-vinyl acetate (EVA) copolymers: A general review.
IEEE Electr. Insul. Mag., 9, 1, 30–38, 1993.
5. Ghani bin, M.S.H., Development of hot melt ethylene vinyl acetate (EVA)
adhesive for packaging industry. PhD Thesis, Universiti Malaysia Pahang,
Malaysia, 2015.
6. Arsac, A., Carrot, C., Guillet, J., Rheological characterization of ethylene
vinyl acetate copolymers. J. Appl. Polym. Sci., 74, 2625–2630, 1999.
7. Park, S., Yim, C., Lee, B.H., Choe, S., Properties of the blends of ethylene-
vinyl acetate and ethylene-α-olefins copolymers. Macromol. Res., 13, 243–
252, 2005.
8. Polymer Properties Database, 2015. http://polymerdatabase.com/home.html.
9. Slark, A.T., Recent developments in acrylic/polyurethane reactive hot melt
(rum) adhesives. Proceedings, Hot-Melt Adhesives, Society for Adhesion and
Adhesives, One-day Symposium, Society of Chemical Industry, Belgrave
Square, London, 5th December 2002, 2002.
10. P.R. Lakshmanan and B.J. Monachino, Thermoplastic polyamide hot melt
adhesive composition. US Patent 4,359,556, 1982.
11. Zhou, P., Development of Polypropylene-Based Polyolefin Hot Melt
Adhesive for Personal Care Product Applications. Ip.com.disclosure num-
ber: IPCOM000141004D, 2006.
Hot Melt Adhesives 365

12. Kraton Tackiyer for Adhesives. Company brochure. http://www.kraton.com/


products/pdf/Adhesive%20Brochure.pdf.
13. Satas, D., Handbook of Pressure Sensitive Adhesive Technology, Van Nostrand
Reinhold, New York, 1982.
14. Schlademan, J.A., Compatibility and performance of hydrocarbon tackifier
resins, 1999. https://www.pstc.org/files/public/Schlademan.pdf.
15. Napolitano, M.J., TAPPI Hot Melt Symposium, Notes, pp. 275–287, 1998.
16. Donker, C., PSTC Annual Technical Seminar, Proceedings, pp. 149–164, 2001.
17. Abhay Deshpande Terpene Resins in Pressure Sensitive Adhesives. Arizona
Chemical Company. https://www.pstc.org/files/Deshpande_Abhay.pdf.
18. Mathers, R.T. and Meier, M.A.R., Green Polymerization Methods, Wiley-
VCH, Weinheim, Germany, 2011.
19. C.T. Gonzenbach, Phenol–terpene cyclic polyolefin polymer. US Patent
3,383,362, 1968.
20. Wang, S.-F., Furuno, T., Cheng, Z., Studies on the synthesis and properties of
terpene–phenol–aldehyde resin with a high softening point. J. Wood Sci., 46,
143–148, 2000.
21. Fengel, D. and Wegener, G., Wood, Chemistry, Ultrastructure, Reactions,
Walter de Gruyter, Berlin 1989.
22. Petrie, E.M., Special Chem. https://adhesives.specialchem.com/selection-guide/
hot-melt-adhesive-formulation-for-hygiene-applications/antioxidants.
23. Merrill, N., Wax for hot melt adhesive applications. Adhes. Sealants Ind., 14,
4, 37–39, 2007.
Part B
POLYMER MATRIX MATERIALS
FOR BIOFIBER COMPOSITES

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (367–388) © 2019
Scrivener Publishing LLC
16
Modification of Natural Fibers
and Polymeric Matrices

16.1 Introduction
Recently, the incorporation of lignocellulosic materials as reinforcing
agents or as fillers in polymer matrix composites has received increased
attention. Although natural fibers have a number of advantages over glass
fibers, the strong polar character of their surface makes them incompati-
ble with strongly nonpolar thermoplastic matrices. Consequently, natural
fibers do not disperse easily in hydrophobic thermoplastic polymers. This
is due to hydrophilicity and strong intermolecular hydrogen bonding pres-
ent in natural fibers. As a result, biofibers tend to agglomerate during the
compounding process and do not mix well with synthetic polymer matri-
ces. The low compatibility and lack of interfacial adhesion between biofi-
bers and the polymer matrix result in composites having low mechanical
properties [1–4].
Hence, fiber pretreatment is necessary to enhance compatibility between
the natural lignocellulosic fibers and the polymer matrix materials, a con-
dition imperative for the enhanced adhesion between the fiber and the
matrix. The efficiency of a fiber-reinforced composite depends on the
fiber–matrix interface and the ability to transfer stress from the matrix to
the fiber. This stress transfer efficiency plays a dominant role in determin-
ing the mechanical properties of the composite and also in the material’s
ability to withstand environmentally severe conditions [5].
Lu et al. identified over 40 coupling agents that are used in wood–fiber–
polymer composites (WFPC) [6]. Accordingly, the coupling agents can
be classified into organic, inorganic, and organo-inorganic substances.
Organic coupling agents include isocyanates, anhydrides, amides, imides,
acrylates, chlorotriazines, epoxides, organic acids, monomers, polymers,
and copolymers. Only a few inorganic coupling agents, such as silicates,
are used in WFPC. Organic–inorganic agents include silanes and titanates.

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (369–388) © 2019
Scrivener Publishing LLC

369
370 Adhesives for Wood and Lignocellulosic Materials

Some functionalized agents, such as maleated polypropylene (PP)


(MAPP), maleated styrene–ethylene/butylene–styrene (SEBS-MA), and
styrene–maleic anhydride (SMA), also act as compatibilizers in WFPC
[7–9].
Organic coupling agents contain normally mono-, bi-, or multifunc-
tional groups in their molecular structure such as isocyanates, chlorine,
epoxides, acrylates, acids, etc. The results can be of two types:

(1) Reaction between these functional groups with the


hydroxyl groups of the biofiber surfaces can result in a
reduction of the surface energy of biofibers. The biofibers
become hydrophobic and therefore compatible with the
polymer matrices.
(2) Alternately, organic coupling agents can modify the poly-
mer matrix by graft copolymerization, thus resulting in
strong adhesion, or even cross-linking, at the interface.

Inorganic coupling agents possibly act as dispersing agents to counter-


act the surface polarity of the wood fibers and improve the compatibility
between wood fiber and matrix polymer.
Based on the facts above, there are two strategies for achieving fiber–
polymer compatibility. These are:

1. Fiber-based strategies
2. Matrix-based strategies

The fiber-based strategies are physical and chemical treatment of fibers


to make them more hydrophobic so that they are compatible with the poly-
mer matrices. Conversely, the strategy for the treatment of polymer matri-
ces relies on converting their surface to become hydrophilic to ensure their
favorable interaction with the natural fibers.

16.2 Strategies to Treat the Biofibers


for Compatibility
16.2.1 Physical Methods
Stretching, calendaring, and thermal treatments are some of the physical
methods employed to promote the affinity between the biofibers and the
polymer matrix materials. These methods do not change the chemical
Natural Fibers and Polymeric Matrices 371

composition of the fibers but can change the surface characteristics of the
fiber in such a manner as to facilitate their mechanical anchorage with the
matrix polymers [10, 11].
Other physical methods are discharge processes such as corona and
cold plasma. Corona treatment effects activation due to surface oxidation.
It is a process that changes the surface energy of the cellulose fibers and
increases the amount of aldehyde groups on their surfaces. To explain the
nature of any chemical change occurring on the wood surface, Sakata et al.
examined the dyeing of wood and its components with Schiff ’s reagents
[12]. The results showed a higher dyeing ability for corona‐treated samples
compared to untreated ones, indicating the increase of aldehyde groups by
corona treatment.
Electric discharge methods are also known to be very effective for
“non-active” polymer substrates such as polystyrene (PS), polyethylene
(PE), PP, etc. The electric discharge methods are successfully used for
cellulose–fiber modification to decrease the melt viscosity of cellulose–
PE composites and improve mechanical properties of cellulose–PP com-
posites [11, 13–15].

16.2.2 Steam Explosion Treatment


In the steam explosion process, lignocellulosic materials are subjected
to a high-pressure steam at high temperature, and pressure is suddenly
released, causing a shock followed by a sudden mechanical disruption of
the material by a violent discharge (explosion) of the material into a col-
lecting tank. This process has been applied to many lignocellulosic materi-
als including plant fibers, to enhance dispersibility in and adhesion to the
polymer matrix [16].

16.3 Chemical Methods


Chemical pretreatment of the natural fiber can eliminate limitations of
incompatibility of the natural fibers with the polymer matrix and enhance
the interaction between them. This results in an improved performance of
fiber-reinforced composites. Different surface treatment methods such as
mercerization (alkali treatment), isocyanate treatment, maleic anhydride
(MAH) treatment, acrylation, benzoylation, permanganate treatment,
acetylation, silane treatment, and peroxide treatment have been applied on
the fiber to improve the fiber–matrix adhesion and strength of the resulting
composites.
372 Adhesives for Wood and Lignocellulosic Materials

Many chemicals have been reported to modify natural fibers to


improve performance: anhydrides, acid chlorides, ketene carboxylic
acids, isocyanates, aldehydes, β-propiolactone, acrylonitrile, and epox-
ides. The treated natural fibers can be made into a composite, having
good dimensional stability and biological resistance, after reaction [17].

16.3.1 Mercerization
Mercerization is an alkali treatment widely used for modifying cellulose
fibers in cotton materials. Mercerization causes an increase of the amount
of cellulose II at the expense of cellulose I. Furthermore, because of this
treatment, fibers are reported to become more circular in cross section
[18]. Mercerization can be considered as a pretreatment for activation for
subsequent modification steps [19–22] or pretreatment for matrix-based
compatibilization strategies [14, 22].
It is possible to use mercerization on natural fibers to ensure improve-
ment of their tensile properties [23]. Due to alkali treatment, there is an
increase in the amount of amorphous cellulose at the expense of crystalline
cellulose. By this treatment, there is a removal of hydrogen bonding in the
network structure. The reaction that takes place during this treatment is
shown below [24].

Fiber-OH + NaOH → Fiber-O-Na+ + H20

Increase in the fiber surface roughness caused by mercerization improves


the adhesion and the potential for stress transfer through mechanical
interlocking at the fiber–matrix interface [25–27]. This, in turn, influences
mechanical properties of the composites not only via fiber–matrix interac-
tion but also via changing the intrinsic fiber properties themselves.
Aqueous solutions of NaOH are used for mercerization. Other alkalis
such as KOH and LiOH are also used. The alkali concentration influences
the degree of swelling and degree of lattice transformation into cellulose II
[28]. Alkali solution not only affects the cellulosic components inside the
plant fiber but also affects the non-cellulosic components such as hemicel-
luloses, lignin, and pectins [29].
Jacob et al. reported that maximum tensile strength resulted from the 4%
NaOH treatment at room temperature [30]. Mishra et al. found that del-
ignification of natural fibers occurs at high concentration of alkali, result-
ing in damage of the fibers [31]. As a result, the tensile strength of such
composites decreased drastically beyond certain optimum NaOH concen-
trations. Soleimani et al. found considerable improvements in mechanical
Natural Fibers and Polymeric Matrices 373

properties in the PP–flax composites when a combination of NaOH treat-


ment of the flax fibers is coupled with a subsequent matrix modification by
MA–PP coupling [32].

16.3.2 Acetylation of Natural Fibers


Acetylation has been a well-known method to confer dimensional stabil-
ity to solid wood since 1946. This method has been used for the surface
modification of biofibers for use in fiber-reinforced composites. During
acetylation, the hydroxyl groups (–OH) of the fiber constituents are con-
verted into acetyl groups (CH3CO), which, in turn, imparts hydrophobic
character to the surface of biofibers [33].
The reaction involves esterification of all the three hydroxyl groups of
anhydro-D-glucose when it is carried out in a homogeneous phase (i.e.,
when cellulose is dissolved). But in the case of fibers and wood, the reac-
tion is heterogeneous and a catalyst is required to accelerate the reaction
and also to achieve uniformity of the product. The catalysts employed are
sulfuric acid, pyridine, potassium and sodium acetate, etc. A more recent
acetylation method is based on isopropenyl acetate. Acetic acid is the
by-product. Acetylation imparts hydrophobic character to the biofibers.
The hydrophobic biofiber becomes then compatible with the thermoplastic
polymer matrix material [19]. The reaction is depicted in Figure 16.1.
The reaction lowers the surface energy of the fiber and makes the sur-
face nonpolar so that the fibers become compatible with the plastic matrix
[34, 35].

16.3.3 Silane Coupling Agents


Silane coupling agents were initially developed in the 1940s for use in glass
fiber-reinforced polymers for many years. These coupling agents have also
been used with biofibers as compatibilization strategy. A large variety of

Bio-Fiber
O=

O=

CH3 C HO
O=

C OH
O
CH3 C O + CH3
CH3 C HO
=

Acetylated fiber Acetic acid


O

Figure 16.1 Acetylation of lignocellulosic fibers.


374 Adhesives for Wood and Lignocellulosic Materials

Si

R'O OR'
OR'
R = Alkyl or aryl functional groups such as Aminopropyl,
Mthacryloxypropyl, Glycidyloxypropyl etc
OR' = Methoxy, ethoxy, acetoxy etc

Figure 16.2 Structure of silane coupling agent.

silanes are now available for different applications. All these molecules
share the same basic structure (Figure 16.2).
The alkoxy group OR is hydrolyzed during the fiber pretreatment, lead-
ing to the formation of silanols. The silanol then reacts with the hydroxyl
group of the fiber, forming ether bonds (Si–O–C). Furthermore, the silane
molecules contain the functional group R, which is compatible with the
polymer matrix. The appropriate R groups are chosen for different polymer
matrices. By such mechanisms, composite mechanical properties could be
improved due to enhanced fiber–matrix interaction.
Further better dispersion can be ensured for the fiber/filler throughout
the matrix phase [19, 36].
The fiber/filler pretreatment conditions are adapted to suit the silane
chemical structure. Depending on this, the silanes are either applied in
organic-based solvents, such as acetone/acetic acid, carbon tetrachloride,
methanol/water, ethanol/ water, or ethanol, or water-based solvents [19].

16.3.4 Benzoylation Treatment


Treatment of biofibers with benzoyl chloride decreases the hydrophilic
character of the natural fibers and improves the interaction with hydro-
phobic PS. Benzoylation of fiber improves fiber matrix adhesion, thereby
considerably increasing the strength of composites, decreasing their water
absorption and improving their thermal stability [37, 38].

16.3.5 Acrylation of Natural Fibers


Acrylation of natural fibers can be carried out by treatment with acrylic
acid in the presence of free radical initiators. Sreekala et al. used acrylic
acid for natural fiber surface modification with no apparent success
Natural Fibers and Polymeric Matrices 375

Bio-fiber OH + CH2=CH-CN Bio-fiber O CH2-CH2-CN


Acrylonitrile

Figure 16.3 Fiber modification by acrylonitrile.

[39, 40]. But Li et al. reported that the tensile strength of acrylic-acid-treated


flax fiber–HDPE composites could be improved and water absorption of
the composites could be decreased [41]. Acrylonitrile has also been used to
modify fibers. The reaction of AN with the fiber hydroxyl groups occurs as
shown in Figure16.3 [42].
Mishra et al. studied the grafting of AN on sisal fibers using a combina-
tion of NaIO4 and CuSO4 as initiator in an aqueous medium at tempera-
tures between 50 and 70°C [43]. It was found that AN-grafted sisal fibers
absorbed the less water compared to untreated fibers.

16.3.6 Treatment with Isocyanates


As discussed in Chapter 7, isocyanates can react readily with the hydroxyl
groups (e.g., cellulose and lignin) in fibers to form urethanes. Paul et al.
have employed isocyanates as coupling agents for cellulosic fibers used in
fiber-reinforced composites [44]. The reaction between fiber and isocya-
nate coupling agent is shown in Figure 16.4 [45].
The mechanical properties of the composites reinforced with isocyanate-
treated wood fibers and PVC or PS as polymer matrix can be improved.
Maldas and Kotka used polymethylene polyphenyl isocyanate (PMPPIC)
in pure state or as solution in plasticizer [46]. PMPPIC is chemically linked
to the cellulose matrix through strong covalent bonds.
Both PMPPIC and PS contain benzene rings, and their delocalized
π-electrons provide strong interactions, so that there is an adhesion
between PMPPIC and PS. Comparing both treatment with silanes and
treatment with isocyanates, it is obvious that the isocyanate treatment is
more effective than the treatment with silane. Equal results are obtained,
when PMPPIC is used for the modification of either the fibers or the poly-
mer matrix [46].

R-NCO + HO Biofiber RNHCO-O Biofiber

Figure 16.4 Fiber modification by isocyanate.


376 Adhesives for Wood and Lignocellulosic Materials

Isocyanate groups bond to wood fiber through urethane linkages,


which are more stable to hydrolysis than esterification [47, 48]. Due
to the difference in molecular structure, the reactivity of isocyanate
decreases in the following order: PMPPIC, toluene diisocyanate (TDIC),
hexamethylene diisocyanate (HMDIC), and ethyl isocyanate (EIC) [49].
Composites prepared with PMPPIC as a coupling agent had the highest
tensile strength, compared with those made with other types of coupling
agents. TDIC has better coupling effectiveness than HMDIC and EIC.
The Kokta group in Canada made a number of investigations on isocya-
nate, alkoxysilane, and anhydride coupling agents. Through their efforts,
PMPPIC has been successfully used as an important coupling agent in
melt-blended composites.
Joseph et al. studied the chemical treatment of the cardanol derivative
of toluene diisocyanate (CTDIC) in sisal fiber–LDPE composites [50].
CTDIC was prepared by cardanol and TDIC in such a way that one free
isocyanate group is available for further reaction. It was demonstrated that
CTDIC reduced the hydrophilic nature of the sisal fiber, thereby enhancing
the tensile properties of the sisal fiber–LDPE composites.
George et al. treated pineapple leaf fiber with PMPPIC solution at
50oC for 30 min to improve the fiber–matrix interfacial adhesion [51].
Comparing silane and isocyanate-treated wood fiber–PS composites,
it was reported that isocyanate treatment was more effective than silane
treatment in enhancing the mechanical properties of cellulose fiber–PS
composites [46, 48]. Sreekala et al. reported that TDIC-treated oil palm
fiber–PF composites had lower tensile strength and Young’s modulus than
permanganate-, peroxide-, and alkaline-treated fiber–PF composites, but
had higher tensile strength than and similar Young’s modulus to silane-
and acrylic-acid-treated fiber–PF composites [38, 39].
Karmarker et al. synthesized a novel vinyl monomer with isocyanate
functionality [52]. The maximum grafting yield achieved in this new cou-
pling agent is ~9% as against 1–2% reported for maleated PP. The func-
tional group in this coupling agent gets grafted as a single-monomer unit
without any oligomerization, which further improves its efficiency as cou-
pling agent. Also, the isocyanate group of this coupling agent is less reac-
tive to water; this is very important as side reactions with residual moisture
in wood can be avoided.

16.3.7 Peroxide Treatment


Peroxide R–O–O–R contains the divalent ion O–O. Organic peroxides
tend to decompose homolytically to free radicals of the form RO*. RO*
Natural Fibers and Polymeric Matrices 377

then reacts with the hydrogen group of the matrix and cellulose fibers.
Most commonly used peroxides for the treatment of natural fibers are
benzoyl peroxide (BPO) and dicumyl peroxide (DCP). The main advan-
tage of peroxide treatment is the quick decomposition of a peroxide
yielding free radical that can react with the hydroxyl group of the fiber
and with the matrix resulting in good fiber/matrix adhesion along the
composite interphase. Like acetylation and benzoylation treatments,
fibers are pretreated with alkali before treating with peroxides. Higher
temperature is more suitable for achieving the complete decomposition
of peroxide [39].

16.3.8 Permanganate Treatment


Potassium permanganate (KMnO4) in acetone solution is mostly used for
permanganate treatment of natural fibers. The fibers are soaked in solu-
tion and its concentration is carefully controlled to form highly reactive
Mn2+ ions that can react with the cellulose hydroxyl groups and form
cellulose–manganate to initiate graft copolymerization. The hydrophilic
tendency of fibers after permanganate treatment is reduced and chemical
interlocking at the interphase is enhanced, providing better fiber/matrix
adhesion [53].

16.3.9 MAH Treatment


Maleated coupling agents such as MAH coupling agents are frequently
used to modify the fiber surfaces and provide efficient interaction with the
functional surface of the fiber and matrix. This method is mostly used to
modify fiber surfaces destined for pairing with a PP matrix. MAH reacts
with the hydroxyl groups in the amorphous region of fiber structure. This
reaction produces brush-like long-chain polymer coating on the fiber sur-
face and reduces its hydrophilic nature [54]. The covalent bonds between
the hydroxyl groups of the fiber and the anhydride groups of MAH bridge
the interface, causing efficient interlocking [55]. Further details are given
in Section 16.4.

16.3.10 Treatment with Chlorotriazines


Chlorotriazines and their derivatives were used as compatibilizers in wood
polyester composites [56, 57]. The active chlorine in chlorotriazines and
their derivatives react with the hydroxyl groups in wood fibers to form
ether linkages, while the carbon–carbon double bonds on their alkyl chains
378 Adhesives for Wood and Lignocellulosic Materials

can graft onto the polymer matrix by forming covalent bonds. Therefore,
chlorotriazines and their derivatives could build a bridge between wood
fiber and polyester matrix, resulting in higher mechanical properties.

16.3.11 Additives
Additives are usually required with coupling agents during the coupling
treatment, especially in graft copolymerization. The most widely used ini-
tiators are organic peroxides, including DCP, BPO, lauroyl peroxide (LPO),
tertbutyl peroxy benzoate (TBPB), and di-tertbutylperoxide (DTBPO).
DCP is usually used with N,N m-phynylene bismaleiimide, MAPP,
PMPPIC, and silanes; and BPO is used with MA, SA, silane A-1100, and
chlorotriazines. TBPB is used as a free radical initiator of MA and acry-
lates. LPO and DTBPO can be used in the silane coupling agents. In graft
reactions, the concentration of peroxide is usually between 0.5% and 1% by
weight. Excess peroxide may adversely affect the mechanical properties of
the composite because molecular chain scission of the polymer and cellu-
lose occurs when peroxide is too abundant. DCP has also been found to be
a better initiator for MA compared with BPO because the free radicals of
DCP have superior thermal stability that leads to better graft performance.
The free radical initiator, 2,2 -azobisisobutyronitrile (also called Vazo), is
usually combined with gamma radiation for graft reaction of styrene and
vinyl monomers. Stearic acid and its metallic salts are used to improve the
dispersibility of wood fibers in the matrix [6].

16.4 Functionalization of Matrices for Compatibility


Matrix-based strategies for improving the fiber–matrix interaction in bio-
fiber composites rely on the use of polymeric coupling agents. They are
obtained by introducing appropriate functional groups on the polymer
matrices to promote the interaction between the biofibers and the ther-
moplastic polymers. They play a very significant role in improving com-
patibility and adhesion between polar wood fibers and nonpolar polymer
matrices. This process can either allow the polar hydroxyl groups (-OH) of
wood fibers to react with reagents, such as PMPPIC, which have a linear
molecular structure similar to the polymer matrix, or create a chemical
interaction between chemicals such as MA and the matrix [46, 48, 49]. Thus,
these coupling agents are functionalized polymers with a backbone that
should be compatible with the matrix, while the functional groups grafted
to this backbone can interact with the reinforcements. Compatibility with
Natural Fibers and Polymeric Matrices 379

the matrix is therefore achieved by using polymers that are similar to or


identical with the matrix material.
According to the exhaustive review on the coupling agents by Lu et
al., MAH is grafted onto polyolefin matrix in the presence of free rad-
ical initiators to modify the polymer matrix [58]. The graft copolymers
so obtained, e.g., MAPE, MAPP, SEBS-MA, and SMA, are subsequently
used as coupling agents [Lu et al. (2000) and the cross references con-
tained therein].
PP can be functionalized with MAH in the presence of DCP. The chem-
ical reaction involved is shown in Figure 16.5.
The succinic anhydride groups of MAH graft polymers are highly reac-
tive and can form covalent bonds to polar polymer (cellulose) backbones
and end groups.
Rozman et al. showed that compounding MAH-treated EFB (palm
empty fruit bunch) with PP matrix in the melt in the presence of a per-
oxide resulted in marked change of mechanical properties due to grafting
reaction of MAH double bond with PP chains [59].
Unfortunately, the grafting reaction of MAH onto PP in the presence
of an organic peroxide initiator is often accompanied by undesirable side
reactions like chain scissions, which lead to severe degradation of the PP

CH3 CH3 CH3

R-O-O-R 2RO. + CH CH2 CH CH2 CH

CH3 CH3 CH3


CH CH2 C CH2 CH + R-OH
.

β−Scission

CH3 CH3 CH3


CH2 O= C C =O
CH CH + CH2 C O
PP-g-MAA
Ideal reaction
O= C C =O
O
Undesirable reaction

Figure 16.5 MAA-g-polypropylene.


380 Adhesives for Wood and Lignocellulosic Materials

backbone. Analysis of such MAA-grafted PP has demonstrated that most


MAH groups have been grafted onto the chain ends of PP macromole-
cules [60].
To mitigate this issue, electron-donating comonomers such as styrene
and α-methylstyrene have been used to activate MAH through charge
transfer complexes for improving grafting efficiency and reducing chain
[61]. Recently, MAH-modified PP or MAPP is a popular coupling agent
for WFPC. Two kinds of MAPP are used in WFPC. One is the MAPP with
a high-molecular weight (M > 30,000) such as 63H, 13H, and Hercoprime
G. The other type, such as Epolene E-43 (or 47L) and 15L, has a low molec-
ular weight (M < 20,000).
Methacrylates can be used in graft reactions. For example, glycidyl
methacrylate (GMA) and HEMA have been used to modify wood fiber
and polymer [46, 62]. GMA is a suitable unsaturated monomer with an
epoxide group, which can react with the hydroxyl group of cellulose, while
the unsaturated moiety can be easily grafted to polyolefin chain by melt
radical reaction. Rozman et al. showed that compounding MAH-treated
EFB with PP matrix in the melt in the presence of a peroxide resulted in
marked change of mechanical properties due to grafting reaction of MAH
double bond with PP chains [59, 63]. Pracella et al. studied structure–
property relationships of PP/Hemp composites modified by means of
grafting reactions of the components, either the fiber or the PP matrix,
with GMA to improve the fiber–matrix interactions [64]. The reaction
involving GMA is shown in Figure 16.6.
GMA is first grafted onto PP via the acrylic double bonds. The glycidyl
group can react with the OH groups of hemp fiber as shown above.

CH3

CH2-C CH3
n O
O=

Hemp
CH2 C C O CH2 HC CH2 + HO

CH3

CH2-C CH3 OH
n
O=

CH2 C C O CH2 CH CH2 O HempGMA

Figure 16.6 Fiber modification by GMA.


Natural Fibers and Polymeric Matrices 381

Grafting was initially conducted in solution or in batch in the melt state.


However, it is now common to produce grafted PP by reactive extrusion,
using twin screw extruders as continuous reactors. A major advantage of
the reactive extrusion process is avoidance of the use of solvent, but the
high temperatures encountered in extrusion can also lead to secondary
reactions, like β-scissions for the PP. Moreover, in such conditions, the
reaction yields are generally lower than those obtained in solution [65, 66].
By far, the most important group among these coupling agents, as
judged by the abundance of references to it in the literature, are polymers
with grafted MAH groups. Sobczak et al. published an exhaustive overview
on the subject [19].
Besides the MAH grafted polymers, other functionalized polyolefins that
have been investigated are oxidized PE, carboxylated PE, acrylic acid func-
tionalized PE, N-vinyl-formamide-grafted PP, and vinyltriethoxysilane-
functionalized PP.
Various isocyanate-functionalized polymers such as PP based on
PMPPIC, polymeric methylene diphenyl diisocyanate, and polybutadiene
isocyanate have been reported [19]. Karmarkar et al. produced a novel
modified coupling agent by grafting m-isopropenyl-α,α-dimethylbenzyl-
isocyanate (mTMI) onto PP (Figure 16.7) [67].
Isocyanate links wood fiber through the urethane structure (or a carba-
mate), which is more stable to hydrolysis than esterification (Figure 16.8) [47,
68]. Due to the difference in molecular structure, the reactivity of isocyanate
decreases in the following order: PMPPIC, TDIC, HMDIC, and EIC [49].
The delocalized π-electrons of the benzene rings in PMPPIC and TDIC lead
to the stronger interaction with PS and other polymer matrices compared
with HMDIC and EIC without π-electrons. Moreover, the cellulose phase and
the polymer phase (PS or FVC) are continuously linked by PMPPIC at the
interface, while the discrete nature of TDIC, HMDIC, and EIC makes them
inferior in this respect [69]. Composites with PMPPIC as a coupling agent
had the highest tensile strength, compared with those made with other types
of coupling agents. Thus, PMPPIC is the best coupling agent in the isocya-
nates, while TDIC has better coupling effectiveness than HMDIC and EIC.

H3C CH3
NCO

H 3C CH2

Figure 16.7 mTMI unsaturated aliphatic isocyanate.


382 Adhesives for Wood and Lignocellulosic Materials

CH3 CH3 CH3

C CH2 C CH2 C CH2

H
H3C CH3 H3C CH3
C C
+ Wood fiber

H3C C CH3 H3C C CH3

NCO NCO

CH3 CH3 CH3

C CH2 C CH2 C CH2

H
H3C CH3 H3C CH3
C C

H3C C CH3 H 3C C CH3

NHC-O-Wood NHC-O-Wood
=O

=O

Figure 16.8 Isocyanate modification of fiber.

16.5 MAH Grafted Polyolefins as Matrix Additives


MAH grafted polyolefins have been employed as a coupling agent for com-
patibilization of natural fibers with polymer matrices to effect improved
material properties as judged by the effects on the material properties of
biocomposites as reported by Sobczak et al. [19].
It has to be mentioned here that not only MAH but also its isostructural
analogues (maleic, fumaric, citraconic, and itaconic acids and their amide,
imide, ester, and nitrile derivatives) are used as polyfunctional monomers
in the synthesis of reactive macromolecules. These reactive polymers have
been employed to prepare high-performance engineering, bioengineering,
and nano-engineering materials. However, these functionalized reactive
macromolecules have not yet been so far employed for natural fiber com-
posites [70].
Natural Fibers and Polymeric Matrices 383

16.6 Reactive Extrusion System


Reactive extruder systems, namely, twin-screw extruders, are employed
as processing equipment for carrying out chemical reactions to introduce
appropriate functional groups onto thermoplastic polymer matrices. This is
to modify the hydrophobic polymers so that they are compatible with the
hydrophilic biofibers. Thus, twin extruders act as continuous-flow reactors
to achieve such a purpose [71, 72]. In this process, thermoplastics, usually
in the form of beads or pellets, are softened and mixed with the fiber trans-
ported by means of a single or two rotating screws, compressed and forced
out of the chamber at a steady rate through a die (Pickering et al. 2016) [73].
Extensive investigations have been carried out for grafting MAH onto
commercial polymers such as PE [74–79], ethylene–propylene (E–P)
copolymer, PP [76, 78–80] and PS in the presence of peroxide initiators.
The screw speed and temperature should be controlled precisely. High
screw speed can result in air entrapment, excessive melt temperatures, and
fiber breakage. Low speeds, however, lead to poor mixing and insufficient
wetting of the fibers. This method is used on its own or for producing pre-
cursor for biofiber composites. Twin-screw systems have been shown to
give better dispersion of fibers and better mechanical performance than
single-screw extruders [73].

References
1. Salmah, H., Marliza, M.Z., Selvi, E., Biocomposites from Polypropylene
and Corn Cob: Effect Maleic Anhydride Polypropylene, in: Advances in
Civil, Environmental and Materials Research (ACEM 14), Busan, Korea,
August 24–28, 2014.
2. Wu, C.S., Improving polylactide/starch biocomposites by grafting polylac-
tide with acrylic acid: Characterization and biodegradability assessment.
Macromol. Biosci., 5, 352–361, 2005.
3. Aziz, S.H., Ansell, M.P., Clarke, S.J., Panteny, S.R., Modified polyester resins
for natural fibre composites. Compos. Sci. Technol., 65, 525–535, 2005.
4. Acha, B.A., Aranguren, M.I., Marcovich, N.E., Composites from PMMA
modified thermoset and chemically treated woodflour. Polym. Eng. Sci., 43,
999–1010, 2003.
5. Karani, R., Krishnan, M., Narayan, R., Biofiber-reinforced polypropylene
composites. Polym. Eng. Sci., 37, 2, 476–483, 1997.
6. Lu, J.Z., Wut, Q., McNabb, H.S., Jr., Chemical coupling of wood fibre and
polymer composites: A review of coupling agents and treatments. Wood
Fibre Sci., 32, 88–104, 2000.
384 Adhesives for Wood and Lignocellulosic Materials

7. Oksman, K. and Lindberg, H., Influence of thermoplastic elastomers on


adhesion in polyethylene-wood flour composites. J. Appl. Polym. Sci., 68,
1845–1855, 1998.
8. Oksman, K., Lindberg, H., Holmgren, A., The nature and location of
SEBS-MA compatibilizer in polyethylene–wood flour composites. J. Appl.
Polym. Sci., 69, 201–209, 1998.
9. Simonsen, J., Jacobsen, R., Rowell, R., Wood-fibre reinforcement of styrene–
maleic anhydride copolymers. J. Appl. Polym. Sci., 68, 1567–1573, 1998.
10. Adekunle, K.F., Åkesson, D., Skrifvars, M., Biobased composites prepared by
compression molding with a novel thermoset resin from soybean oil and a
natural-fiber reinforcement. J. Appl. Polym. Sci., 116, 1759–1765, 2010.
11. Adekunle, K.F., Surface treatments of natural fibres—A review: Part 1. Open
J. Polym. Chem., 5, 41–46, 2015.
12. Sakata, I., Morita, M., Tsuruta, N., Morita, K., Activation of wood surface
by corona treatment to improve adhesive bonding. J. Appl. Polym. Sci., 49,
1251–1258, 1993.
13. Dong, S., Sapieha, S., Schreiber, H.P., Rheological properties of corona mod-
ified cellulose/polyethylene composites. Polym. Eng. Sci., 32, 1734–1739, 1992.
14. Belgacem, M.N., Bataille, P., Sapieha, S., Effect of corona modification on the
mechanical properties of polypropylene cellulose composites. J. Appl. Polym.
Sci., 53, 379–385, 1994.
15. Ragoubi, M., Bienaimé, D., Molina, S., George, B., Merlin, A., Impact of
corona treated hemp fibres onto mechanical properties of polypropylene
composites made thereof. Ind. Crops. Prod., 31, 344–349, 2010.
16. Satyanarayana, K.G., Steam explosion—A boon for value addition to
renewable resources. Met. News, 22, 35–40, 2004.
17. Rowell, R.M., Acetylation of natural fibers to improve performance. Mol.
Cryst. Liq. Cryst., 418, 1, 153–164, 2004.
18. Carr, C.M., Chemistry of the Textiles Industry, 275 pp, Springer, The
Netherlands, 1995.
19. Sobczak, L., Brueggemann, O., Putz, R.F., Polyolefin composites with natural
fibers and wood-modification of the fiber/filler–matrix interaction. J. Appl.
Polym. Sci., 127, 1–17, 2013.
20. Bledzki, A.K., Mamun, A.A., Jaszkiewicz, A., Erdmann, K., Barley husk and
coconut shell reinforced polypropylene composites: The effect of fibre phys-
ical, chemical and surface properties. Compos. Sci. Technol., 70, 854–860,
2010.
21. Lee, S.Y., Chun, S.J., Doh, G.H., Kang, I.-A., Lee, S., Paik, K.-H., Influence
of chemical modification and filler loading on fundamental properties of
bamboo fibers reinforced polypropylene composites. Compos. Mater., 43,
1639–1657, 2009.
22. Farsi, M., Wood–plastic composites: Influence of wood flour chemical mod-
ification on the mechanical performance. J. Reinf. Plast. Compos., 29, 3587–
3592, 2010.
Natural Fibers and Polymeric Matrices 385

23. Nevel, T.P. and Zeronian, S.H., Cellulose Chemistry and Its Applications,
Wiley, New York, 1985.
24. Kumar, R., Obrai, S., Sharma, A., Chemical modifications of natural fiber for
composite materials. Chem. Sin., 2, 4, 219–228, 2011.
25. Danyadi, L., Moczo, J., Pukanszky, B., Effect of various surface modifications
of wood flour on the properties of PP/wood composites. Composites Part A,
41, 199–206, 2010.
26. Gassan, J., Mildner, I., Bledzki, A.K., Transcrystallization of polypropylene
on different modified jute fibers. Compos. Inter., 8, 443–452, 2001.
27. Vazquez, A., Dominguez, V.A., Kenny, J.M.J., Bagasse fiber–polypropylene
based composites. J. Thermoplast. Compos. Mat., 12, 477–497, 1999.
28. Fengel, D. and Wegener, G., Wood: Chemistry, Ultrastructure, Reactions,
482 pp, de Gruyter, Berlin, 1983.
29. Weyenberg, I.V., Truong, T.C., Vangrimde, B., Verpoest, I., Improving the
properties of UD flax fibre reinforced composites by applying an alkaline
fibre treatment. Composites Part A, 37, 1368–1376, 2006.
30. Jacob, M., Thomas, S., Varughese, K.T., Mechanical properties of sisal/
oil palm hybrid fiber reinforced natural rubber composites. Compos. Sci.
Technol., 64, 955–965, 2004.
31. Mishra, S., Mohanty, A.K., Drzal, L.T., Misra, M., Parija, S., Nayak, S.K.,
Tripathyc, S.S., Studies on mechanical performance of biofibre/glass rein-
forced polyester hybrid composites. Compos. Sci. Technol., 63, 1377–1385, 2003.
32. Soleimani, M., Tabil, L., Panigrahi, S., Opoku, A., The effect of fiber pretreat-
ment and compatibilizer on mechanical and physical properties of flax fiber–
polypropylene composites. J. Polym. Environ., 16, 74–82, 2008.
33. Bledzki, A.K., Mamun, A.A., Lucka-Gabor, M., Gutowski, V.S., The effects
of acetylation on properties of flax fibre and its polypropylene composites.
Express Polym. Lett., 2, 6, 413–422, 2008.
34. L.A. Goettler, Treated fibers and bonded composites of cellulose fibers in
vinyl chloride polymer characterized by an isocyanate bonding agent. US
Patent 4,376,144, 1983.
35. Klason, C., Kuhat, J., Stroemvall, H.E., The efficiency of cellulosic fillers in
common thermoplastics, part I. Filling without processing aids or coupling
agents. Int. J. Polym. Mater., 10, 159–187, 1984.
36. Comyn, J., Structural Adhesives: Developments in Resins and Primers, A.J.
Kinloch (Ed.), 269 pp, Applied Science Pub., London, 1986.
37. Joseph, K., Mattoso, L.H.C., Toledo, R.D., Thomas, S., de Carvalho, L.H.,
Pothen, L., Kala, S., James, B., Natural fiber reinforced thermoplastics, in:
Natural Polymers and Agrofibers Composites, E. Frollini, A.L. Leao, L.H.C.
Mattoso (Eds.), pp. 159–201, Embrapa, USP-IQSC, UNESP, San Carlos,
Brazil, 2000.
38. Li, X., Tabil, L.G., Panigrahi, S., Chemical treatments of natural fiber for
use in natural fiber-reinforced composites: A review. J. Polym. Environ., 15,
25–33, 2007.
386 Adhesives for Wood and Lignocellulosic Materials

39. Sreekala, M.S., Kumaran, M.G., Joseph, S., Jacob, M., Thomas, S., Oil palm
fibre reinforced phenol formaldehyde composites: Influence of fibre surface
modifications on the mechanical performance. Appl. Compos. Mater., 7,
295–329, 2000.
40. Sreekala, M.S., Kumaran, M.G., Joseph, S., Thomas, S., Water sorption in oil
palm fiber reinforced phenol formaldehyde composites. Composites Part A,
33, 763–777, 2002.
41. Li, X., Panigrahi, S., Tabil, L.G., Crerar, W.J., Flax fiber reinforced compos-
ites and the effect of chemical treatments on their properties. Proceedings
of CSAE/ASAE Annual Intersectional Meeting, Winnipeg, Canada, 24–25
September (2004).
42. Mohanty, A.K., Misra, M., Drzal, L.T., Surface modifications of natural fibers
and performance of the resulting biocomposites: An overview. Compos.
Interf., 8, 313–343, 2001.
43. Mishra, S., Mishra, M., Tripathy, S.S., Nayak, S.K., Mohanty, A.K., Graft copo-
lymerization of acrylonitrile on chemical modified sisal fibers. Macromol.
Mat. Eng., 286, 107–113, 2001.
44. Paul, A., Joseph, K., Thomas, S., Effect of surface treatments on the electri-
cal properties of low-density polyethylene composites reinforced with short
sisal fibers. Compos. Sci. Technol., 57, 1, 67–79, 1997.
45. Bledzki, A.K. and Gassan, J., Effect of coupling agents on the moisture
absorption of natural fibre reinforced plastics. Angew. Makromol. Chem.,
236, 129–139, 1996.
46. Maldas, D. and Kokta, B.V., Improving adhesion of wood fiber with poly-
styrene by the chemical treatment of fiber with a coupling agent and
the influence on the mechanical properties of composites. J. Adhes. Sci.
Technol., 3, 529–539, 1989.
47. John, W.E., Isocyanate as wood binders: A review. J. Adhes., 15, 59–67, 1982.
48. Maldas, D., Kokta, B.V., Daneault, C., Influence of coupling agents and
treatments, on the mechanical properties of cellulose fiber–polystyrene
composites. J. Appl. Polym. Sci., 37, 751–775, 1989.
49. Kokta, B.V., Daneault, C., Beland, P., Composites of polyvinyl chloride–wood
fibres I. Effect of isocyanate as a bonding agent. Pol.-Plas. Technol. Eng., 29,
87–118, 1990.
50. Joseph, K., Thomas, S., Pavithran, C., Effect of chemical treatment on the
tensile properties of short sisal fibre reinforced polyethylene composites.
Polymer, 37, 5139–5149, 1996.
51. George, J., Janardhan, R., Anand, J.S., Bhagawan, S.S., Thomas, S., Melt rheo-
logical behavior of short pineapple fiber reinforced low-density polyethylene
composites. Polymer, 37, 5421–5431, 1996.
52. Karmarkar, A., Aggarwal, P., Chauhan, S.S., Technology Package on Wood
Polymer Composites, Institute of Wood Science and Technology, Bangalore,
Personal communication, 2012.
Natural Fibers and Polymeric Matrices 387

53. Rahman, M.M., Mallik, A.K., Khan, M.A., Influences of various surface
pretreatments on the mechanical and degradable properties of photo-
grafted oil palm fibers. J. Appl. Polym. Sci., 105, 3077–3086, 2007.
54. George, J., Sreekala, M., Thomas, S., A review on interface modification
and characterization of natural fiber reinforced plastic composites. Polym.
Eng. Sci., 41, 1471–1485, 2001.
55. Keener, T., Stuart, R., Brown, T., Maleated coupling agents for natural fibre
composites. Composites Part A, 35, 357–362, 2004.
56. Zadorecki, P. and Flodin, P., Surface modification of cellulose fibers. II. The
effect of cellulose fiber treatment on the performance of cellulose–polyester
composites. J. Appl. Polym. Sci., 30, 3971–3983, 1985.
57. Zadorecki, P. and Flodin, P., Surface modification of cellulose fibers. I.
Spectroscopic characterization of surface-modified cellulose fibers and their
copolymerization with styrene. J. Appl. Polym. Sci., 30, 2419–2429, 1985.
58. Lu, J.Z., Wut, Q., McNabb, H.S., Jr., Chemical coupling of wood fibre and
polymer composites: A review of coupling agents and treatments. Wood
Fibre Sci., 32, 88–104, 2000.
59. Rozman, H.D., Saad, M.J., Mohd Ishak, Z.A., Modification of oil palm empty
fruit bunches with maleic anhydride: The effect on the tensile and dimen-
sional stability properties of empty fruit bunch/polypropylene composites.
J. Appl. Polym. Sci., 87, 827–835, 2003.
60. Shi, D., Yang, J., Yao, Z., Wang, Y., Huang, H., Jing, W., Yin, J., Costa, G.,
Functionalization of isotactic polypropylene with maleic anhydride by
reactive extrusion: Mechanism of melt grafting. Polymer, 42, 5549–5557,
2001.
61. Li, Y., Xie, X.-M., Guo, B.-H., Study on styrene-assisted melt free-radical
grafting of maleic anhydride onto polypropylene. Polymer, 42, 3419–3425,
2001.
62. Takase, S. and Shiraishi, N., Studies on composites from wood and polypro-
pylenes. II. J. Appl. Polym. Sci., 37, 3, 645–659, 1989.
63. Rozman, H.D., Kumar, R.N., Abdul Khalil, H.P.S., Abusamah, A., Abu, R.,
Ismail, H., Fibre activation with glycidyl methacrylate and subsequent
copolymerization with diallyl phthalate. Eur. Polym. J., 33, 1213–1218, 1997.
64. Pracella, M., Chionna, D., Anguillesi, I., Kulinski, Z., Piorkowskac, E.,
Functionalization, compatibilization and properties of polypropylene com-
posites with hemp fibres. Compos. Sci. Technol., 66, 2218–2230, 2006.
65. Berzin, F., Flat, J.-J., Vergnes, B., Grafting of maleic anhydride on poly-
propylene by reactive extrusion: Effect of maleic anhydride and peroxide
concentrations on reaction yield and products characteristics. Polym. Eng.,
33, 673–682, 2013.
66. Flat, J.-J., Nouveaux développements dans le greffage radicalaire sur poly-
propylène fondu, PhD Dissertation, University of Strasbourg, Strasbourg,
France, 1991.
388 Adhesives for Wood and Lignocellulosic Materials

67. Karmarkar, A., Chauhan, S.S., Modak, J.M., Chanda, M., Mechanical prop-
erties of wood–fiber reinforced polypropylene. Composites A, 38, 227–233,
2007.
68. Maldas, D. and Kokta, B.V., Effect of coating treatments on the mechanical
behaviour of wood fibre-filled polystyrene composites II. Use of inorganic
salt/polyvinylchloride and isocyanate as coating components. J. Reinf. Plast.
Comp., 8, 2–12, 1990.
69. Maldas, D., Kokta, B.V., Raj, R.G., Daneault, C., Improvement of the mechan-
ical properties of sawdust wood fiber–polystyrene composites by chemical
treatment. Polymer, 29, 1255–1265, 1988.
70. Rzayev, Z.M.O., Graft copolymers of maleic anhydride and its isostructural
analogues: High performance engineering materials. Int. Rev. Chem. Eng.,
3, 153–215, 2011.
71. Cha, J. and White, J.L., Maleic anhydride modification of polyolefin in an
internal mixer and a twin-screw extruder: Experiment and kinetic model.
Polym. Eng. Sci., 41, 1227–1237, 2001.
72. Sathe, S.N., Srinivasa Rao, G.S., Devi, S., Grafting of maleic anhydride onto
polypropylene: Synthesis and characterization. J. Appl. Polym. Sci., 53,
239–245, 1994.
73. Pickering, K.L., Aruan Efendy, M.G., Lee, T.M., A review of recent devel-
opments in natural fibre composites and their mechanical performance.
Composites A, 83, 98–112, 2016.
74. Wang, K.H., Choi, M.H., Koo, C.M., Choi, Y.S., Chung, I.J., Synthesis and
characterization of maleated polyethylene/clay nanocomposites. Polymer,
42, 9819–9826, 2001.
75. Hu, G.H., Flat, J.J., Lambla, M., The synthesis of grafted polymers, in: Reactive
Modifiers for Polymers, Al-Malaika (Ed.), Thomson Science, London, 1997.
76. Braun, D., Braun, I., Kramer, I., Heterogeneous grafting of maleic anhydride
and α-methylstyrene from atactic polypropylene. Angew. Makromol. Chem.,
251, 37–48, 1997.
77. De Roover, D., Devaux, J., Legras, R., Maleic anhydride homopoly-merization
during melt functionalization of isotactic polypropylene. J. Polym. Sci., Part
A: Polym. Chem., 34, 1195–1202, 1996.
78. Schulz, B. and Hartmann, M., Radical grafting of maleic anhydride copoly-
mers to atactic polypropylene. Plaste und Kautschuk, 32, 84–86, 1985.
79. Huang, W., Walia, P., Hu, Y., Zhou, Z., Horstman, N., Huang, T., Effect of
ethylene content on maleic anhydride-grafted polypropylene based block
co-polymers, SPE ANTEC Anaheim 2017/1037, 2017.
80. De Roover, D., Sclavons, M., Carlier, V., Devaux, J., Legras, R., Momtaz, A.,
Molecular characterization of maleic anhydride functionalized polypropyl-
ene. J. Polym. Sci., Part A: Polym. Chem., 33, 829–842, 1995.
17
Polymer Matrix: Unsaturated Polyester

17.1 Introduction
Unsaturated polyesters are copolyesters—that is, polyesters prepared from
a saturated dicarboxylic acid or its anhydride (usually phthalic anhydride)
as well as an unsaturated dicarboxylic acid or anhydride (usually maleic
anhydride). The unsaturation present in this type of polyesters provides a
site for subsequent cross-linking [1, 2].
Unsaturated polyesters are prepared commercially by the reaction of a
saturated diol with a mixture of unsaturated dibasic acid and a “modify-
ing” saturated acid (or corresponding anhydride) [3].
The unsaturation in principle can be derived either from an unsaturated
acid or unsaturated diol. However, for reasons of economy, unsaturated
acids are preferred.
The addition of saturated acids controls the cross-link density and brit-
tleness of the cured product.

17.2 Raw Materials


17.2.1 Diols
Propylene glycol is generally used for linear unsaturated polyesters for
many thermoset composite applications. For applications that require
more flexible or hydrolytically stable resins, dipropylene glycol is used.
Other alcohols like neo-pentyl glycol, di-ethylene glycol, and ethylene
glycol are also used for the production of unsaturated polyester resins
(Figures 17.1 and 17.2). Each of these alcohols contributes to the final poly-
mer characteristics, such as heat distortion temperature (HDT), physical
strength, water uptake, and weather resistance [4].

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (389–401) © 2019
Scrivener Publishing LLC

389
390 Adhesives for Wood and Lignocellulosic Materials

CH3 CH3

HO CH2 CH OH HO CH2 C CH2 OH

Propylene glycol CH3


Neopentyl glycol

CH3 CH3

HO CH2 CH O CH2 CH OH
Diropylene glycol

HO CH2 CH2 O CH2 CH2 OH

Diethylene glycol

Figure 17.1 Diols employed for UP resins.

CH3 CH3 CH3

HO CH CH2 O C O CH CH2 OH

CH3
Propoxylated Bisphenol A

Cyclopentadiene

Figure 17.2 Other speciality chemicals used for UP synthesis.

17.2.2 Cyclopentadiene-Based Resin


Dicyclopentadiene (DCPD) is employed along with maleic anhydride
during the initial stage prior to both isomerization and condensation
polymerization. DCPD improves the solubility of polyester in styrene.
DCPD reduces the cost of the resin. Due to environmental regulations
to reduce styrene emissions in open molding, “low-styrene” resins have
been developed. These resins are based on DCPD and are produced with
lower molecular weight. This allows less styrene to be used to achieve the
desired viscosity. Frydrych et al. found that when UP resins are modified
with DCPD in amounts of 6 and 12 wt.%, DCPD was found to enhance
Polymer Matrix: Unsaturated Polyester 391

the solubility of the polyester in styrene [5]. Penczek reported that incor-
poration of DCPD into UP resin causes improvement in flexural strength,
compression strength, and hardness and increase in HDT value. DCPD
also improves chemical resistance, particularly to water and alkaline
media [6].

17.2.3 Isophthalic-Acid-Based Resin


Isophthalic resins are based on isophthalic acid and maleic anhydride. A
high-molecular-weight resin is obtained with good chemical and thermal
resistance and good mechanical properties. The use of nonpolar glycols
contributes to improved aqueous resistance.

17.2.4 Bisphenol A Fumarate Resins


They are prepared from propoxylated BPA and fumaric acid to obtain
a relatively nonpolar polyester with reduced number of ester linkages.
This contributes to excellent corrosion resistance. Bisphenol A fuma-
rates were once predominantly used for the corrosion-resistant com-
posites. In recent years, they have been replaced by isophthalic resins
for mildly corrosive applications and vinyl esters for more aggressive
environments.

17.2.5 Vinyl Ester


Vinyl esters are produced by the esterification of an epoxy resin with an
unsaturated monocarboxylic acid such as acrylic acid or methacrylic acid;
i.e., the epoxy resin is “end-capped” with an acrylic or methacrylic group.
The reaction product is then dissolved in a reactive solvent, such as styrene,
to approximately 35–45% content by weight. If a multifunctional acrylate
monomer such as trimethylolpropane triacrylate or diethyleneglycol dia-
crylate is used instead of styrene, the product is called “epoxy acrylate”,
which is used for radiation curing.
Vinyl esters can be used in place of polyester and epoxy resins as the
thermoset polymer matrix in composite materials, where its perfor-
mance and cost are intermediate between polyester and epoxy. Vinyl
esters have lower resin viscosity than polyester and epoxy. It is a resin
suitable for marine applications because of its increased corrosion
resistance.
The structure of vinyl ester is given in Figure 17.3.
392 Adhesives for Wood and Lignocellulosic Materials

OH CH3 OH

O=
O=
CH CH2 O C O CH2 CH CH2 O C CH=CH2
CH2=CH C O CH2

CH3

Figure 17.3 Vinyl ester resin.

17.3 Polyesterification Reaction


Maleic anhydride reacts with the diol propylene glycol in two steps:

(1) Formation of a monoester by the reaction of maleic anhy-


dride with a molecule of propylene glycol leaving a free
carboxyl group (Figure 17.4).
O=

CH3
O=

C OH
C
O + OH CH2 CH OH
C CH2 OH
C O CH
=

=
O

CH3

Figure 17.4 Formation of monoester.

(2) The monoester then reacts with a molecule of propylene


glycol to form a diester (Figure 17.5).

CH3
O=

Propyleneglycol C O CH CH2 OH
O=

C OH
C O CH CH2 OH
O=

C O CH CH2 OH
=

CH3
O

CH3

Figure 17.5 Formation of diester.

Due to the higher reactivity of the anhydride group compared with the
acid group, the first step proceeds more rapidly than the second. Further
reactions between the diols and the anhydride will continue in the same
Polymer Matrix: Unsaturated Polyester 393

manner, resulting in chain extension leading to a polyester chain. Thus, a


segment of polyester resulting from the reaction between diol, maleic anhy-
dride, and phthalic anhydride will have the structure shown in Figure 17.6.

CH3 CH3

O O CH CH2 O OC CH=CH CO
OC CH=CH CO O CH CH2

OC CO O CH2 CH O

CH3

Figure 17.6 Unsaturated polyester oligomer.

Appreciable cis–trans isomerization occurs during the polyesterifica-


tion reaction, and accordingly, maleic anhydride is incorporated into the
polymer chain as fumarate.

17.4 Cross-Linking Reaction


Cross-linking of the unsaturated polyester resin occurs at the unsaturated
sites in the polymer chain with vinyl-type monomers by a free radical
mechanism. The cross-linking can take place either at room temperature
or at elevated temperatures depending on the type of free radical initiators
employed for the purpose.

17.4.1 Curing at Elevated Temperatures


The most important initiators employed at elevated temperatures are per-
oxides such as benzoyl peroxide and tert-butylperbenzoate (Figure 17.7).
O=

O=

C O O C

Benzoyl peroxide
O=

C(CH3)3-O-O-C

ter-butylperbenzoate

Figure 17.7 Benzoyl peroxide and tert-butylperbenzoate.


394 Adhesives for Wood and Lignocellulosic Materials

Initiating radicals are formed due to heat as shown in Figure 17.8.

O=
O=

O=
C O O C 2 C O.

Benzoyl peroxide
O=

2 O. 2 . + CO2
C

Figure 17.8 Production of free radicals.

17.4.2 Curing at Room Temperatures


Initiators that are effective for curing of unsaturated polyesters at ambient
temperature normally consist of mixtures of peroxy compounds and an
“accelerator”. In the presence of accelerators, the peroxy compounds rap-
idly decompose without application of heat to generate free radicals. The
most important peroxy compounds for curing unsaturated polyesters at
room temperature are methyl ethyl ketone peroxide (MEKP) and cyclo-
hexanone peroxide (Figure 17.9). These names are rather misleading [3]
since they are not single compounds and have variable compositions.

MEKP
CH3
OH O O
CH3
C CH3 C
C
HO C2H5
C2H5 O-OH C2H5 OH

CH3
CH3 O
O O
O CH3 C
CH3 C
C
C C2H5
C2H5 HO-O
HO O-OH
C2H5 O-OH C2H5

Cyclohexanone peroxides

OH O O O O

O-OH OH HO HO
O-OH
O O

O-OH HO-O

Figure 17.9 Room temperature curing free radical initiators.


Polymer Matrix: Unsaturated Polyester 395

The most common accelerators for MEKP and cyclohexanone peroxide


are salts with variable valences. A widely used metal of this kind is cobalt as
naphthenate, which is readily soluble in the polyester. Octoates are also used.
The decomposition of hydroperoxide in the presence of the metal salt to
give free radicals occurs due to a redox reaction as follows:

R-O-O-H + Co++ → R-O • + Co+++ + OH−

R-O-O-H + Co+++ → R-O-O • + Co++ + H+

17.5 Sheet Molding Compounds Based on UP Resins


Sheet molding compounds (SMCs) consist of unsaturated polymeric
resin, inert fillers, fiber reinforcement, latent catalysts, pigments and
stabilizers, internal release agents, and thickeners. The SMC is produced
in the sheet form sandwiched between two polyethylene sheets. The
manufacture of SMCs is a continuous process. The SMC-formulated
paste is first spread uniformly onto the bottom film. Chopped glass
fibers/natural fibers are randomly deposited onto the paste. The top film
is introduced and the sandwich is rolled into a predetermined thickness.
The sheet is allowed to mature for 48 h. During this period, the material
undergoes thickening and acquires a tack-free leathery consistency,
which can be easily handled during compression molding [7]. The thick-
ening is caused by a chain extension reaction involving MgO as described
in Figure 17.10.

OH C Polyester C OH + MgO OH C Polyester C O MgOH

O O O O

OH C Polyester C O MgOH + OH C Polyester C OH


–H2O
O O O O

OH C Polyester C O Mg O C Polyester C OH

O O O O

Figure 17.10 UP thickening reaction.


396 Adhesives for Wood and Lignocellulosic Materials

A typical formulation for the SMC for general purpose is given


below:

Ortho-phthalic resin 100


Styrene 5
Tert-butylperbenzoate 1.5
Polyethylene powder 5
Zinc stearate 5
Pigment paste 5
Calcium carbonate 160
MgO paste 3

17.6 UV Curable Compositions Based on UP/Vinyl


Ester Resins
When exposed to ultraviolet radiation or sunlight, unsaturated polyes-
ters and vinyl esters can undergo cross-linking reactions in the presence
of suitable photoinitiators. Such a system is ideally suited for producing
glass/biofiber composites.
An important photoinitiator that can be employed for curing thick sec-
tions of polymer–matrix composites is acyl-/bisacylphosphine oxide
(BAPO), the structure of which is presented in Figure 17.11.
In acylphosphine oxide photo-polymerization initiators, the chromo-
phore structure after its photo-cleavage is significantly different from that
before cleavage so that the absorption spectrum largely changes to cause a
decrease in absorption, which is known as photo-bleaching. Because of the
photo-bleaching property of this type of photoinitiators, they are ideally
suited to cure a thick section of UV curable coatings and fiber-reinforced
composites.

O O O
P

BAPO

Figure 17.11 BAPO photoinitiator.


Polymer Matrix: Unsaturated Polyester 397

17.7 Biocomposites Based on UP Matrix


Kumar reported the results on the development of SMCs based on oil palm
fruit pressed fibers for applications in price-driven biofiber-based com-
posites [8]. Kumar et al. developed fire-resistant SMCs made from hybrid
reinforcements of oil palm fiber and glass fiber [8]. Aluminum trihydrate
(aluminum hydroxide) and magnesium hydroxide were employed as the
fire-retardant additive.
The chemical reactions through which these two flame retardants
decompose endothermically are as follows [9]:

2 A1(OH)3 → A12O3 + 3 H2O ΔH = 200 cal/g

Mg(OH)2 → MgO + H2O ΔH = 328 cal/g

Evolution and evaporation of water causes the endotherm.


Kumar also reported the novel development of a photofabrication pro-
cess to produce biofiber-based composites [10]. A vinyl ester was used
together with BAPO as photoinitiator. The cure was effected by exposing
the impregnated biofiber mat to ultraviolet radiation or solar radiation.
Dhakal et al. employed needle-punched, randomly oriented non-woven
hemp fiber fabric of 330 g/m2 weight as the reinforcement [11]. The matrix
material used in this study was based on a commercially available unsat-
urated polyester. The matrix was mixed with curing catalyst, MEKP, at a
concentration of 0.01 w/w of the matrix for curing.
A combination of hand lay-up and compression molding method was
used to prepare the composite samples. After initial room temperature
curing for 24 h, post-curing was carried out at a temperature of 80°C for
3 h. The effect of water absorption on the mechanical properties of the
composites was studied. It was found that the water absorption pattern of
these composites at room temperature is found to follow Fickian diffusion.
Kushwaha and Kumar prepared polyester composites reinforced
by mercerized bamboo mats and determined the effect of silane
treatments on the water absorption properties of the composites [12].
Silanes  employed for treatment were as follows: γ-aminopropyltriethoxy
silane, 3-trimethoxysilylpropyl methacrylate, vinyl-tris(2-methoxye-
thoxy)  silane, bis[3-(triethoxysilyl)propyl]tetrasulfide, 3-aminopropyl
trimethoxy silane, and n-n-octyl-trimethoxy silane. The water-resistant
property of the bamboo fiber-reinforced composites was determined.
398 Adhesives for Wood and Lignocellulosic Materials

Sreekumar et al. studied the mechanical properties of sisal-leaf fiber-


reinforced polyester composites prepared by resin transfer method (RTM)
and compared with the compression molding techniques [13]. RTM is a
novel technology, which bridges the gap between labor-intensive hand
lay-up processes and capital-intensive compression molding. The study
focused on the tensile and flexural behavior of sisal-fiber-reinforced poly-
ester composites as a function of fiber length and fiber content.
Sanadi et al. determined the tensile and impact behavior of polyester
composites reinforced with continuous unidirectional sunhemp fibers
[14]. The analysis of various energy-absorbing mechanisms during impact
fracture showed that fiber pullout and interface fracture were the major
contributions towards the high toughness of these composites. The results
of this study indicated that sunhemp fibers have potential as reinforcing
fillers in plastics in order to produce inexpensive materials with a high
toughness.
Surata et al. used rice husk as reinforcement in composite products
employing unsaturated polyester resins as polymer matrix and MEKP as
catalyst [15]. Composites were made by hand lay-up techniques, with the
variation of the fiber weight fraction of 20%, 30%, 40%, and 50%. Tensile
test specimens were made according to the ASTM D3039, and flexural test
specimens were made according to the ASTM D790M. The results showed
that the tensile and flexural strength of the composites increased as the
fiber weight fraction increased.
Uma et al. made pineapple leaf fiber (PALF)-reinforced polyester com-
posites and determined their tensile, impact, and flexural properties as a
function of fiber loading, fiber length, and fiber surface modification [16].
The tensile strength and Young’s modulus of the composites were found to
increase with fiber content in accordance with the rule of mixtures. The
elongation at break of the composites exhibited an increase by the intro-
duction of the fibers. The mechanical properties are optimum at a fiber
length of 30 mm. The specific flexural stiffness of the composite was found
to be about 2.3 times greater than that of a neat polyester resin. Significant
improvement in the tensile strength was observed for composites with
silane A172-treated fibers. Scanning electron microscopic studies were
carried out to understand the fiber–matrix adhesion, fiber breakage, and
failure topography. The PALF-reinforced polyester composites possess
superior mechanical properties compared to other cellulose-based natural
fiber composites.
Muthukumar et al. carried out an experimental investigation to compare
the mechanical properties of sisal-, jute-, and kenaf-fiber-reinforced com-
posites with those of glass-fiber-reinforced polyester matrix composites
Polymer Matrix: Unsaturated Polyester 399

[17]. Hybrid composites were manufactured by the hand lay-up technique.


The tensile, flexural, and impact tests were carried out on different com-
posite samples as per the ASTM standards. It was observed that the tensile
strength of a jute/glass fiber composite is 1.94 and 1.59 times more than
that of sisal/glass and kenaf/glass composites, respectively. The flexural
load carrying capacity of sisal/glass composite is 3.4 and 2.83 times greater
than those of jute/glass and kenaf/glass composites, respectively. It can also
be seen that the impact strength of a jute/glass composite is almost equal
to that of a kenaf/glass composite and 1.13 times more than that of a sisal/
glass composite.
Natural fibers have been known for their good acoustic damping prop-
erties, and therefore, these materials could be used as a sound insulation
in many applications. Abdullah et al. reported on the investigation on
the sound absorption coefficient of sugarcane baggase fiber, banana fiber,
and its hybrid-based composites under various fiber volume fractions
[18]. The sound absorption frequencies were measured using the two-
microphone transfer function technique in the impedance tube that has
a 100-mm diameter for low frequency and 28 mm for high frequency,
0 Hz to 4000 Hz, respectively. The result indicated that in low and high
frequency, the combination of different natural fibers produced better
sound absorption coefficient rather than using each natural fiber individ-
ually. The results also demonstrated that the higher amounts of fiber vol-
ume fraction affect frequency broadening, hence promising better sound
absorbing capacity.
Prasanna Venkatesh et al. carried out an investigation on the tensile, flex-
ural, and impact properties and water absorption of sisal/unsaturated poly-
ester composite material [19]. Initially, the optimum fiber length and weight
percentage were determined/estimated. To improve the tensile, flexural, and
impact properties, sisal fiber was hybridized with bamboo fiber. This work
shows that the addition of bamboo fiber in sisal/unsaturated polyester com-
posites of up to 50% by weight results in increasing the mechanical proper-
ties and decreasing the moisture absorption. The treated hybrid composite
was compared with an untreated hybrid composite, with the former show-
ing a 30% increase in tensile strength, a 27.4% increase in flexural strength,
and a 36.9% increase in impact strength, along with a decrease in moisture
absorption. The fractured surface of the treated fiber composite specimen
was studied using scanning electron microscopy.
Prasad et al. prepared a lightweight composite material using banana
and empty fruit bunch fiber (banana-EFB) as reinforcement in polyester
resin matrix and studied the mechanical properties [20]. The composites
were formulated up to a maximum fiber volume fraction of about 0.37,
400 Adhesives for Wood and Lignocellulosic Materials

resulting in a mean tensile strength of 43 MPa and a tensile modulus of


1.06 GPa, which are 36% and 68% higher than those of the plain polyester,
respectively. The specific flexural modulus of the composite was found to
be 1.42 times that of polyester resin and the work of fracture in impact was
found to be 141.7 J/m.

References
1. Boenig, H.V., Unsaturated Polyesters: Structure and Properties, Elsevier,
Amsterdam, 1964.
2. Parkyn, B., Lamb, F., Clinton, B.V., Polyesters; Unsaturated Polyesters and
Polyester Plasticizers, Elsevier, New York, 1967.
3. Saunders, K.J., Organic Polymer Chemistry, Chapman and Hall, London,
1973.
4. Ashton-Acton, Q., Anhydrides—Advances in Research and Application,
Scholarly Editions, Atlanta, Georgia 2013.
5. Frydrych, A., Ostrysz, R., Penczek, P., The effect of built-in dicyclopenta-
diene on selected properties of unsaturated polyester resins. Polimery, 44,
745–749, 1999.
6. Penczek, P., Abramowicz, D., Rokicki, G., Ostrysz, R., Unsaturated polyes-
ter resins modified with vegetable oil and dicyclopentadiene. Polimery, 49,
767–773, 2004.
7. Kumar, R.N., Rozman, H.D., Abusamah, A., Chin, T.H., Sheet moulding
compounds based on palm fruit pressed fibre, National Seminar on
Utilization of Palm Tree and Other Palms, Forest Research Symposium on
Oil Palm By-Products from Agra-Based Industries, Kuala Lumpur, PORIM
Bulletin. 11, 99, 5, 1985.
8. Kumar, R.N., Wei, L.M., Rozman, H.D., Abusamah, A., Fire resistant SMC
from hybrid reinforcements of oil palm fibre and glass fibre. Int. J. Polym.
Mater., 37, 43–52, 1997.
9. Mohammed, M., Ansari, M.N.M., Pua, G., Jawaid, M., Saiful Islam, M., A
review on natural fiber reinforced polymer composite and its applications.
Int. J. Polym. Sci., 2015, 15 pp, 2015. Article ID 243947.
10. Kumar, R.N., Hee, K.C., Rozman, H.D., Photofabrication of bio-fibre-based
polymer matrix composites, J. Appl. Polym. Sci., 95, 1493–1499, 2005.
11. Dhakal, H.N., Zhang, Z.Y., Richardson, M.O.W., Effect of water absorption
on the mechanical properties of hemp fibre reinforced unsaturated polyester
composites, Comp. Sci. Technol., 67, 1674–1683, 2007.
12. Kushwaha, P.K. and Kumar, R., Studies on water absorption of bamboo–
polyester composites: Effect of silane treatment of mercerized bamboo.
Polym.-Plast. Technol. Eng., 49, 1, 45–52, 2009.
Polymer Matrix: Unsaturated Polyester 401

13. Sreekumar, P.A., Joseph, K., Unnikrishnan, G., A comparative study on


mechanical properties of sisal-leaf fibre-reinforced polyester composites pre-
pared by resin transfer and compression moulding techniques. Comp. Sci.
Technol., 67, 453–461, 2007.
14. Sanadi, A.R., Prasad, S.V., Rohatgi, P.K., Sunhemp fibre-reinforced polyester.
J. Mat. Sci., 21, 4299–4304, 1986.
15. Surata, I.W., Suriadi, I.G.A.K., Arnis, K., Mechanical properties of rice husks
fiber reinforced polyester composites. Int. J. Mat. Mech. Manuf., 2, 2, 165,
2014.
16. Uma Devi, L., Bhagawan, S.S., Sabu, T., Mechanical properties of pineapple
leaf fiber-reinforced polyester composites. J. Appl. Polym. Sci., 64, 1739–1748,
1997.
17. Muthukumar, V., Venkatasamy, V., Mariselvam, A., Sureshbabu, A.,
Senthikumar, N., Fernando, A.A.G., Comparative investigation on mechan-
ical properties of natural fibre reinforced polyester composites. Appl. Mech.
Mater., 592–594, 92–96, 2014.
18. Abdullah, A.H., Azharia, A., Salleh, F.M., Sound absorption coefficient of
natural fibres hybrid reinforced polyester composites. J. Teknol., 76, 9, 31–36,
2015.
19. Prasanna Venkatesh, R., Ramanathan, K., Srinivasa Raman, V., Tensile, flex-
ural, impact and water absorption properties of natural fibre reinforced poly-
ester hybrid composites. Fibres Text. East. Eur., 24, 3, 117, 2016.
20. Prasad, A.V.R., Rao, K.M., Nagasrinivasulu, G., Mechanical properties of
banana empty fruit bunch fibre reinforced polyester composites. Indian J.
Fibre Text. Res., 34, 2, 162–167, 2009.
18
Polymer Matrix: Epoxy Resins

18.1 Introduction
The term “epoxy resin” represents a broad class of cross-linked polymers
with the cross-linking derived from the oxirane or epoxy group:

C C
(18.1)

This ring approximates as an equilateral triangle, which makes it


strained, thus highly reactive.
Epoxies were developed, independently, principally by Ciba AG
(Switzerland) and Devoe and Reynolds Co. (USA) [1].
Epoxy resins are employed extensively in structural and speciality com-
posite applications since they offer a unique combination of properties that
are not attainable from other thermosetting resins. They are available in a
wide variety of physical forms from low-viscosity liquid to high-melting
solids. High strength, low shrinkage, excellent adhesion to various sub-
strates, outstanding electrical insulation properties, chemical and solvent
resistance, and relatively low cost are some of the noteworthy attributes of
the epoxy polymers. They cure with ease. No volatiles or by-products are
evolved during the cure [2].
The primary cross-linking of epoxy resin takes place through the ring
opening reaction of the epoxy group.

18.2 Resin Preparation


General-purpose epoxy resins are prepared by reacting epichlorohy-
drin and bisphenol A to give diglycidyl ether of bisphenol A (DGEBA)
with a molecular weight of 340 (Figure 18.1). Higher-molecular-weight

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (403–423) © 2019
Scrivener Publishing LLC

403
404 Adhesives for Wood and Lignocellulosic Materials

CH3

HO C OH O

CH3 Cl CH2 HC CH2

Bisphenol-A Epichlorohydrin

(18.2) (18.3)

Figure 18.1 Raw materials for epoxy resin.

resins in the liquid, semisolid, to hard crystalline forms can be pro-


duced by increasing the ratio of bisphenol A to epichlorohydrin during
manufacturing. They are linear polyethers with glycidyl end groups
(Figure 18.3).
The details of preparation and the mechanism of reaction as well as the
relationship between molecular weight, the softening point, and epoxy
equivalents are given by Saunders [1].

18.3 Characteristics of Epoxy Resins


Epoxy resins used in commercial composite applications can be catego-
rized as those suitable for structural and high-temperature applications
and those best suited to nonstructural or low-temperature applications.
A primary indicator of service temperature is the glass transition tem-
perature (Tg). The Tg is the temperature below which a polymer exists
in the glassy state, whereas above this temperature, individual segments
are able to move relative to each other in what is termed as the “rubbery
state”.

18.3.1 Epoxy Equivalent


Another key characteristic that determines the resin suitability for use is
the epoxy equivalent (EEW), which is defined as the weight of the resin
per epoxy group. The equivalent weight of the polymer is used to calcu-
late the stoichiometric ratio between the epoxy resin and the curing agent.
Dividing the molecular weight of the resin by the number of epoxy groups
per molecule is the equivalent weight of the resin.
Polymer Matrix: Epoxy Resins 405

18.3.2 Enhancement of Properties


Rigidity can be built into the cured matrix in several ways:

(1) Through incorporation of aromatic groups


(2) By Increasing the number of reactive sites (epoxy groups)
per molecule
(3) By reducing the distance between reactive sites

18.3.3 Types of Epoxy Resins


The three primary classes of epoxies used in composite applications are as
follows:

(1) Phenolic glycidyl ethers


(2) Aromatic glycidyl amines
(3) Cycloaliphatic epoxies

18.3.4 Bisphenol A Glycidyl Ethers


They are formed by the condensation of epichlorohydrin with a phenolic
group. Different types of resins can be formed by the choice of the appro-
priate structure of phenol-containing molecule and the number of pheno-
lic groups per molecule.
The first commercial resin in this category is the DGEBA, which is widely
used today [1, 3, 4]. The structure of DGEBA is shown in Figure 18.2.

CH3 O
HO C OH + 2 Cl H2C HC CH2

CH3 Epichlorohydrin
Bisphenol A NaOH

O CH3 O

H2C CH CH2 O C O H2C HC CH2

CH3
Diglycidylether of Bisphenol A (DGEBA)

Figure 18.2 Structure of DGEBA.


406 Adhesives for Wood and Lignocellulosic Materials

18.4 Preparation of DGEBA Epoxy Resin


The preparation consists of first producing DGEBA as given by structure
in Figure 18.2. This is then followed by a chain extension to increase the
molecular weight to produce epoxy oligomer (Figure 18.3).
The advantage of this epoxidation strategy is the possibility of tailor-
ing the prepolymer’s length and molecular weight. The molecular weight
and softening point depends on the ratio of epichlorohydrin and bisphe-
nol A as mentioned earlier. The chain extension reactions are given below
(where ECH is epichlorohydrin and BPA is bisphenol A):
The typical experimental details of the preparation of epoxy resin are
given by Saunders [1].
Various types of epoxy resins are available from multiple suppliers. The
primary distinction between these resins is their viscosity, which can range
from 14 Pa.s to 5000–14,000 Pa.s at 25°C. The molecular weight increases
as the ratio of ECH to BPA decreases. The viscosity depends on the molec-
ular weight. Lower-molecular-weight materials have lower viscosities.
Aromatic benzene rings of bisphenol A can be replaced by cyclohexane
rings to produce a cycloaliphatic material. This results in a low-viscosity,
moderately reactive resin with a structure analogous to the DGEBA type.

18.4.1 Curing Agents


The epoxy resin as depicted by the structure in Figure 18.3 is linear and
therefore thermoplastic in nature. The uncross-linked resins will have
poor mechanical, chemical, and heat resistance. However, by opening the
epoxy ring, three-dimensional cross-linked structures can be obtained.
The carbon atoms of the epoxy ring are electrophile because of different
electronegativity of carbon and oxygen. This atomic arrangement thus
shows enhanced reactivity. Thus, epoxies or the epoxy functions are highly
reactive and usually undergo ring-opening reactions with nucleophiles [5].
This can be achieved by using appropriate hardeners or curing agents. Most
of the curing agents in common use can be classified into three groups [1]:

(a) Tertiary amines


(b) Polyfunctional amines
(c) Acid anhydrides
CH3 OH CH3
O

H 2C CH CH2 O C O H2C CH O C OH

CH3 CH3

ECH

O CH3 OH CH3

H 2C CH CH2 O C O H 2C CH O C O

CH3 CH3
O
BA/ECH H 2C CH CH2

CH3 OH CH3
O

H2C CH CH2* O C O H2C CH *O C O

CH3 CH3
O
n
Polymer Matrix: Epoxy Resins

H2C CH CH2

Figure 18.3 Epoxy oligomer based on bisphenol A.


407
408 Adhesives for Wood and Lignocellulosic Materials

18.4.1.1 Tertiary Amines


Examples of tertiary amines used as curing agents for epoxy resins include
benzyldimethylamine (6), 2-(dimethylaminomethyl)-phenol (DMP) (7),
2,4,6-tris(dimethylaminomethyl)-phenol (DMP-30) (8), triethanolamine
(9), and N-n-butylimidazole (10) (Figure 18.4) [1].
The reaction between an epoxy resin and a tertiary amine is thought to
proceed as shown in Figure 18.5 [1].
The fact that chemically bound nitrogen is found in the final product
confirms the above reaction scheme. Tertiary amines are called “catalytic”
curing agents since they only induce the linkage of epoxy groups to each
other and do not participate in the reaction [1].

18.4.1.2 Polyfunctional Amines


Polyfunctional primary amines are an important class of hardeners for
epoxy resins. Primary amines react with the epoxide group to form a
hydroxyl group and a secondary amine. The secondary amine can fur-
ther react with an epoxide to form a tertiary amine and an additional
hydroxyl group. The amine type will alter both the processing properties
like mechanical, temperature, and heat resistance of cured copolymer net-
works [5].

CH2-N(CH3)2 OH

CH2-N(CH3)2

Benzyldimthylamine (BDMA) 2-(Dimethylaminomethyl)- Phenol


6 7

OH N
N(CH2CH2OH)3
N(CH3)2-CH2 CH2-N(CH3)2
Triethanolamine N
9
CH2-N(CH3)2 C4H9
2,4,6-tris(dimethylaminomethyl)-phenol N-n-butylimidazole
8 10

Figure 18.4 Tertiary amines curing agents.


Polymer Matrix: Epoxy Resins 409

(i) Formation of quaternary base

O
O
+
R3N: H2C CH R3N CH2 CH

(ii) Protonation of quaternary base with the formation of an anion

O OH

+ + + R'O
CH2 + R'OH R3N CH2 CH
R3N CH

O OH

+ + + OH
R3N CH2 CH + H2O R3N CH2 CH

(iii) Polymerization through epoxy group (initiated by the above-formed anion)

O
O
O
H2C CH
R'O R'O CH2 CH
H2C CH

O O

CH2 CH O CH2 CH
O
O
CH2 CH
H2C CH
R'O CH2 CH O
etc
R'O CH2 CH

11

Figure 18.5 Scheme of reaction of epoxy resin with tertiary amines.


410 Adhesives for Wood and Lignocellulosic Materials

Both aliphatic and aromatic compounds having at least three active


hydrogen atoms present in the primary and/or secondary amine groups
are widely used as curing agents for epoxy resins. Examples of such poly-
functional amines are shown in Figure 18.6.
The reaction between an epoxy resin and a polyamine results in a cross-
linked polymeric structure as illustrated by the reaction with triethylene
tetramine (with six reactive hydrogens) (Figure 18.7):

H2N-CH2-CH2-NH2 H2N-(CH2)2-NH-(CH2)2-NH2 H2N-(CH2)2-NH-(CH2)2-NH2-(CH2)2-NH2

Ethylenediamine Diethylenetriamine Triethylenetetramine


12 13 14

NH2

H2N CH2 NH2


NH2
4,4' Diaminodiphenylmethane
15 m-Phenylenediamine (MPD)
17

O
=

H2N S NH2
=

O
4,4' Diaminodiphenylsulphone
16

Figure 18.6 Polyamines as curing agents.

H2N-(CH2)2-NH-(CH2)2-NH-(CH2)2-NH2 + 6 H2C CH

CH OH
OH

CH2 CH2
CH
(CH2)2 N (CH2)2 N (CH2)2 N
N
CH CH2 CH2
CH2 CH2

OH CH OH
CH OH CH OH

Figure 18.7 Polyamines–epoxy curing reaction.


Polymer Matrix: Epoxy Resins 411

In general, aliphatic amines provide fast cures and are effective at room
temperature while aromatic amines are somewhat less reactive and give
products with higher heat distortion temperatures.

18.4.1.3 Calculations of the Proportion of Amines for Curing Epoxy


Resins
The desired stoichiometric quantity of amine by weight per 100 gram of
epoxy resin is calculated from the following equation:

Equivalent weight of amine


× 100
Equivalent weight of epoxy resin
Molecular weight of amine
Equivalent of amine =
Number of active hydrogen atoms
Molecular weight of epoxy resin
Equivalent of epoxy resin =
Number of epoxy groups

Example:
The molecular weight of diethylenetriamine (DETA)

H2N CH2 CH2 NH CH2 CH2 NH2

is around 103 g/mol, and the number of active hydrogen atoms of DETA
is 5.
Hence, the equivalent weight of DETA = 103/5 = 20.6.
The molecular weight of the epoxy resin (DGEBA, EPON 828) is about
340 g/mol, and the number of epoxy groups of the epoxy resin is 2.
Hence, the equivalent weight of epoxy = 340/2 = 170.
The desired stoichiometric quantity of amine by weight per 100 gram of
epoxy resin is = 20.6/170 = 0.121.
In accordance with the above equations, therefore, the stoichiometric
DETA/epoxy ratio is = (103/5)/(340/2) = 12/100 (w/w).

18.4.1.4 Special Amines


There are curing agents that have less than three hydrogen atoms available
for reaction with epoxy groups. Examples are diethanolamine (18), piperi-
dine (19), and dimethylaminopropylamine (20) (Figure 18.8).
412 Adhesives for Wood and Lignocellulosic Materials

H2
C
H2C CH2

H2C CH2
N
NH(CH2CH2OH)2 H
Diethanolamine Piridine
18 19

(CH3)2N-CH2-CH2-CH2-NH2
Dimethylaminopropylamine
20

Figure 18.8 Special amines.

These curing agents react in two steps:

(1) Firstly, the active hydrogen atoms of primary and second-


ary amines react in a manner similar to those present in
polyfunctional amines.
(2) The resulting tertiary amine is sufficiently reactive to ini-
tiate polymerization of epoxy groups as a catalytic curing
agent described earlier.

18.4.1.5 Acid Anhydrides


Cyclic anhydrides widely employed as curing agents for epoxy resins are as
follows: maleic anhydride (21), dodecenylsuccinic anhydride (22), endometh-
ylenetetrahydrophthalic anhydride (23), hexahydrophthalic anhydride (24),
phthalic anhydride (25), and pyromellitic dianhydride (26) (Figure 18.9).

18.4.1.6 Anhydride Curing Mechanism


The interaction between an acid anhydride and epoxy resin is complex. In
general, two main types of reaction occur:

(a) Opening of the anhydride group to form carboxy groups


(b) Opening of the epoxy ring

(i) Reaction of the anhydride group:


The anhydride groups may open to form one or two carboxy groups by

(a) Water (traces of which may be present in the system) or


Polymer Matrix: Epoxy Resins 413

O
CH C CH3 CH3 CH3
O
O
CH3 CH2 CH2 CH CH2 C=CH C CH C
CH C
O
O CH3
CH C
Maleic anhydride Dodecenylsuccinic anhydride O
21 22

H O
C H O
HC C C
CH C
O H 2C CH
CH2
O
HC CH C
H 2C CH C
C
CH3 H O C
H O

Endomethyl enetetrahydrophthalic anahydride Hexahydrophthalic anhydride


23 24

O O O
=

=
=

C C C
O
C O O
C C
=

O
=
=

O O
Phthalic anhydride
25 Pyromelliticdi anhydride
26

Figure 18.9 Anhydrides for epoxy resin curing.

(b) Hydroxyl groups, which are either present as pendant


groups in the original resin or which may be produced by
reaction (iia) below:
O=

O=

(a) C O + H2O C OH
C C OH
=O

O=

OH
O=

O=

C O C OH
(b) + CH2 CH CH2
C C O
O=

O=

CH2 CH CH2
414 Adhesives for Wood and Lignocellulosic Materials

(ii) Reaction of the epoxy group


The epoxy group may be opened by reaction with (a) carboxy groups
[formed by reaction (1a,b)] or (b) hydroxyl groups [which may be present
as pendant groups in the original resin or which may be produced by reac-
tion (iia)].
OH
O
O=

O=
C OH C O CH2 CH
(a) + H2C CH
C O C O
O=

O=
CH2 CH CH2 CH2 CH CH2

OH

OH O CH2 CH
O

(b) CH2 CH CH2 + H2C CH CH2 CH CH2

The reaction between epoxy groups and anhydride is slow and hence
addition of a catalyst, usually a tertiary amine, accelerates the reaction by
the following mechanism:
O
O=

+ H2C CH
O=

C O NR3
+ R3N C
C C O-
O=

O=
O=

C O
O=

+
C NR3 C
O=

+
O=

C O CH R3 NR3
CH2 C
O=

C O CH2 CH
O-
O=

CO O
CO-O-

18.5 Other Types of Epoxy Resins


18.5.1 Epoxidized Novolac
Novolac resins derived from phenol or substitute phenols can be allowed to
react with epichlorohydrin to produce epoxidized novolacs such as epoxy
phenol novolacs (EPN) and epoxy cresol novolacs (ECN) (Figure 18.10).
Polymer Matrix: Epoxy Resins 415

H2C H2C H2C


O O O
HC HC HC

CH2 CH2 CH2

O O O
CH2 CH2

Figure 18.10 Epoxidized novolac.

These are highly viscous or solid resins and possess high temperature and
chemical resistance but low flexibility [5].
High epoxy resin functionality and high cured Tg characterize these
resins and differentiate them from difunctional bisphenol A resin. Phenol
novolacs are highly viscous liquids, while cresol novolacs are typically
solids at room temperature. They are of general interest because excellent
temperature performance can be achieved at a relatively modest cost.

18.5.2 Tetrabromo Bisphenol A Epoxy Resins


An important variant is the epoxy resins produced from tetrabromo
bisphenol A.
These brominated resins are used to impart flame retardancy to the
final product and are commonly used in electrical applications. A variety
of products are available with wide ranges of both bromine content and
molecular weight. This category of resins ranges from pure diglycidyl ether
of tetrabromo bisphenol A to high-molecular-weight analogs similar to
those available with standard bisphenol A.
Flame retardants are either additive or reactive. Reactive flame retar-
dants are added during the polymerization process and become an integral
part of the polymer. TBBPA is mainly used in epoxy resins on the market as
a reactive flame retardant. It has to be pointed out that halogenated flame
retardants in particular are often toxic or even carcinogenic by themselves.
Environmental persistence and the ability to bio-accumulate add more
concerns [6].
Recently, researchers have synthesized epoxy resin by incorporat-
ing tris(2-hydroxypropyl)borate (THPB) and OPOSS (octaminophenyl
416 Adhesives for Wood and Lignocellulosic Materials

polyhedral oligomeric silsesquioxane) having boron and silicon in the cur-


ing system. It was found that there was an increase in the limited oxygen
index value of the epoxy resin with an improved heat-resistant layer and a
decrease in the release of flammable gas [5].

18.5.3 Epoxidized Vegetable Oils


There is a growing interest in the use of vegetable oils, which are sustain-
able resources needed for transition to green economy. Vegetable oils are
triglycerides of saturated and unsaturated fatty acids. The double bonds
present in the fatty acid moieties of vegetable oils can be epoxidized to
form epoxidized vegetable oils.
Epoxidation of double bonds in the oil is carried out by peracids. During
the epoxidation of double bonds, an oxygen atom is transferred from the
peracid to the C=C double bond, thus forming an oxirane ring. The epox-
idation agent of choice is m-chloroperbenzoic acid (mCPBA) (Figure
18.11).
Example
Other peracids used are peroxybenzoic acid, peroxyacetic acid, and per-
fluoroacetic acid. The latter is prepared in situ from acetic acid and hydro-
gen peroxide. The reactivities of the peroxy-acids are given in Figure 18.12.
The above method is employed for the epoxidation of the double bonds
in vegetable oils and natural rubber. Epoxidation of natural rubber gives
epoxidized natural rubber (ENR). The reaction of alkenes with peracids is
often called the Prilezhaev (Prileschajew) reaction [7]. Prileshajev epoxida-
tion consists of the formation in situ of a peracid by the action of hydro-
gen peroxide on organic acid (acetic or formic) in the presence of a strong
Bronsted acid (H2SO4 typically).
O=

O
Cl
C CHCl3
CH2 CH=CH2 + OOH CH3 CH2 CH CH2
CH3

m-Chloroperbenzoic acid +
mCPBA
O=

Cl
C
OH

m-Chlorobenzoic acid

Figure 18.11 Peroxy-acids for epoxidizing unsaturated compounds.


Polymer Matrix: Epoxy Resins 417

O=
O=

O=
C Cl C
C
OOH OOH
CH3 OOH CF3

Peracetic acid m-Chloroperbenzoic acid Perfuoroacetic acid


mCPBA

Increasing reactivity

Figure 18.12 Reactivity of peroxy-acids.

A greener epoxidation method is using lipase-catalyzed chemo-


enzymatic oxidation. This strategy presents several advantages com-
pared to the Prileshajev process. Typical experimental l conditions are
given by the work of Tellez et al. [8].

18.5.4 Epoxidized Natural Rubber


Preparation
ENR is prepared by performic epoxidation. Concentrated NR is diluted
with distilled water to have 20% DRC. Nonionic surfactant (Emulwin W or
Terric 16A-16) is gradually added to maintain the stability of latex. Formic
acid (0.75 mole equiv. to isoprene unit) and 30% H2O2 (0.75 mole equiv.
to isoprene unit) are then slowly dropped in while the mixture is heated
to 50°C. The mixture solution is stirred at 50°C for 2–9 h. After the reac-
tion, the mixture is cooled down to room temperature and sodium sulfite
is added to remove the excess H2O2 followed by neutralization with 25%
ammonia solution. MeOH is then added to precipitate the product. The
white solid bulk obtained is washed with water several times, rinsed with
MeOH, and dried under reduced pressure.

18.6 Green or Sustainable Epoxy Matrix


Bisphenol A is reported to have severe impact on human health as it is able
to alter the human immune and reproductive system behavior within brain
chemistry. Bisphenol A also has estrogenic properties. Thus, several efforts
are being made to replace the traditional DGEBA biobased epoxy mono-
mers or oligomers with comparable high-performance materials.
Natural polyphenols (condensed tannins, lignin) have potential to sub-
stitute bisphenol A. Biobased epoxy resins were synthesized from catechin
418 Adhesives for Wood and Lignocellulosic Materials

molecule or condensed tannins, or one of the repetitive units in natural


flavonoid biopolymers with improved thermal and mechanical proper-
ties, and also from nonfunctionalized gallic acid (from the hydrolysis of
gallotannins). Catechin as a model compound of procyanidin-type con-
densed tannins has been epoxidized either by reaction with epichlorohy-
drin or by alkylation with unsaturated halogenated compound followed
by its oxidation [9, 10]. Benyahya et al. have proposed the direct use of
condensed tannins from green tea for the synthesis of aromatic biobased
epoxy oligomers [11]. More recently, two studies by Jahanshahi et al. [12,
13], one on the formation and testing of commercial polyflavonoid tannin
epoxy resins, were also published, the second of which was of particular
interest as it involved an epoxidized tannin–acrylic polymer that allowed
rapid curing of the epoxy tannin without having to use a hardener [12, 13].
The tannin-based epoxy acrylate resin was prepared from glycidyl ether
tannin (GET) and acrylic acid by reaction between GET and acrylic acid
in the presence of a catalyst and hydroquinone and tested successfully for
block shear strength [13].
Hydrolyzable tannins can also be used to prepare epoxy oligomers.
Thus, gallic acid epoxidation with epichlorohydrin was first reported
by Tomita and Yonezawa [14]. The addition occurs on both the carbox-
ylic acid group and also on at least one phenol group in the presence
of an ammonium-type phase transfer catalyst in anhydrous medium.
Nouailhas et al. also recently proposed another synthesis pathway leading
to epoxy prepolymers from gallic acid allylation followed by double bond
oxidation [10].
Cardanol is a further biocompound that has been tried for bioepoxy
synthesis. A patent exists on the use of commercial epoxidized cardanol
from Weller et al. [15], suggesting the use of cardanol to prepare biobased
epoxy cross-linked polymers. Epoxy monomers and oligomers were also
prepared using cardanol by enzymatic double bond epoxidation with
lipase and acetic acid in toluene [16–19].
Only few publications deal with the modification of lignin for use as
epoxides [20–23]. Three processes have been described in the literature to
prepare epoxidized lignin [24]. The main one is based on the oxypropyla-
tion of lignin to yield a lignin with more accessible hydroxyl groups. These
hydroxyl groups are then converted into epoxide groups by reaction with
epichlorohydrin, and cross-linking occurs by addition of diamines [24].
A variety of epoxide prepolymers that have been obtained from biobased
polyols like glycerol and sorbitol are commercially available [25, 26]. These
are mainly of the aliphatic glycidyl ether type. Equally, there are reports
on the synthesis of epoxide prepolymes from glycerin [27, 28]. Diglycidyl
Polymer Matrix: Epoxy Resins 419

ethers of isosorbide as well as oligo-glycidyl ethers of isosorbide have also


been synthesized, leading to oligomers of epoxidized isosorbide by reac-
tion with epichlorohydrin [29–34]. Lactic acid also seems usable to pre-
pare either linear or hyperbranched epoxide prepolymer, depending on the
experimental conditions used [35]. Finally, epoxy monomers from sugars
can also be prepared. They are usually obtained by double bond oxidation,
needing the use of strong oxidative agents in large excess [28].
Furthermore, the industry can use epichlorohydrin derived from
biobased glycerol, which is an inexpensive and abundant source from soap
industry. Epichlorohydrin from such a source, named biobased epichloro-
hydrin, is now commercially available.
Researchers are also interested in epoxidized plant oils and fatty acids
as alternatives because they display interesting and comparable thermal
properties to commercial DGEBA. ENR is another interesting biobased
material that has been employed as toughening agent with petroleum-
derived epoxy resin [36–40].

18.7 Epoxy-Matrix-Based Biofiber Composites


Natural-fiber-based epoxy resin composites provide a unique combina-
tion of great versatility, high performance, and production at cheaper rates
compared to synthetic thermoplastics. Study also shows that the physical
and mechanical properties of epoxies are improved in the greater majority
of cases by the incorporation of natural fibers.
Jawaid and Abdul Khalil developed oil palm-jute-reinforced epoxy
matrix composites and studied their dynamic mechanical properties
and thermal degradation [41]. Thermogravimetric analysis indicated an
increase in thermal stability when jute fibers are incorporated along with
oil palm empty fruit bunches (EFB). A DGEBA-type epoxy resin was used
together with a reactive polyamide as curing agent.
Oil palm EFB and woven jute (Jw) fiber-reinforced epoxy matrix com-
posites were prepared by Jawaid, employing a hand lay-up method [41].
An epoxy resin was used as matrix, and trilayer hybrid composites of EFB/
woven jute/EFB and woven jute/EFB/woven jute were prepared by keeping
the EFB/woven jute fiber weight ratio constant at 4:1. The chemical resis-
tance properties of the EFB/woven jute/EFB and woven jute/EFB/woven
jute fiber-reinforced hybrid composites were studied by using three acids,
three alkalis, and three solvents. It was found that all the composites are
resistant to various chemicals. It was observed that a marked reduction in
void content of the composites occurred with the hybridization of oil palm
420 Adhesives for Wood and Lignocellulosic Materials

EFB with woven jute fiber. A density and moisture content study demon-
strated that density and moisture content property was enhanced in hybrid
composites. The hybrid composites showed lower impact properties than
pure EFB composites. The Izod impact strength of EFB/woven jute/EFB
and woven jute/EFB/woven jute hybrid was found to be 72.4 and 57.9 J/m,
respectively. The impact fracture surface of the composites was studied
with the help of scanning electron microscopy.
Ojinmah et al. presented the results on the effect of a semi-nanofiller from
rice husk on the mechanical properties of ENR [42]. This was studied and
compared with carbon black-filled epoxidized natural rubber composites.
The particle size distribution of the milled rice husk was determined using
optical microscopy and digital imaging technique. The semi-nanopowder
obtained by ball milling of the rice husk was characterized for pH, iodine
value, cellulose content, lignin content, hemicellulose content, moisture
content, bulk density, and loss on ignition. The result of the physicochem-
ical analysis showed excellent values that compared favorably with other
literature reports. Optical microscopy and digital imaging of the rice husk
nanofiller showed a spherical particle of average size 221 nm. The results
of the mechanical testing presented similar trends and close values for
both carbon black-filled ENR and rice husk semi-nanofilled ENR at filler
loading in the range 10–50 phr. Consequently, rice husk nanofiller can be
a good substitute for carbon black for products where cost, compression
set, and rebound resilience properties are critical.

References
1. Saunders, K.J., Organic Polymer Chemistry, Chapman and Hall, London,
1973.
2. Boyle, M.A., Martin, C.J., Neuner, J.D., Epoxy Resins, ASM Handbook, vol. 21,
pp. 78–89, Composites, ASM International, Almere, The Netherlands, 2001.
3. H.A. Newey and E.C. Shokal, Glycidyl ether composition and method of
using same. US Patent 2,575,558, 1951, Assigned to Shell Development Co.
4. E.C. Shokal, H.A. Newey, T.F. Bradley, Nitrogen containing polyethers
and process for curing glycidyl polyethers to resinous products. US Patent
2,643,239, 1953, Assigned to Shell Development Co.
5. Shirude, S.R., Shambharkar, S.Y., Bhosale, H.B., Patil, V.U., Patil, A.G., A sur-
vey on epoxy resins. Int. J. Innov. Res. Sci. Eng. Technol., 6, 14861–14867,
2017.
6. Rakotomalala, M., Wagner, S.S., Döring, M., Recent developments in halo-
gen free flame retardants for epoxy resins for electrical and electronic appli-
cations. Materials, 3, 4300–4327, 2010.
Polymer Matrix: Epoxy Resins 421

7. Prileschajew, N., Oxydation ungesättigter Verbindungen mittels organischer


Superoxyde. Chem. Ber., 42, 4, 4811–4815, 1909. http://www.chemgapedia.
de/vsengine/vlu/vsc/en/ch/2/vlu/oxidation_reduktion/epxo_per_1.vlu.
html.
8. Tellez, G.L., Vigueras-Santiago, E., Hermandez-Lopez, S., Characterization
of linseed oil epoxidized at different percentages. Superf. Vacio, 22, 5–10,
2009.
9. B. Boutevin, S. Caillol, C. Burguiere, S. Rapior, H. Fulcrand, H. Nouailhas,
Novel method for producing thermosetting epoxy resins. Patent WO
2010136725, 2010. Assigned to Centre National de la Recherche Scientifique,
France.
10. Nouailhas, H., Aouf, C., Le Guerneve, C., Caillol, S., Boutevin, B., Fulcrand,
H., Synthesis and properties of biobased epoxy resins. Part 1: Glycidylation
of flavonoids by epichlorydrin. J. Polym. Sci., Part A: Polym. Chem., 49,
2261–2270, 2011.
11. Benyahya, S., Aouf, C., Caillol, S., Boutevin, B., Fulcrand, H., Functionalized
green tea tannins as phenolic prepolymers for bio-based epoxy resins. Ind.
Crops Prod., 53, 296–307, 2014.
12. Jahanshahi, S., Pizzi, A., Abdolkhani, A., Doosthoseini, K., Shakeri, A., Lagel,
M.-C., Delmotte, L., MALDI-TOF and 13C-NMR and FT-MIR and strength
characterization of glycidyl ether tannin epoxy resins. Ind. Crops Prod., 83,
177–185, 2016.
13. Jahanshahi, S., Pizzi, A., Abdolkhani, A., Shakeri, A., Analysis and testing of
bisphenol-A-free bio-based tannin epoxy-acrylic adhesives. Polymers, 8, 143,
2016.
14. H. Tomita and K. Yonezawa, Epoxy resin and process to prepare the same.
European Patent EP 0095609, 1983. Assigned to Kanegafuchi Chemical Ind.
15. C.G. Weller, E.J. Siebert, Z. Yang, R.K. Agarwal, W.E. Fristad, B.D. Bammel,
Autodeposition compositions for polymeric coatings of reduced gloss, good
corrosion resistance and uniform appearance. Patent WO 2003026888,
2003.
16. Devi, A. and Srivastava, D., Enzymatic epoxidation and polymerization
of cardanol obtained from a renewable resource and curing of epoxide-
containing polycardanol. J. Appl. Polym. Sci., 102, 2730–2737, 2006.
17. Unnikrishnan, K.P. and Thachil, E.T., Synthesis and characterization of
cardanol-based epoxy systems. Des. Monomers Polym., 11, 593–607, 2008.
18. He, J., Xu, S., Xu, X., Study on the anacardiol modified epoxy resin. Tuliao
Gongye, 29, 5–11, 1999.
19. Nieu, H.N., Tan, T.T.M., Huong, N.L., Epoxy–phenol–cardanol–formalde-
hyde systems: Thermogravimetry analysis and their carbon fiber composites.
J. Appl. Polym. Sci., 61, 2259–2264, 1996.
20. Simionescu, C.I., Rusan, V., Macoveanu, M.M., Cazacu, G., Lipsa, R., Vasile,
C., Stoleriu, A., Ioanid, A., Lignin/epoxy composites. Compos. Sci. Technol.,
48, 317–323, 1993.
422 Adhesives for Wood and Lignocellulosic Materials

21. Feldman, D., Lacasse, M., Beznaczuk, L.M., Lignin–polymer systems and
some applications. Prog. Polym. Sci., 12, 271–276, 1987.
22. Tai, S., Nagata, M., Nakano, J., Migita, N., Studies on utilization of lignin. IV.
Epoxidation of thiolignin. Mokuzai Gakkaishi, 13, 102–107, 1967a.
23. Tai, S., Nakano, J., Migita, N., Studies on utilization of lignin. V. Adhesive
from lignin epoxide. Mokuzai Gakkaishi, 13, 257–262, 1967b.
24. Koike, T., Progress in development of epoxy resin systems based on wood
biomass in Japan. Polym. Eng. Sci., 52, 701–717, 2012.
25. Shibata, M. and Nakai, N., Preparation and properties of biocomposites
composed of bio-based epoxy resin, tannic acid, and microfibrillated cellu-
lose. J. Polym. Sci., Part B: Polym. Phys., 48, 425–433, 2010.
26. Takada, Y., Shinbo, K., Someya, Y., Shibata, M., Preparation and properties of
bio-based epoxy montomorillonite nanocomposites derived from polyglyc-
erol polyglycidyl ether and -polylysine. J. Appl. Polym. Sci., 113, 479–484,
2009.
27. Acierno, D., Russo, P., Savarese, R., Oxidation of maltitol and sorbitol as a way
to bioepoxies, Polymer Processing Society, PPS-24, Salerno, Italy, 2008.
28. Sachinvala, N.D., Winsor, D.L., Menescal, R.K., Ganjian, I., Niemczura, W.P.,
Litt, M.H., Sucrose-based epoxy monomers and their reactions with dieth-
ylenetriamine. J. Polym. Sci., Part A: Polym. Chem., 36, 2397–2413, 1998.
29. A. East, M. Jaffe, Y. Zhang, L. Catalani, Thermoset epoxy polymers from
renewable resources such as anhydrosugars. Patent WO 2008147473, 2008a.
Assigned to New Jersey Techn. Inst.
30. A. East, M. Jaffe, Y. Zhang, L. Catalani, Ethers of bisanhydrohexitols. US
Patent 20080021209, 2008b. Assigned to New Jersey Techn. Inst.
31. Achet, D., Delmas, M., Gaset, A., Biomass as a source of chemicals. VI.
Synthesis of new polyfunctional ethers of isosorbide in solid–liquid hetero-
geneous mixtures. Biomass, 9, 247–254, 1987.
32. Feng, X., East, A.J., Hammond, W., Jaffe, M., Sugar-based chemicals for envi-
ronmentally sustainable applications, in: Contemporary Science of Polymeric
Materials, vol. 1061, L. Korugic-Karasz (Ed.), pp. 3–27, ACS Symp. Ser,
Oxford University Press, 2010.
33. Feng, X., East, A.J., Hammond, W.B., Zhang, Y., Jaffe, M., Overview of
advances in sugar-based polymers. Polym. Adv. Technol., 22, 139–150, 2011.
34. I. Wiesner, J. Kriz, V. Bruthans, J. Kolinsky, Glycidyl ethers of phenols and
polyphenols for synthesis of epoxy resins. Patent CS 113977, 1965.
35. Ma, S., Liu, X., Jiang, Y., Tang, Z., Zhang, C., Zhu, J., Bio-based epoxy resin
from itaconic acid and its thermosets cured with anhydride and comono-
mers. Green Chem., 15, 245–254, 2013.
36. Kumar, R.N., Kong, W.C., Abusamah, A., UV radiation curing of surface
coatings based on ENR-cycloalipahatic di-epoxide-glycidyl methacry-
late system by cationic photointiators—Optimization of process variables
through response surface methodology. J. Coat. Technol., 71, 896, 79–88,
1999.
Polymer Matrix: Epoxy Resins 423

37. Kumar, R.N. and Mehnert, R., Electron beam curing of the system consisting
of cycloaliphatic diepoxide-epoxidized natural rubber–glycidyl methacry-
late in presence of cationic initiators. Macromol. Mat. Eng., 286, 449–455,
2001a.
38. Kumar, R.N., Mehnert, R., Scherzer, T., Bauer, F., CP MAS NMR spectros-
copy to the characterization of UV/EB cured epoxidized natural rubber
blends. Macromol. Mat. Eng., 286, 598–604, 2001b.
39. Kumar, R.N., Woo, C.K., Abusamah, A., UV curing of surface coating sys-
tem consisting of cycloaliphatic di-epoxide–ENR–glycidyl methacrylate by
cationic photointiators—Charaterization of the cured film by FTIR spectros-
copy. J. Appl. Polym. Sci., 73, 1569–1577, 1999.
40. Rosli, W., Daud, W., Kumar, R.N., Salleh, M.Z., Mahmood, M.H., Studies on
cationic curing of cycloaliphatic diepoxide-epoxidized palm oil (EPO) for
surface coatings. Eur. Polym. J., 39, 593–600, 2003.
41. Jawaid, M. and Abdul Khalil, H.P.S., Effect of layering pattern on the dynamic
mechanical properties and thermal degradation of oil palm–jute fibres rein-
forced epoxy hybrid composites. Bioresources, 63, 2309–2322, 2011.
42. Ojinmah, N., Uchechukwu, T.O., Ezeh, V.O., Ogbobe, O., Studies on the
effect of rice husk semi-nano filler on the mechanical properties of epoxi-
dized natural rubber composites. Eur. J. Adv. Eng. Techn., 4, 164–171, 2017.
19
Polymer Matrix: Polyethylene

19.1 Introduction
Polyethylene (PE) is one of the most widely used thermoplastic in the
world. PEs are classified into several different categories. The main forms
of PE are High-density polyethylene (HDPE), low-density polyethylene
(LDPE), medium-density polyethylene (MDPE), ultrahigh molecu-
lar weight polyethylene (UHMWPE), linear low-density polyethylene
(LLDPE), and very low-density polyethylene (VLDPE). These are divided
based on density and branching. Generally, the most used PE grades are
HDPE, LDPE, and MDPE [1, 2].
PE being the major commodity plastics, it is used as the matrix in bio-
composites. PE-based biocomposites have been extensively used in both
structural and nonstructural applications due to low cost, recyclability, and
processing advantages.
PE is produced from ethylene monomer. Ethylene is generally pro-
duced via steam cracking of crude oil derivatives. Hence, improving
polymer production techniques with the objective to reduce manufac-
turing cost continues to be an area of research, development, and process
improvement [3].
Figure 19.1 shows schematic structures for the three PEs, with the main
features exaggerated.
During the 1950s, three different catalysts were invented to produce
polyethylene at low pressure and temperature in contrast to the previously
employed high-pressure process for the production of LDPEs. The new
ethylene polymers had densities in the region of 960 kg/m3, and became
known as HDPE.

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (425–440) © 2019
Scrivener Publishing LLC

425
426 Adhesives for Wood and Lignocellulosic Materials

(a) (b) (c)

LDPE LLDPE HDPE

Figure 19.1 Schematic molecular structure. (a) LDPE, (b) LLDPE, and (c) HDPE [2].

19.2 High-Pressure Process


PE was initially produced commercially under very high pressure
(~3000  atm). The polymerization reaction was initiated by free radicals.
The product was LDPE. Because of the high operating pressure, the capital
investment cost for equipment to withstand the high pressure was very
high. Further, the operating costs of the process such as power to com-
press reactants to the high operating pressure were also high. However,
variants of this technology as licensed by ExxonMobil Chemical Co. and
Basell Polyolefins continue to be used worldwide today [3]. For instance, a
very flexible and branched LDPE, typically in the range of 910–940 kg/m3,
is manufactured commercially in a high-pressure free radical polymeriza-
tion reactor in the presence of a mixture of initiators (e.g., peroxides, oxy-
gen, and azo compounds). Practically, almost all LDPE is manufactured in
stirred autoclaves or in tubular reactors at high pressures [4].

19.3 Low-Pressure Processes—Catalysts


for Polymerization
Technologies were subsequently developed to produce PE under lower
operating pressures. Special catalysts were invented, which allowed pro-
duction of PE by coordination polymerization in slurry and solution phase
reactors.
Polymer Matrix: Polyethylene 427

These reactors operate at a lower pressure compared to the earlier


high-pressure free radical process. The slurry-phase “loop reactor” by
Chevron Phillips Chemical Co. operates at 40 atm. Further, the use of
co-monomers such as 1-butene and 1-hexene allowed the production
of LLDPE and HDPE. The special catalysts cited above are described
below.

19.3.1 Ziegler–Natta Catalysts


Ziegler-Natta catalysts have gained reputation the world over in the
polymerization industry for the production of PE and PP [5, 6]. These
catalysts have been categorized into two major classes based on their sol-
ubility [7]:

(i) Heterogeneous catalysts: These catalysts consist of an


organic metallic compound of a metal from group I–III
of the periodic table such as triethylaluminum [Al(C2H5)3]
and a transition metal compound from groups IV–VIII
such as titanium tetrachloride [8–12].
(ii) Homogeneous catalysts are based on metallocenes. They
consist of minute particles of positively charged metal
ions sandwiched between two cyclopentadienyl anions,
which have five atoms per ring. A cyclopentadienyl anion
is derived from cyclopentadiene. They are also known as
single-site catalysts because they have only one active site
per catalyst particle, which are all identical. Metallocene
catalysts were discovered in the 1950s. They are “sandwich
compounds” in which a π-bonded metal atom is situated
between two aromatic ring systems. Despite the early dis-
covery, it took more than 20 years before technology was
developed that allowed the economical large-scale pro-
duction of metallocene polyolefins [13–15]. Metallocenes,
in combination with the conventional aluminum alkyl
co-catalysts used in Ziegler systems, are indeed capable of
polymerizing ethylene, but only at a very low activity. Only
with the discovery and application of methylaluminoxane
(MAO) (post-metallocenes) in the Institute for Technical
and Macromolecular Chemistry, University of Hamburg,
in 1977 was it possible to enhance the activity, surpris-
ingly, by a factor of 10,000. Therefore, MAO plays a crucial
part in catalysis with metallocenes [16].
428 Adhesives for Wood and Lignocellulosic Materials

19.4 Production of PE
19.4.1 Solution Process
The ethylene is dissolved (25%) in a diluent such as cyclohexane and
charged into the reactor at ca. 10 MPa. The reaction is carried out at a
temperature range of 200–300°C and a residence time is ca. 2 min. Ninety-
five percent of the feed is converted to PE in the reactor. The catalyst is a
mixture of VOCl3 and TiCl4 activated by an aluminum alkyl. The PE solu-
tion leaving the reactor is treated with a deactivating agent and the mix-
ture then passes through a bed of alumina where the deactivated catalyst
residues are adsorbed. Two depressurization stages follow, in which sol-
vent and unreacted monomers are volatilized [2]. Solution processes have
been developed by various companies including Du Pont, Dow, DSM, and
Mitsui for the manufacture of LLDPE or HDPE/LLDPE on a swing basis.

19.4.2 Slurry Process


Polymerizations may be conducted in diluents in which PE is insoluble
at the process temperature. Such processes are termed slurry (or suspen-
sion) processes. Diluents must be inert toward the catalyst system and are
usually saturated hydrocarbons such as propane, isobutane, and hexane.
Slurry processes typically operate at temperatures from about 80 to 110°C
and pressures of 200–500 psig. PE precipitates as it is formed, resulting in
a suspension of polymer in diluent. The catalysts most commonly used
in slurry processes are chromium on silica or supported Ziegler–Natta
catalysts.
Polymerizations that use supported chromium (Phillips) catalysts are
conducted predominantly in slurry processes. The slurry process originally
developed by Phillips Petroleum (now Chevron Phillips) has been called
the “particle form loop slurry process” and the “slurry loop reactor pro-
cess” for production of HDPE and LLDPE (10). Hexene-1 is most often
used as co-monomer for LLDPE in the Phillips process.
Another well-known slurry (suspension) process was developed by
Hoechst in Germany in the mid-1950s. Hoechst was the first licensee
to use the catalyst and process developed by Karl Ziegler for producing
low-pressure linear PE in 1955. Hoechst was eventually absorbed into the
company that is known today as LyondellBasell.
The Hoechst slurry process was improved over the years and has evolved
into what is now called the Hostalen process. Hostalen is a slurry-
cascade process that is capable of producing a wide range of molecular
Polymer Matrix: Polyethylene 429

weight distributions of HDPE. The modern Hostalen process employs


two continuous stirred tank reactors that can be run in series or in paral-
lel to produce unimodal and bimodal HDPE [17].

19.4.3 Gas Phase Fluidized Bed Reactor


Gas phase ethylene polymerizations are typically conducted in fluid-
ized beds at pressures of 200–500 psig and temperatures of 80–110°C.
Gas phase processes for PE were developed originally by Union Carbide
(now Dow) and later by Naphtachimie (now INEOS). These processes
are called the Unipol and Innovene processes, respectively. The pre-
dominant catalyst used in each process is of the supported Ziegler–Natta
type, though the catalysts are produced by completely different chem-
istries. Most of the PE made in gas phase processes employs Ziegler–
Natta catalysts. There are, however, a few instances where supported
chromium and single-site catalysts are used. The number of fluidized
bed plants increased after the announcement of Union Carbide’s Unipol
process for LLDPE in 1977. These plants were dual-purpose plants
(“swing plants”) with the ability to produce either LLDPE or HDPE
according to demand. The fluidized bed process can produce a very wide
range of melt flow indexes (MFIs) and densities. Butene was originally
used as the co-monomer for LLDPEs, which was followed later by hex-
ene as high-performance copolymers. BP uses 4-methylpentene for its
high-performance LLDPEs.
The reactor has a typically designed shape so as to reduce the exit gas
velocity. The particles can therefore fall back into the fluidized bed without
being carried over. The gas enters the reactor through a distributor plate,
which provides an even distribution of gas and must also prevent powder
falling through when the gas flow is stopped. The fluidized bed functions
as a continuous stirred-tank back-mix reactor.
Reaction temperature is 80–100°C and the pressure is in the range of
0.7–2.0 MPa. The polymer is removed via cyclone, from which residual
monomers are recovered and recompressed [2].

19.5 Compatibilizers
Compatibilizers such as maleated ethylene, maleated propylene, and a
few acrylic-grafted linear polymers are also produced for applications in
biofiber-reinforced polymer matrix composites to enhance the adhesion
between the fiber and polymer matrix [18–22].
430 Adhesives for Wood and Lignocellulosic Materials

19.6 Relevant Property of PE


19.6.1 Melt Flow Index
MFI is a measure of how many grams of a polymer flow through the die
in 10 min. The test is performed at a given temperature depending on the
plastic. The force used to push the plastic through the system is supplied by
a weight that sits on top of a ram.
The thermoplastics materials industry uses the MFI as basic tool for
quality control and acceptance of incoming products. PEs are routinely
characterized by their density and MFI. The MFI test was originally chosen
for LDPE to give a measure of the melt characteristics under conditions
related to its processing. It is carried out by applying a standard force to a
piston and measuring the rate of extrusion (in g/10 min) of the PE through
a standard die. This is commonly used for polyolefins (polyethylene HDPE,
LDPE, LLDPE, and polypropylene PP). ASTM D1238 and ISO 1133 give
the details [23, 24].

19.7 Treatment and Functionalizing of Biofibers


and Matrix Materials
This has been dealt with in detail in Chapter 16 (“Modification of Natural
Fibers and Polymeric Matrices”).

19.8 Biocomposites Based on PE


19.8.1 Kenaf-Based Biocomposites
En et al. reported that the high tensile properties of kenaf fibers (KFs) can
reinforce the polyolefin matrices considerably [25]. They also considered
that the presence of KFs may affect the crystallization kinetics of PE. It is
known that the reinforcements may act as the nucleation sites, which may
consequently promote crystallization. The effect of KFs on the crystalliza-
tion kinetics of the PE matrix was examined by DSC.
Ku et al. made HDPE-based KF composites using a twin-screw extruder
[26]. Foaming agent was incorporated in the formulation to decrease the
weight of the polyolefin/KF composites. Thermal expandable microcap-
sule was considered as the best choice as a foaming agent for polyolefin/KF
composites due to its easy process control compared to other chemical and
Polymer Matrix: Polyethylene 431

mechanical foaming processes. Three types of PE-g-MAH, as a compatibi-


lizer, with different grafting degrees of maleic anhydride (MAH) were pro-
duced by reactive extrusion. The chemical structure and grafting degree of
PE-g-MAH were identified by infrared spectrometry and back titration.
The optimum grafting degree of MAH for PE-g-MAH was confirmed
by measuring the mechanical properties and specific gravity. It might be
caused by a difference in the interface properties. The interface adhesion
properties of the composite were determined by contact angle measure-
ments and SEM images of the fracture after the tension test.
Behjat et al. reported on the biodegradability of several blends of cel-
lulose derived from bast part of kenaf (Hibiscus cannabinus L.) plant,
with different thermoplastics, LDPE and HDPE [27]. The blends were
prepared by a melt blending machine. Polyethylene glycol (PEG) was
used as plasticizer. Biodegradability of these blends was measured using
soil burial test in order to study the rates of biodegradation of these
polymer blends. It was found that the cellulose/LDPE and cellulose/
HDPE blends were biodegradable to a considerable extent. The bio-
composites with high content of cellulose had higher degradation rate.
In addition, biodegradability of the biocomposites made up using PEG
was superior to those of the biocomposites fabricated without PEG, due
to the improved wetting of the plasticizer in the matrix polymer. The
results were also supported by scanning electron microscopy (SEM).
Biodegradability of polymers such as 3-hydroxyl butyrate and its copo-
lymers and aliphatic polyesters is indeed good, but their higher cost
compared to petroleum-based commodity plastics prevents them from
a larger commercial usage and finds applications only in niche sectors
[28]. Synthetic thermoplastics blended with lignocellulosic fibers are
cost-wise more attractive.
Ismail et al. made composites based on HDPE/soya powder/kenaf core
fiber by incorporation of kenaf core fiber at different loadings into HDPE/
soya powder matrix [29]. An internal mixer (Haake PolyLab) was employed
at a temperature of 180°C and a rotor speed of 50 rpm. The compatibilizer
[maleated polyethylene (MAPE)], was used to improve the matrix–fiber
interaction and hence the mechanical properties of the composites. The
effects of operating variables such as kenaf core fiber loading, amount of
MAPE compatibilizer on processing torque, mechanical properties, and
water absorption were studied. The results indicated that tensile strength
and elongation at break decreased with increasing filler loading, whereas
stabilization torque, tensile modulus, flexural strength, and flexural mod-
ulus exhibited an increasing trend. The presence of MAPE also enhanced
the tensile and flexural properties of the composites.
432 Adhesives for Wood and Lignocellulosic Materials

The performance of kenaf powder (KP) as filler for recycled high-


density polyethylene (rHDPE)/natural rubber (NR) thermoplastic elas-
tomer composites was investigated by Viet et al. [30]. The composites
with different filler loading were prepared in a Haake internal mixer.
Increasing KP loading in rHDPE/NR/KP biocomposites reduced the ten-
sile strength and elongation at break but increased the stabilization torque
and the tensile modulus. SEM study of fracture surface indicated that
fibrillation of rHDPE was reduced and detachment of KP from polymer
matrix was present particularly at high filler loading. These observations
were responsible for the deterioration of tensile strength and elongation
at break of rHDPE/NR/KP biocomposites. Water absorption study also
showed that the water absorption of these biocomposites increased with
increasing KP content.
Akubue et al. produced kenaf-reinforced PE composites for ballistic
protection [31]. Alkali-treated and silane-coupled non-woven mat of KFs
were cut to the required dimensions and oriented vertically and horizon-
tally in combinations with a virgin high-density polyethylene (VHDPE).
The Box-Behnken three-variable experimental design was adopted for
producing the composites.
Conditions employed were as follows: molding temperature: 160°C to
200°C; molding time: 60 to 80 min; fiber volume fraction: 10% to 30%.The
pressure for heating and cooling was controlled at 12 MPa. The responses
as tensile and flexural values were determined and optimized. The com-
posite sample of VHDPE ballistic test were blended based on the optimum
settings of temperature (200°C), molding time (80 min), and fiber volume
fraction (30%), and tested with a Jojef Magnum rifle gun. The mechanical
and ballistic properties of the composite panel of VHDPE were determined.
The studies revealed that fiber volume at 30% protected against Armour
Level Protection Class of NIJ standard level III-A for VHDPE composite.

19.8.2 Sisal-Fiber-Based Biocomposites


Sisal fiber is obtained from the leaves of the plant Agave sisalana, which
was originated from Mexico and is now mainly cultivated in East Africa,
Brazil, Haiti, India, and Indonesia [32, 33]. It is grouped under the broad
heading of the “hard fibers”, among which sisal is placed second to manila
in durability and strength [34].
Sisal fiber is a promising reinforcement for use in composites on
account of its low cost, low density, high specific strength and modulus, no
health risk, easy availability in some countries, and renewability. In recent
years, there has been an increasing interest in finding new applications for
Polymer Matrix: Polyethylene 433

sisal-fiber-reinforced composites that are traditionally used for making


ropes, mats, carpets, fancy articles, and others.
Li et al. have reviewed the developments of sisal fiber and its composites
[35]. The properties of sisal fiber itself, the interface between sisal fiber
and matrix, the properties of sisal-fiber-reinforced composites, and their
hybrid composites have been reviewed. Suggestions for future work are
also given.
Joseph et al. have published a detailed review on the research work
published in the field of sisal-fiber-reinforced polymer composites with
special reference to the structure and properties of sisal fiber, processing
techniques, and the physical and mechanical properties of the composites
[36]. PE and sisal fibers can be chemically modified to improve their com-
patibilities, i.e., to increase the hydrophobic character of the sisal fiber and
the hydrophilic character of HDPE [33, 37, 38]. Sisal was mercerized with a
NaOH solution and acetylated and the PE was oxidized with KMnO4 solu-
tion [39]. The chemically modified fibers were characterized by Fourier
transform infrared spectroscopy (FTIR) and 13C nuclear magnetic reso-
nance spectroscopy (13C NMR). The composites were prepared by extru-
sion of modified and unmodified materials containing either 5 or 10 wt%
fibers. The morphology of the obtained materials was evaluated by SEM.
The fiber chemical modification improves its adhesion with matrix, but
no benefit could be derived as a result of oxidation of HDPE. Flexural and
impact tests demonstrated that the composites prepared with modified
sisal fibers and unmodified PE improved mechanical performance com-
pared to pure PE.
Although the recycling capacity for plastics has been progressively
increased, the fraction of plastics that end up in a landfill is still very sig-
nificant. HDPE is one of the most widely used polymers, having a broad
amount of applications such as bottles, containers, and consumer goods.
Since post-consumer HDPE from bottles cannot be used again and has
high melting viscosity, HDPE becomes an interesting source of recycled
material. The use of natural fibers as reinforcement in post-consumer
HDPE can be a way to produce high-strength materials from recycled
HDPE.
Chianelli et al. investigated the mechanical properties of sisal-fiber-
reinforced recycled HDPE [40]. The work showed the usefulness of sisal
fiber as reinforcement to recycled HDPE matrix. The tensile and flexural
properties of sisal-recycled HDPE remain the same with no significant det-
riment to ultimate strength.
Sisal fibers and finely powdered HDPE were surface functionalized
using dichlorosilane under R-F plasma conditions to improve interfacial
434 Adhesives for Wood and Lignocellulosic Materials

adhesion between the two dissimilar substrates. The functionalized PE


(70%) and sisal (30%) were compounded on four different ways using a
thermokinetic mixer and injection molded into composite specimens for
test. Some improvements in mechanical properties of the composites due
to the plasma treatments were achieved. Scanning electronic microscopy
(SEM) data indicated that some compatibilization of the two plasma-
modified phases had taken place as compared to non-plasma-treated
composite [41].

19.8.3 Flax-Fiber-Based Biocomposites


Flax (Linum usitatissimum), also known as common flax or linseed, is a
member of the genus Linum in the family Linaceae. It is a food and fiber
crop cultivated in cooler regions of the world. Flax belongs to the bast fibers.
The textiles made from flax are known in the Western countries as linen,
and traditionally used for bed sheets, underclothes, and table linen. The oil
is known as linseed oil. In addition to referring to the plant itself, the word
“flax” may refer to the unspun fibers of the flax plant. The plant species is
known only as a cultivated plant and appears to have been domesticated
just once from the wild species Linum bienne, called pale flax.
Flax-fiber-reinforced plastic composites have attracted increasing inter-
est because of the advantages of flax fibers, such as low density, relatively
high toughness, high strength and stiffness, and biodegradability. Among
plastics, PE was considered as a suitable material for use as a matrix in
composites. LLDPE and HDPE were the main matrices investigated in this
research [42, 43].
Flax straw and flax fibers are renewable resource and have potential
for use in the manufacturing industry. Li et al. studied the properties of
flax-fiber-reinforced (including 58% flax shives by weight) PE (HDPE
and LLDPE) [42]. The material was processed through extrusion and
injection molding to biocomposites. Five surface modification meth-
ods on flax fibers were carried out and scanning electronic micrographs
were taken to analyze the surface characteristic. The tensile strength of
biocomposites increased and moisture absorption decreased to varying
degrees after the fiber surface modifications. Among those surface modi-
fication techniques, acrylic acid treatment showed a relatively good result
in reducing moisture absorption and enhancing tensile properties of bio-
composites. It was also found that with increased fiber content (from
10% to 30%), the tensile strength and moisture absorption of biocom-
posites increased.
Polymer Matrix: Polyethylene 435

19.8.4 Hemp-Fiber-Based Biocomposites


Hemp  fiber is obtained from the bast of the plant  Cannabis sativa  L. It
grows easily—to a height of 4 m—without agrochemicals and captures large
quantities of carbon. Production of hemp is restricted in some countries,
where the plant is confused with marijuana. 
Lu et al. prepared hemp fiber composites with recycled HDPE matrix
at various compositions ranging from 20% to 40% of fiber volume frac-
tion [44]. The fiber–matrix interface was improved using 5% by weight
NaOH-treated hemp fiber in each composite system. The surface mor-
phology and chemical composition of hemp fiber after chemical treatment
were determined by SEM and FTIR. This study indicated that hemp fiber–
rHDPE composites could achieve maximum tensile strengths of the order
of 60 MPa. Among the tested samples, the composites with 40% of fiber
volume fraction gave the best mechanical properties with regard to tensile
strength, elastic modulus, and flexural strength and modulus.
Lu et al. conducted investigations on the properties of natural fiber com-
posites (NFCs) produced from industrial hemp fiber [45]. The research
program was aimed at utilizing materials that are renewable and avail-
able in perpetuity for the future use in the design and construction of
civil structures and commercial buildings. The program was conducted
at the Sustainable Material and Renewable Technology Laboratory at the
University of North Carolina. The NFCs were made from recycled ther-
moplastic polymers using extrusion and compression molding techniques.
The effects of fiber/matrix volume fraction on composite’s tensile, flexural
strength, and modulus were investigated. The surface morphology of the
natural fiber both before and after treatment was examined by using SEM.
The results of the study indicate that hemp fiber composites with rHDPE
have many desirable mechanical properties such as tensile strength, elastic
modulus, flexural strength, and modulus of rupture at 40% of fiber vol-
ume fraction. It was observed that the tensile strength of hemp fiber with
rHDPE composite reached 60.2 MPa at the 40% of fiber volume fraction,
which outperforms hemp composites with virgin thermoplastic matrix
reported in previous studies.
Kukle et al. employed short hemp fibers to reinforce LLDPE matrix
composites with fiber content in a range from 30 to 50 wt% [46]. Four dif-
ferent preprocessing technologies were investigated. The properties of the
composites such as tensile strength and elongation at break, tensile modu-
lus, microhardness, and water absorption were determined. Capillary vis-
cosimetry was used for fluidity evaluation. MFI was evaluated by standard
methods.
436 Adhesives for Wood and Lignocellulosic Materials

19.8.5 Miscellaneous
Leduc et al. prepared NFCs based on Agave fibers (Agave tequilana) and
LDPE by injection molding [47]. The agave fibers result from A. tequilana
Weber Azul after cooking, shredding, and sap extraction of the sugar-based
components. These fibers are currently discarded. Fiber revalorization
schemes had to be found in order to solve this accumulation problem.
Development of biofiber–polymer matrix composites was considered to
be ideal.
MAH grafted polyethylene (MAPE) was employed as coupling agent.
Studies were made to determine the effect of fiber content, coupling agent
addition, and presence of weld line on mechanical properties. It was found
that fiber concentration increased the tensile and flexural moduli but
decreased the impact strength. In all cases, MAPE addition enhanced the
tensile properties of the composites up to an optimum content between
2% and 5% on a fiber weight basis. The presence of a weld line substantially
reduced tensile properties, and this effect could be interpreted in terms of
molecular entanglement and orientation changes.
Raj et al. reported that tensile strength and tensile modulus were
improved substantially when maleated ethylene was used in the prepara-
tion of HDPE/cellulose-fiber-reinforced composites (at 10% and 30% fiber
concentration) [48]. The enhancement in the properties was attributed to
the coupling reaction (ester linkage) between the maleated ethylene and
the hydroxyl group of cellulose, which improves the bonding between the
fiber and matrix.
Carrasco et al. determined the mechanical properties of HDPE/wood
fiber composites made with two types of coupling agents—Epolene C-18
and Silane 174 [21]. Use of silane coupling agents resulted in composites
with better properties than the composites made from Epolene-treated
wood fibers. The influence of coupling agents on the mechanical proper-
ties of HDPE/wood fibers has also been reported by Saheb et al. [49]. It has
been observed that incorporation of wood fibers in HDPE resulted in an
increase in the stiffness and decrease in tensile strength for untreated wood
fibers. Treatment of wood fibers with silane coupling agent and polyisocy-
anate resulted in an increase in tensile strength.
Raj et al. and Saheb et al. have compared the tensile and impact proper-
ties of LLDPE/wood fiber composites with mica and glass fiber composites
and have shown the potential advantage of using wood fibers as reinforce-
ment in terms of material cost and specific properties [48, 49]. Pretreated
wood fiber produced a significant improvement in tensile strength and
modulus. Grafting of aspen chemithermomechanical pulp was found to
Polymer Matrix: Polyethylene 437

improve the mechanical properties of LLDPE composites as noted by


Beshay et al. and Saheb [18, 49].
The influence of various chemical treatments on the properties of sisal/
PE composites has been investigated by Joseph et al. [36]. The chemical
treatments included treatment with sodium hydroxide, isocyanate, and
peroxide. The enhancement in the properties was ascribed to the bonding
between sisal fiber and the PE matrix. Treatment with the cardanol deriv-
ative of toluene isocyanate was found to be better than other treatments
as evidenced by the decrease in the hydrophilic nature of the composite.
The composites exhibited better dimensional stability and retention of
properties even after aging, which was ascribed to the improved moisture
resistance.

References
1. Khanam, P.N. and AlMaadeed, M.A.A., Processing and characterization
of polyethylene-based composites. Adv. Manuf. Polym. Compos. Sci., 1, 2,
63–79, 2015.
2. Whiteley, K.S., Polyethylene, Vol. 29, in: Ullmann’s Encyclopedia of Industrial
Chemistry, Wiley-VCH Verlag, Weinheim, Germany, 2012.
3. Tham, C.M., Production of Polyethylene Using Gas Fluidized Bed Reactor,
pp.  1–20, National University of Singapore, Singapore, 2009, http://www.
klmtechgroup.com/PDF/Articles/articles/Fluidized-Bed-Reactor.pdf.
4. Azmi, A. and Aziz, N., Low density polyethylene tubular reactor modelling:
Overview of the model developments and future directions. Int. J. Appl. Eng.
Res., 11, 9906–9913, 2016.
5. Fregonese, D., Mortara, S., Bresadola, S., Ziegler–Natta MgCl2-supported
catalysts: Relationship between titanium oxidation states distribution and
activity in olefin polymerization. J. Mol. Catal. A, Chem., 172, 89–95, 2001.
6. Kashiwa, N.J., The discovery and progress of MgCl2-supported TiCl4 cata-
lysts. J. Polym. Sci. Part A: Polym. Chem., 42, 1–8, 2004.
7. Shamiri, A., Chakrabarti, M.H., Jahan, S., Hussain, M.A., Kaminsky, W.,
Aravind, P.V., Yehye, W.A., The Influence of Ziegler–Natta and metallocene
catalysts on polyolefin structure, properties, and processing ability. Materials,
7, 5069–5108, 2014.
8. Mandal, B.M., Fundamentals of Polymerization, World Scientific, Hackensack,
NJ, 20132013.
9. Natta, G., Kinetic studies of α-olefin polymerization. J. Polym. Sci., 34, 21–48,
1959.
10. G. Natta, P. Pino, P. Mazzanti, Regular linear head-to-tail polymerizates of
certain unsaturated hydrocarbons and filaments comprising said polymer-
izates, US Patent 3,715,344, 1973, US Patent Office, Alexandria, Virginia.
438 Adhesives for Wood and Lignocellulosic Materials

11. Ziegler, K., Organometallic Chemistry, ACS Monograph 147, H. Zeiss (Ed.),
pp. 194–269, Reinhold Publishing, New York, 1960.
12. Kaminsky, W., Funck, A., Hähnsen, H., New application for metallocene
catalysts in olefin polymerization. J. Chem. Soc., Dalton Trans., 41, 8803–
8810, 2009.
13. Kesti, M.R., Coates, G.W., Waymouth, R.W., Homogeneous Ziegler–Natta
polymerization of functionalized monomers catalyzed by cationic Group IV
metallocenes. J. Am. Chem. Soc., 114, 9679–9680, 1992.
14. Klimke, K., Parkinson, M., Piel, C., Kaminsky, W., Spiess, H.W., Wilhelm, M.,
Optimisation and application of polyolefin branch quantification by melt-
state 13C-NMR spectroscopy. Macromol. Chem. Phys., 207, 382–395, 2006.
15. Collins, R.A., Russell, A.F., Mountford, P., Group 4 metal complexes for
homogeneous olefin polymerisation: A short tutorial review. Appl. Petrochem.
Res., 5, 3, 153–171, 2015.
16. Kaminsky, W., Highly active metallocene catalysts for olefin polymerization.
J. Chem. Soc., Dalton Trans., 0, 1413–1418, 1998.
17. Kuehl, R. and ten Berg, G., Handbook of Petrochemicals Production Processes,
R.A. Meyers (Ed.), McGraw-Hill, New York, 2005.
18. Beshay, A.D., Kokta, B.V., Maldas, D., Daneault, C., Use of wood fibres
in thermoplastic composites II: Polyethylene. Polym. Comp., 6, 261–271,
1985.
19. Maiti, S.N. and Singh, K., Influence of wood flour on the mechanical proper-
ties of polyethylene. J. Appl. Polym. Sci., 32, 4285–4289, 1986.
20. Raj, R.G., Kokta, B.V., Maldas, D., Daneault, C., Use of wood fibers in ther-
moplastics. VII. The effect of coupling agents in polyethylene–wood fiber
composites. J. Appl. Polym. Sci., 37, 1089–1103, 1989.
21. Carrasco, F., Saurina, J., Arnau, J.J., Natural biofibre composites, in:
Proceedings 6th European Conference on Composite Materials, p. 483, Elsevier
Applied Science, France, 1993.
22. Berenbrok, P.A. and Liles, B.E., Dispersion of fibres in the composites,
in: Proceedings Special Areas Annual Technical Conference—ANTEC,
Toronto, vol. 3, pp. 2931–2933, Society of Plastics Engineers, Brookfield,
CT, 1997.
23. ASTM D1238, Standard Test Method for Melt Flow Rates of Thermoplastics by
Extrusion Plastometer. American Society of Testing Materials International,
West Conshohocken, Pennsylvania, 1990.
24. ISO 1133 Melt Mass-Flow Rate and Melt Volume-Flow Rate Thermoplastics,
International Standards Organisation, 2005.
25. En, H.-L. and Porter, R.S., Composite of polyethylene and kenaf, a natural
cellulose fibre. J. Appl. Polym. Sci., 54, 1781–1783, 1994.
26. Ku, S.G. and Kim, Y.C., Effect of grafting degree of maleic anhydride on
the physical properties of expandable HDPE/KF. Proceedings Annual Int.
Conference on Chemical Processes, Ecology & Environmental Engineering
(ICCPEE’16), Pattaya, Thailand, April 28–29, 2016.
Polymer Matrix: Polyethylene 439

27. Behjat, T., Russly, A.R., Luqman, C.A., Yus, Y., Nor Azowa, I., Effect of PEG
on the biodegradability studies of Kenaf cellulose–polyethylene composites.
Int. Food Res. J., 16, 243–247, 2009.
28. Avella, M., De Vlieger, J.J., Errico, M.E., Fischer, S., Vacca, P., Volpe, M.G.,
Biodegradable starch/clay nanocomposite films for food packaging applica-
tions. Food Chem., 93, 467–474, 2005.
29. Ismail, H., Hamid Abdullah, A., Abu Bakar, A., Kenaf core reinforced
high-density polyethylene/soya powder composites: The effects of filler load-
ing and compatibilizer. J. Reinf. Plast. Comp., 29, 2489–2497, 2010.
30. Viet, C.X., Ismail, H., Rashid, A.A., Takeichi, T., Kenaf powder filled recycled
high density polyethylene/natural rubber biocomposites: The effect of filler
content. Int. J. Integr. Eng., 4, 1, 22–25, 2010.
31. Akubue, P.C., Igbokwe, P.K., Nwabanne, J.T., Production of Kenaf fibre rein-
forced polyethylene composite for ballistic protection. Int. J. Sci. Eng. Res., 6,
8, 1–7, 2015.
32. Nilsson, L., Reinforcement of concrete with sisal and other vegetable fibres,
Stockholm: Swedish Building Research Summaries. D-14, 68, 1975.
33. Mattoso, L.H.C., Ferreira, F.C., Curvelo, A.A.A., Sisal fibre: Morphology and
applications in polymer composites, in: Lignocellulosic–Plastics Composites,
A.L. Leâo, F.X. Carvalho, E. Frollini (Eds.), pp. 21–51, USP/UNESP, São
Paulo, 1997.
34. Weindling, L., Long Vegetable Fibres, Columbia University Press, New York,
1947.
35. Li, Y., Mai, Y.-W., Ye, L., Sisal fibre and its composites: A review of recent
developments. Comp. Sci. Technol., 60, 2037–2055, 2000.
36. Joseph, K., Tolêdo Filho, R.D., James, B., Thomas, S., Hecker de Carvalho, L.,
A review on sisal fibre reinforced polymer composites. Rev. Bras. Eng. Agr.
Amb., 3, 3, 367–379, 1999.
37. Joseph, K., Thomas, S., Pavithran, C., Effect of ageing on the physical and
mechanical properties of sisal fibre reinforced polyethylene composites.
Comp. Sci. Technol., 53, 99–110, 1995.
38. Joseph, J., Thomas, S., Pavithran, C., Effect of chemical treatment on the
tensile properties of short sisal fibre-reinforced polyethylene composites.
Polymer, 37, 5139–5149, 1996.
39. Fávaro, S.L., Ganzerli, T.A., de Carvalho Neto, A.G.V., da Silva, O.R.R.F.,
Radovanovic, E., Chemical, morphological and mechanical analysis of sisal
fiber-reinforced recycled high-density polyethylene composites. Express
Polym. Lett., 4, 8, 465–473, 2010.
40. Chianelli, R., Jr., Reis, J.M.L., Cardoso, J.L., Castro, P.F., Mechanical charac-
terization of sisal fibre-reinforced recycled HDPE composites. Mater. Res.,
16, 1393–1397, 2013.
41. Martin, A.R., Manolache, S., Mattoso, L.H.C., Rowell, R.M., Denes, F.,
Plasma modification of sisal and high-density polyethylene composites:
Effect on mechanical properties. Proceedings, Third International Symposium
440 Adhesives for Wood and Lignocellulosic Materials

on Natural Polymers and Composites—ISNaPol/2000, Workshop on Progress


in Production and Processing of Cellulosic Fibres and Natural Polymers
May, 14–17, 2000.
42. Li, X., Panigrahi, S., Tabil, L.G., A study on flax fibre-reinforced polyethylene
biocomposites. Appl. Eng. Agric., 25, 525–531, 2009.
43. Li, X., Development of flax fibre-reinforced polyethylene biocomposites by
injection moulding, PhD Dissertation, University of Saskatchewan, Canada,
2008.
44. Lu, N., Swan, R.H., Jr., Ferguson, I., Composition, structure, and mechani-
cal properties of hemp fibre reinforced composite with recycled high-density
polyethylene matrix. J. Comp. Mater., 46, 16, 1–11, 2012.
45. Lu, N., Asce, M., Korman, T.M., Engineering sustainable construction mate-
rial: Hemp-fibre reinforced composite with recycled HDPE matrix. J. Archit.
Eng., 19, 3, 2013.
46. Kukle, S., Vidzickis, R., Zelca, Z., Belakova, D., Kajaks, J., Influence of
hemp fibres pre-processing on low density polyethylene matrix compos-
ites properties. IOP Conf. Ser.: Mater. Sci. Eng., 123, 012022, 2016, doi:
10.1088/1757-899X/123/1/012022.
47. Leduc, S., Ureña, J.R.G., González-Núñez, R., Quirarte, J.R., Riedl, B.,
Rodrigue, D., LDPE/agave fibre composites: Effect of coupling agent and
weld line on mechanical and morphological properties. Polym. Polym.
Comp., 16, 2, 115–123, 2008.
48. Raj, R.G., Kokta, B.V., Daneault, C., Wood flour as a low-cost reinforcing
filler for polyethylene: Studies on mechanical properties. J. Mater. Sci., 25,
1851–1855, 1990.
49. Saheb, D.N. and Jog, J.P., Natural fibre polymer composites: A review. Adv.
Polym. Tech., 18, 351–363, 1999.
20
Polymer Matrix: Polypropylene

20.1 Introduction
Polypropylene (PP) has recently become an attractive polymer for many
engineering applications. Relatively low price, excellent chemical resis-
tance, good processability, and the possibility of modifying its mechanical
properties by adding fillers and reinforcements have made it an important
commodity plastics for applications in automotive, land transport, home
appliances, and other industries. Due to its relatively low processing tem-
perature, PP is considered as one of the primary polymeric matrix materials
of choice for thermoplastic–biofiber composites (natural-fiber-reinforced
composites). Low processing temperature is essential because of low ther-
mal stability of natural fibers.

20.2 PP Manufacture
Propylene is the raw material for the production of PP. Even minute
amounts of impurities affect the polymerization reaction. The raw
material refining process is the first step in the polymerization of pro-
pylene. The impurities such as water, oxygen, carbon monoxide, carbon
dioxide, carbonyl sulfide, and the like are to be removed from the pro-
pylene and other monomers as well as the solvents and other auxiliary
materials. The polymerization of propylene and, if necessary, ethylene
and other monomers is carried out in the presence of a catalyst. Most of
the main industrial catalysts are in a granular shape. The polymeriza-
tion reaction occurs at the active centers of the catalyst particles. The
PP that is formed precipitates out, and the catalyst splits into primary
particles.
The process of producing crystalline PP was first commercialized in 1957
by Montecatini in Italy and Hercules in the United States. Since then, the
process has undergone a number of revolutionary changes. Accordingly,

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (441–466) © 2019
Scrivener Publishing LLC

441
442 Adhesives for Wood and Lignocellulosic Materials

the current production of PP is based on low-pressure processes using a


Ziegler–Natta catalyst that produces a product with an isotactic content
of 90% or more. After the production of PP, an “after treatment process”
is necessary. This is a process for eliminating catalyst residue, the solvent
and atactic polymers from the PP particles. The process for eliminating the
catalyst is known as “deashing”.
The PP production technology can be classified into three generations.
In the first generation process, there are steps for deashing and atactic poly-
propylene (AP) removal. The second-generation process is characterized
by the absence of deashing (non-deashing). The third-generation process
is non-deashing and non-AP removal [1].
In addition, classification can also be done according to the method of
carrying out the polymerization: solvent processes, bulk polymerization
processes, and vapor phase polymerization processes. These details are
given in Figure 20.1.

20.2.1 Catalysts
Details of Ziegler–Natta catalysts and the metallocene catalysts are dealt
with in Section 19.3.1 [2].

20.2.2 α- and β-Forms of PP


Isotactic PP is one of the most common polymeric materials for natural-
fiber-reinforced composites. Commercial-grade PP crystallizes into the stable
α-form with sporadic β-form crystalline structure formation. However,
when special crystallization procedures are applied, or specific nucleators
are added, the β-form can become a predominant crystalline form [3, 4].
Recently, β-PP has attracted the interest of numerous scholars because it
possesses some advantageous mechanical properties, such as high tough-
ness, drawability, and low thermal deformation temperature compared to
α-PP. Wu et al. have reported surface modification of flax fibers by MAPP
and vinyltrimethoxy silane (VTMO) to alter the interfacial bonding of PP
resins [5, 6].
However, no study on the mechanical properties of fiber-reinforced
β-PP composites currently exists in the literature. The study by Wu
et  al. involved investigations on the effects of treatments of MAPP
and VTMO on the mechanical properties of flax/β-PP [5]. The inter-
facial performances of flax/α-PP composites were also evaluated for
comparison.
Polymer Matrix: Polypropylene 443

1st Generation
Solvent polymerization process

Monomer Recovery

Polymerization Degassing Deashing Drying PP

Solvent Recovery

AP, Ash
(a)
2nd Generation
a) Solvent polymerization process (Non-deashing)

Monomer Recovery

Polymerization Degassing Drying PP

Solvent Recovery

AP
(b)
b) Bulk polymerization process (Non-solvent)

Monomer Recovery
AP, Ash
Polymerization Extraction PP

(c)
3rd Generation
Vaper phase polymerization process (Non-deashing, Non-AP)

Polymerization PP
(d)

Figure 20.1 PP manufacturing process: (a) 1st generation solvent polymerization


process, (b) 2nd generation solvent polymerization process (no dehashing), (c) Bulk
polymerization process (no solvent), (d) 3rd generation vapour phase polymerization
process (no deaashing and no ATP produced). (Review: Sumitomo Kagaku 2009).

20.2.3 Polymerization Methods


20.2.3.1 Solvent Polymerization Process
Figure 20.1a shows the first-generation solvent polymerization process
developed by Sumitomo Chemical [7]. The Sumitomo Chemical solvent
polymerization process was originally developed by the Italian company
444 Adhesives for Wood and Lignocellulosic Materials

Montecatini, the first in the world to industrialize PP. Sumitomo Chemical


made substantial technical improvements subsequently.
The solvent process for the polymerization of propylene is carried out
at a temperature of 50 to 80°C and a pressure of approximately 1 MPa in
hexane, heptane, or another inert hydrocarbon solvent. PP particles were
separated from unreacted propylene. Deashing (decomposition and elimi-
nation of the catalyst using alcohol) is carried out. PP is washed with water
and centrifugally separated and dried. A large amount of alcohol and water
are used for deashing. In addition. AP, which is produced as a secondary
product at 10% of the amount polymerized, is separated (first generation).
This process is complicated and costly.
In the second generation, the deashing process was omitted because of
improvements in catalyst activity, and the large amounts of alcohol and
water became unnecessary. Special catalysts have been developed that gave
a high level of stereo-regularity that enabled reduction in the proportion of
AP (see Section 20.2.1).

20.2.3.2 Bulk Polymerization Process


In the bulk polymerization process, liquefied propylene is directly used
without any solvents such as hexane and heptane. The energy costs for the
steam, electricity, etc. required for recovering the solvent may be greatly
reduced. The bulk polymerization represents the second-generation pro-
cess [7]. The typical process conditions for the bulk polymerization are a
temperature of 50 to 80°C and a pressure that is roughly the vapor pressure
of propylene. It changes according to the temperature, but is in a range of
2 to 4 MPa. The polymerization reaction is rapid, and the retention time is
shortened leading to increase in the volumetric efficiency. The reactor size
for obtaining the same production capacity can therefore be smaller than
the conventional reactors.

20.2.3.3 Gas Phase Polymerization Process


The first commercial gas phase processes for polyolefin production
appeared in the late 1960s [8]. Amoco developed the first stirred-bed gas
phase reactor for PP manufacture in the 1970s [9–12]. The gas phase pro-
cess involves the feeding of liquid monomer into the reactor, which evap-
orates upon entering. The system involves mechanical agitation, rather
than fluidization. Companies that license PP gas phase processes include
BP Amoco (Innovene), Dow (Unipol), Novolen Technology Holdings
(Novolen), and Basell (Spheripol, Catalloy).
Polymer Matrix: Polypropylene 445

Fresh monomer, catalyst, and modifiers (chain-transfer agent, etc.) are


fed to the reactor section. Hydrogen is added to control the molecular
weight. Traditionally, the resulting polymer powder stream then under-
goes catalyst deactivation (deashing). However, complete elimination of
deashing and AP removal operations because of the rapid improvement
in catalyst performance. Special catalysts have been developed that gave
a high level of stereo-regularity; i.e., tacticity control agents, fed with the
special catalyst, are commonly used to increase the isotactic content of the
PP [13], which enabled reduction in the proportion of AP (see Section
20.2.1). Further simplifications were achieved in the process, and it
achieved a position as the third-generation process capable of manufactur-
ing high-performance products with diverse levels of quality. Additional
modifiers are then added, depending on the intended use of the product.
Finally, the polymer product is pelletized.
The typical operating conditions are a temperature of 50 to 80°C and a
pressure in the range of 1 to 2 MPa. Various types of reactors, such as stirred
tanks and fluidized beds, have been developed by various companies.
In the fluidized bed process, gaseous propylene contacts a solid catalyst in
a fluidized bed reactor. The process can be separated into three different areas:
purification and reaction, resin degassing and pelletizing, and vent recovery.

20.3 Biofiber Composites Based on PP


20.3.1 Kenaf-Based Composites
Kenaf (Hibiscus cannabinus), originating from Africa, has traditionally
been a source of bast fiber in India and China, which together account for
more than 75% of worldwide kenaf production. Kenaf is a very important
natural fiber with robust mechanical properties.
Aji et al. [14] have published an exhaustive review highlighting the fiber
surface modification to enhance adhesion between the reinforcing fibers
and the polymeric matrix materials, mechanical and tribological proper-
ties, green composites, etc. Another review has been published by Kamal
et al. [15].
Oil palm and kenaf are regarded as an industrial and commodity crop
in Malaysia. They are also grown in other parts of the world. They are
important plants cultivated in Malaysia for the oil (for palm oil) and fiber
and hence promoted in government policies. Kenaf was first introduced
in Malaysia in the early 1970s and was highlighted in the late 1999 by the
National Economic Action Council [16].
446 Adhesives for Wood and Lignocellulosic Materials

In 2000, Malaysian Agricultural Research and Development Institute


(MARDI) coordinated a fast track research and development projects.
MARDI has successfully developed variety screening, agronomic practices
for Kenaf cultivation, harvesting and mechanization, fiber processing, and
some downstream applications such as animal feed and biocomposites
[16]. Presently, research on kenaf is highly focused on the production of
biofuel, biocomposite materials, and bioproducts [16].
Many research projects have been carried out to determine the mechan-
ical properties of kenaf-fiber-reinforced PP [17–22] and some described
PP composites with high kenaf fiber content [23].
The general characteristics of kenaf fibers and the mechanical properties
of its polymer composites as well as applications of such materials have
been reviewed by Rashdi et al. [24].
The effect of maleic anhydride (MAH) as a coupling agent has been
intensively investigated [19–21], and a significant increase in flexural and
tensile strength was reported for maleated polypropylene (MAPP) (2 wt%
of MAH) [25]. However, no significant difference could be observed
between kenaf-fiber-reinforced PP and kenaf-fiber-reinforced MAPP with
respect to the tensile and flexural moduli.
Anuar et al. employed a nonwoven processing technique to prepare nat-
ural fiber PP composites [26]. This method has not received attention so
far. The advantage of this method is that natural fiber loading in a nonwo-
ven composite can be as high as 50–70% in contrast to only up to 30% in
the case of injection molding and resin transfer molding techniques cur-
rently adopted widely.
With such a high fiber loading, kenaf fibers function as the domi-
nant loading transfer component in kenaf–PP natural fiber composite
(KPNC). However, too much fiber loading affects the wetting and hence
poor interfacial bonding and lower strength [27, 28]. The manufacture of
KPNC involved three steps, namely, carding, needle punching, and heat
compression.
Sixty percent fiber content was found to be the optimum for flexural,
tensile, and compression strength of KPNC. It has to be mentioned that
tensile strength of kenaf fibers was significantly higher than that of empty
fruit bunch (EFB) fibers.
With particular focus on packaging applications, David et al. deter-
mined the elastic and dynamic characteristics of natural-fiber-based com-
posites, more particularly the flexural behavior of kenaf-reinforced PP
(kenaf/PP) composites [29]. The mass fraction of kenaf was varied from
0 to 60 wt% in 10% increments. The results obtained indicated that the
elastic modulus of the kenaf/PP composites increased while the ultimate
Polymer Matrix: Polypropylene 447

tensile stress of the composites decreased with increasing fiber loadings.


The fracture toughness of the kenaf/PP composites decreases by 56% as the
fiber fraction is increased from 10 to 60 wt%.
In response to the current increasing demand for flame-retardant com-
posite materials, Ismail et al. studied the flame retardancy and mechanical
properties of kenaf-filled PP containing ammonium polyphosphate (APP)
[30]. PP burns rapidly with a relatively smoke-free flame and without leav-
ing a char residue due to its wholly aliphatic hydrocarbon structure [31].
Further addition of organic fibers such as kenaf increases the flammability
of the PP composite [32]. Flame retardant should therefore be added into
natural-fiber-filled PP composite in order to reduce the flammability.
Over the last decade, the most commonly used additive types are inor-
ganic compounds, halogenated compounds, and phosphorus compound.
Halogenated compounds as flame retardant are said to function primarily
by a vapor phase flame inhibiting mechanism through radical reaction.
As a result, combustion products can cause corrosiveness and also evolve
toxic products [33]. Therefore, the demand to halogenated flame retardant
has decreased due to its harmfulness to the environment.
APP is one of the phosphorus-based flame retardants commonly used
in PP. It is an intumescent flame retardant that forms during burning a
foamy multicellular char on the polymer surface, which insulates the sur-
face from atmospheric oxygen and inhibits the conduction of heat to the
surface, thereby hindering the polymer from burning. Determination of
limiting oxygen index (LOI) and thermogravimetry were employed to
characterize the flame retardancy.
Dimeski et al. carried out research focused on development of a light-
weight, environmentally friendly, cost-effective composite material based
on natural, i.e., kenaf, fibers applied as partitioning panels in buildings [34].
The research has shown that high-fiber/low-matrix kenaf/PP composites
can fulfill the end user requirements for partitioning boards in offices and
dwelling houses to a large extent. The results have shown that the com-
posites exhibit an average flexural strength of 30.1 MPa and a modulus of
9.0 GPa, an impact toughness of 43.1 kJ/m2, and a compression strength
of 17.4 MPa.
Generally, it is reported that the tensile properties decreased with
increasing immersion time in water [35–38]. Further, the mechanical
degradation increased as the number of cyclic immersions increased.
Accordingly, Haniffah et al. carried out the continuous and cyclic tests on
kenaf/PP composites [39]. The aim of this study was to compare the effect
of domestic bleach (sodium hypochlorite 5.25% solution) and water under
cyclic immersion on tensile properties of kenaf/PP composites for potential
448 Adhesives for Wood and Lignocellulosic Materials

bathroom application. All samples underwent four cycles of immersion


and conditioning with each cycle consisting of 3 days of immersion and
4  days of conditioning. After the fourth day of conditioning, the speci-
mens were tested using a universal testing machine (Instron). The other set
of specimens underwent immersion for 12 days and was conditioned for
4 days in order to evaluate the tensile properties of the composites under
cyclic and continuous immersion. Both continuous and cyclic immersions
in bleach affected the tensile strength of the kenaf-fiber-reinforced PP
composites. Increased number of cycles of immersion and conditioning
caused degradation of tensile strength and modulus. However, continuous
immersion in the bleach reduced the tensile strength of composites more
compared to cyclic immersion.
Sanadi et al. made the kenaf/PP biocomposites by blending in a thermo-
kinetic mixture kenaf fibers and PP [19, 23]. Biofiber weight fractions were
varied up to a maximum of 60%. A MAPP was used to improve the inter-
action and adhesion between the nonpolar matrix to the polar lignocellu-
losic fiber. The specific tensile moduli of 50% by weight (39.5 by volume)
of kenaf/PP composite compared favorably with a 40% by weight of glass-
fiber=reinforced PP injection-molded composite. These results suggested
that kenaf fibers are viable alternative to inorganic/mineral-based reinforc-
ing fibers as long as right processing conditions are employed.
Wambere et al. employed a film stacking technique followed by the
hot-pressing method for preparing kenaf/PP biocomposite [40]. Modified
film of PP [containing 98.5% by weight of commercial PP of very high melt
flow (800 dg/min) and 1.5% MAH-grafted PP] was first prepared by an
undisclosed technique. This modified PP film improved the impregnation
of the kenaf fiber mat and enhanced the adhesion between the nonpolar
matrix and polar lignocellulosic fibers. The modified PP films were stacked
between three layers of random kenaf fiber mat. Composites were made
by pressing the mat in a hot press at 180°C for 2 min at 130 bar pressure
followed by cooling under pressure for 3 min.

20.3.2 Oil-Palm-Fiber-Based Composites


Biomass residues after extracting the palm oil are oil palm trunks (OPT),
fronds (OPF), kernel shell, EFB, and pressed fruit fibers. These residues are
valuable source of lignocellulosic natural fibers. Ridzuan et al. prepared
EFB fiber–PP biocomposites incorporating magnesium hydroxide as flame
retardant [41]. Biocomposites were made from both alkali-treated and
untreated EFB–PP. Similarly, composites were made using PP alone as well
as with MAH-grafted PP. Significant improvement in flexural modulus
Polymer Matrix: Polypropylene 449

(133%) and flame-retardant property (60%) of alkali-treated EFB fibers–


PP composite with MAPP and FR has been reported.
Kittikorn et al. employed four different methods to modify the surface
of the EFB of oil palm fibers, namely, propionylation, vinyltrimethoxysi-
lanization, PP-g-MA dissolution modification, and PP-g-MA blending
and integration into PP matrix [42, 43]. The composites were subjected
to a water absorption process at different temperatures. An increase in
fiber content and temperature uptake increased the water uptake of all
biocomposites. However, the biocomposites containing modified fibers
showed reduced water uptake, rate of diffusion, and sorption, proving
that the surface modification is effective in reducing the penetration into
composites.
Rozman et al. prepared PP hybrid composites using oil palm EFB and
glass fibers as reinforcing agents in PP matrix [44, 45]. Three types of
coupling agents, i.e., maleic-anhydride-modified PP (commercial name,
Epolene, E-43), polymethylenepolyphenyl isocyanate (PMPPIC), and
3-(trimethoxysilyl)-propylmethacrylate (TPM), were attempted. Overall,
E-43 and TPM had imparted considerable improvements in the flexural
and tensile properties. However, only slight improvements in some cases
were shown for those composites treated with PMPPIC.
OPF meal was reacted with γ-methacryloxypropyl-trimethxy silane.
The adduct formed was mixed with glycidyl methacrylate prior to hot
pressing. Well-consolidated boards were produced by the process. All the
silane-modified samples performed well in all the tests conducted. Modulus
of rupture, modulus of elasticity, and impact strength were reduced after
some levels of chemical loading (weight percent gain). All modified sam-
ples showed greater dimensional stability than the untreated samples in
water, toluene, and chloroform tests [46].
Hydrolysis of OPEFB ash creates a base solution that can be utilized
in an alkaline treatment process to increase the mechanical properties of
natural composites. Fatra et al. investigated the effect of various extracts
of OPEFB ash on the tensile strength, flexural strength, and water absorp-
tion of an OPEFB fiber–PP composite [47]. The experimental design
used was the Response Surface Method–Central Composite Design to
determine the optimum conditions. The statistical analysis of the date
showed that the highest flexural strength (30.216 MPa) was achieved at
5 wt% alkaline concentration, 12 h soaking time, and 3 cm fiber length.
The lowest water absorption (0.324%) was achieved at 10 wt% alkaline
concentration, 24 h soaking time, and 2 cm fiber length.
Radiation treatment was also found to enhance the properties of wood
fiber–thermoplastic composites [48–50]. This study evaluated the effects of
450 Adhesives for Wood and Lignocellulosic Materials

different irradiation techniques and some reactive additives on the proper-


ties of EFB–PP composite.
Khairul Zaman et al. evaluated the effects of different irradiation tech-
niques and reactive additives on the properties of EFB–PP composites
[51]. Different modes of irradiation were investigated. Mono-, di-, and
trifunctional monomers such as 2-ethylhexyl acrylate, 1,6-hexadiol diac-
rylate, tripropylene glycol diacrylate, and trimethylol propane triacrylate
(TMPTA) were employed as reactive additives. All samples were irradiated
at room temperature with an EB accelerator voltage of 3 MeV, a beam cur-
rent of 1 mA, and a dose of 1 kGy. Conditions of irradiation are as follows:
(1) no irradiation, (2) both EFB and PP were irradiated, (3) PP alone was
irradiated, and (4) EFB alone was irradiated.
These studies showed that radiation can enhance the mechanical prop-
erties of EFB fiber–PP composites compared to unirradiated PP–biofiber
composite. By irradiating PP alone, the composites can be easily processed
using the commonly used polymer mixers. The addition of reactive addi-
tives such as acrylate monomers acts as processing or lubricating agents.
The addition of reactive additives also further enhanced the properties of
the irradiated EFB–PP composites, with TMPTA giving the best results.
Abdul Khalil et al. carried out studies on PP-based biocomposite made
from 25-year-old oil palm biomass (OPB) fiber at different fiber loadings
(10%, 20%, 30%, 40%, and 50%) [52]. The types of OPB used are oil palm
EFBs, OPF, and OPT. Transmission electron microscopy confirmed that
the cell wall structures of the various oil palm fibers have different cell wall
thicknesses and exhibit the same ultrastructure as that of wood. The fibers
consist of middle lamella, primary, and thick secondary walls with differ-
ent thicknesses for different types of fibers. The secondary wall is differ-
entiated into an S1 layer, a unique multilamellae S2 layer, and an S3 layer.
OPB fibers were compounded with PP using a Brabender DSK 42/7 twin-
screw extruder. The mechanical properties such as tensile, flexural, and
impact strengths of the OPB–PP composite are determined. The melt flow
index (MFI) of the composite materials was also studied. Generally, the
results showed that lower fiber loading (10%) exhibited the highest tensile
strength and MFI properties as compared to higher fiber loading (50%).
Evidence of a fiber–matrix interphase was analyzed using scanning elec-
tron microscopy (SEM).
The physical and mechanical properties of oil palm empty fruit bunch
(OPEFB)-reinforced PP matrix composites with different fiber loadings
(10, 15, 20, 25, and 30 wt%) have been studied [53]. Two different types of
chemical treatments, namely, sodium periodate (NaIO4) oxidation and urea
(CO(NH2)2) coupling reaction, have been used to improve the properties
Polymer Matrix: Polypropylene 451

of the composites. In the current research, OPEFB-fiber-reinforced PP


composites were manufactured using a high-voltage hot compression
technique. Scanning electron micrographs of the fractured surfaces were
taken to study the fiber/matrix interface adhesion. Reduced fiber agglom-
eration and improved interfacial adhesion were observed under a scanning
electron microscope in the case of oxidized OPEFB fiber and improved the
compatibility in the case of urea-treated oxidized OPEFB-fiber-reinforced
PP composites. Tensile and flexural modulus significantly increased with
fiber loading.
However, tensile and flexural strength decreased beyond 25% by weight
fiber loading. It was found that oxidized OPEFB-fiber-reinforced PP
composites absorbed less water compared to raw and urea-treated oxi-
dized OPEFB-fiber-reinforced PP composites. Overall, chemically treated
OPEFB-fiber-reinforced PP composites showed better fiber/matrix inter-
actions as observed from the good dispersion of fibers in the matrix system.
Unlike the untreated OPEFB-fiber-reinforced PP composites, all treated
OPEFB-fiber-reinforced composites had the same tendency of increasing
the mechanical properties of composites.
Soil burial tests were carried out to evaluate the effect of biodegradation
on the mechanical properties (tensile, flexural, and impact) and the mass
loss of OPT fiber-filled RPP composites, as compared to control samples
(virgin PP and RPP without filler) [54]. The composite samples were pre-
pared using 30% w/w of OPT fiber reinforcement. Compounding was car-
ried out using a Haake Rheodrive 500 twin-screw compounder operating
at 190°C and 8 MPa for 30 min. The effect of biodegradation was performed
in a perspex plastic apparatus for 12 months. Assessments of the mechani-
cal properties and the percentage of mass loss were carried out at 3, 6, and
12 months of exposure in soil. The mechanical properties (tensile, flexural,
and impact) of materials deteriorated with an increase in exposure time.
Biodegradation of the composite increased with the duration of burial,
i.e., from 0 to 12 months. The tensile properties, flexural properties, and
impact strength of the composites decreased by about 38–47%, 37–50%,
and 47%, respectively, as compared to the value before the biological test.
Khalid et al. [55] isolated cellulose from oil palm EFB as per ASTM D
1104 [56] and employed the same for making PP–cellulose biocomposites.
Similarly, the EFB fibers were also used for preparing the biocomposite.
The cellulose fibers and the EFB fibers were blended in different ratios up
to 50 wt% with PP using a Brabender twin-screw compounder. Effects of
cellulose and EFB fibers on the mechanical properties of PP were investi-
gated. Studies on the morphological properties and the influence of fiber
loading on the properties of PP–cellulose and PP–EFBF composites were
452 Adhesives for Wood and Lignocellulosic Materials

also conducted. The PP–cellulose composite gave better results in compar-


ison with PP–EFBF composite. The changes in mechanical and morpho-
logical properties with different cellulose and fiber loading were discussed.
Wirjosentono et al. [57] achieved good compatibility between the oil
palm EFB fibers and PP by modifying the organic fiber surface with acrylic
acid. This has been confirmed from SEM, GPC, and IR. The surface modi-
fication was carried out using a solution technique and reactive processing
in a screw extruder. The first technique resulted in an increase of tensile
strength of the composites with increasing oil palm EFB fiber loading up
to 20% (by weight) and in the presence of 3% (by weight) of acrylic acid
and 0.01 mol ratio of dicumyl peroxide. However, the elongation at break
showed a decreasing trend. A similar observation was also made for com-
posites prepared in a single-screw extruder.
Jani et al. carried out studies on the oil palm EFB fiber–PP compos-
ites and employed untreated PP as well as PP mixed with MAPP [58].
Commercial MAPP (Epolene 43) and MAPP samples synthesized at the
laboratory were used for the purpose. The effectiveness of the MAPP as
a coupling agent for EFB fiber composites was determined. Several char-
acterizations on MAPP-treated composites were carried out including
Fourier transform infrared (FTIR) spectra, acid number, and Gd tests to
determine the action of the anhydride. The flexural and impact properties
of the treated composite samples produced in the laboratory using both
Epolene 43 and MAPP samples produced in the laboratory were signifi-
cantly better than those of the untreated composites.
Rozman et al. made oil palm EFB–PP composites using a twin‐screw
extruder as the compounding equipment [59]. Two levels of EFB were
employed: 40% and 60% of the total weight of the sample. Three types of
coupling agent, MAH‐modified PP (commercial name Epolene E‐43),
PMPPIC, and TPM, were used. Overall, all coupling agents imparted
considerable improvements in the flexural properties, with E‐43 showing
the highest enhancement. However, only E‐43 was observed to improve
impact strength and tensile properties of the composites. All composites
with coupling agents showed lower water absorption and thickness swell-
ing. The absorption and swelling decreased as the loading of the coupling
agents was increased.
Surface treatment of the OPF fiber was carried out using superheated
steam (SHS) in order to increase the compatibility between the OPF fiber
and PP [60]. Biocomposites were prepared using PP and SHS-treated
or untreated OPF fiber by the injection molding method to observe the
effects of fiber surface treatment on the mechanical properties of the bio-
composites. The microstructure of the untreated and SHS-treated OPF
Polymer Matrix: Polypropylene 453

fiber was characterized by using X-ray diffraction (XRD) and SEM. XRD
results showed that the crystallinity of OPF fiber was increased after SHS
treatment, indicating tougher fiber properties. The thermal degradation
behavior of untreated and SHS-treated OPF fiber was studied using ther-
mogravimetric (TG) and derivative thermogravimetric (DTG) analyses.
TG/DTG analyses revealed the improved thermal stability of OPF fiber
after the SHS treatment. The superior mechanical properties were obtained
for the SHS-treated fiber-reinforced biocomposites as compared with
untreated fiber-reinforced biocomposites due to the enhanced interfacial
interaction between the OPF fiber and PP.
Abdan in his book emphasized the feasibility of processing the com-
posites prepared from oil palm EFB/PP using injection molding processes
[61]. The effects of fiber size, fiber content, melt flow rate, and concen-
tration of MAPP on the mechanical, physical, rheological, and thermal
properties of EFB/PP composites were reported. The effects of types and
concentrations of reactive additives on the irradiated EFB/PP composites
were also investigated particularly with respect to the rheological behavior
and dynamic mechanical thermal characteristics. The EFB/PP composites
were prepared from thermomechanically pulped EFB fiber and PP resin.
The internal mixer was used to mix, and the injection molding machine
was employed to form the specimens in accordance to the ASTM stan-
dards. Electron beam was used to irradiate the EFB/PP. The EFB fiber size
and fiber content significantly affected the mechanical and physical prop-
erties of EFB/PP composites. The rheological properties were significantly
influenced by radiation. Fiber orientation was studied using SEM showing
EFB skin core structure.
Ramli et al. carried out studies on the effects of fiber type, fiber load-
ing, and coupling agent on the performance of OPB fiber–PP composites
[62]. Fiber composition and fiber morphology were evaluated by SEM and
energy-dispersive X-ray spectroscopy. Fiber was compounded with PP by
means of a twin-screw compounder. MAPP was used as a coupling agent
during compounding. The incorporated fiber contents for OPB composites
were up to 40% (by weight). The compounded samples were prepared into
test specimens by an injection molder. The composites were characterized
by tensile testing, flexural testing, impact testing, MFI, SEM, thermograv-
imetric analysis (TGA), and differential scanning calorimetry (DSC). The
most significant effect on strength and modulus was found by the addition
of coupling agent. This was attributed to the thermodynamic segregation
of the MAPP toward the interface, resulting in the formation of covalent
bonding to the –OH groups of the fiber surface. Composites with MAPP
also provided better thermal stability.
454 Adhesives for Wood and Lignocellulosic Materials

Kittikorn et al. studied the effect of chemical surface modifications


on EFB oil palm fiber/PP composite properties [42, 43]. FTIR spectra of
propionylated fiber and PP-g-MA-modified fiber showed the presence
of a carbonyl group of esters while vinyltrimethoxysilane-treated fiber
showed a peak of silicate, confirming that the modifications were suc-
cessful. PP-g-MA-modified fiber PP composite gave the highest mod-
ulus of 0.71 GPa at 20% fiber content in contrast to the modulus value
of 0.56 GPa for unmodified fiber PP composite. By DSC analysis, PP-g-
MA-modified fiber and vinyltrimethoxysilane-treated fiber PP compos-
ite at the same fiber content of 5% showed the highest crystallinity of 46%
and 44%, respectively, whereas unmodified fiber PP composite showed a
lower crystallinity of 38%. The DMTA analysis showed that after 60°C,
the modified fiber PP composites exhibited a higher stiffness than pure
PP-based composites.

20.3.3 Flax-Fiber-Based Composites


Isotactic PP is one of the most common polymeric materials for natural-fiber-
reinforced composites. Commercial-grade PP crystallizes into the stable
α-form with sporadic β-form crystalline structure formation. However,
when special crystallization procedures are applied, or specific nucleators
are added, the β-form can become a predominant crystalline form [2, 4].
Recently, β-PP has attracted the interest of numerous scholars because it
possesses some advantageous mechanical properties, such as high tough-
ness, drawability, and low thermal deformation temperature compared
to α-PP. Wu et al. [5] have reported surface modification of flax fibers by
MAPP and VTMO to alter the interfacial bonding of PP resins [6].
However, no study on the mechanical properties of fiber-reinforced
β-PP composites currently exists in the literature. The study by Wu et al.
(2016) involved investigations on the effects of treatments of MAPP and
VTMO on the mechanical properties of flax/β-PP. The interfacial perfor-
mances of flax/α-PP composites are also evaluated for comparison.
Zhu et al. have reviewed recent developments on flax-fiber-reinforced
polymeric composites [63]. The properties of flax fibers, as well as
advanced fiber treatments such as mercerization, silane treatment, acy-
lation, peroxide treatment, and coatings for the enhancement of flax/
matrix incompatibility are presented. The characteristic properties and
characterizations of flax composites on various polymers including PP
and polylactic acid, epoxy, bio-epoxy, and bio-phenolic resin are dis-
cussed. A brief overview is also given on the recent nanotechnology
applied in flax composites.
Polymer Matrix: Polypropylene 455

20.3.4 Sisal-Based PP Composites


The dynamic mechanical properties of short sisal-fiber-reinforced PP
composites containing both untreated and treated fibers have been stud-
ied by Joseph et al. with reference to fiber loading, fiber length, chemical
treatments, frequency, and temperature [64, 65]. By the incorporation of
short sisal fiber into PP, the storage moduli (E ) and loss moduli (E ) have
been found to be increasing whereas the mechanical loss factor (tan δ) has
been found to be decreasing. The storage modulus decreases with increase
in temperature. The treated fiber composites show better properties com-
pared to the untreated system. The Arrhenius relationship has been used
to calculate the activation energy for the glass transition. The use and lim-
itations of various theoretical equations to predict the tan  δ  and storage
modulus of the fiber-reinforced plastic composites have been discussed.
Cole–Cole analysis has been carried out to understand the phase behavior
of the composite samples. A master curve for the modulus of the blend is
drawn by applying the time–temperature superposition principle.
Common methods for manufacturing natural-fiber-reinforced ther-
moplastic composites, injection molding and extrusion, tend to degrade
the fibers during processing. Jayaraman reported on the development of a
simple manufacturing technique for sisal-fiber-reinforced PP composites
that minimizes fiber degradation and can be used in developing coun-
tries [66]. Composite sheets with a fiber length greater than 10 mm and
a fiber mass fraction in the range 15% to 35% exhibited good mechanical
properties.
Joseph et al. carried out investigation on the fiber breakage analysis
during melt mixing and solution mixing processes employed for the prepa-
ration of sisal fiber–PP matrix composites [64, 65]. A large amount of fiber
breakage was observed during melt mixing. The fiber breakage analysis
during composite preparation by melt mixing was carried out using optical
microscopy. A polynomial equation was used to model the fiber‐length
distribution during melt mixing. The experimental mechanical proper-
ties of sisal/PP composites were compared with existing theoretical mod-
els such as the modified rule of mixtures, parallel and series models, the
Hirsch model, and the Bowyer–Bader model. The dependence of the ten-
sile strength on the angle of measurement with respect to fiber orientation
was also modeled.
Treated sisal fibers were used by Pimenta et al. as reinforcement for
PP  composites, with MAH-grafted PP (MAPP) as coupling agent [67].
The composites were made by melt processing of PP with the fiber in a
heated roller followed by multiple extrusions in a single-screw extruder.
456 Adhesives for Wood and Lignocellulosic Materials

Injection-molded specimens were produced for the characterization of


the material. In order to improve the adhesion between fiber and matrix
and to eliminate odorous substances, sisal fibers were treated with boiling
water and with NaOH solutions at 3 and 10 wt%. The mechanical proper-
ties of the composites were assessed by tensile, bending, and impact tests.
Additionally, the morphology of the composites and the adhesion at the
fiber–matrix interface were analyzed by SEM. The fiber treatment led to
very light and odorless materials. Fiber treatment caused an appreciable
change in fiber characteristics, yet the mechanical properties under tensile
and flexural tests were not influenced by that treatment. Only the impact
strength increased in the composites with alkali-treated sisal fibers.

20.3.5 Hemp-Based PP Composites


Hemp (Cannabis sativa L.) has been cultivated for at least 6000 years, and
it may be one of the oldest non-food crops. The most common purpose of
hemp cultivation is to isolate the fibers present in the bark on the hemp stem
surface, for production of ropes, textiles, and paper. Markiewicz et al. pre-
pared composites with isotactic PP as the matrix and lignocellulosic fibers
derived from hemp and flax plants as the reinforcement [68]. The electric
permittivity ε and the ac conductivity σac of these composites were mea-
sured in the frequency range from 100 Hz to 13 MHz at room temperature.
The linear relationship between the reciprocal relative permittivity (1/ε )
and the volume fraction of the lignocellulosic material was established. The
loading of the PP with the lignocellulosic materials increases the electric
permittivity ε and improves the electrical conductivity σac. These compos-
ites can replace dielectrics in microchips and printed circuit boards used
in the electronic industry. Some composites were designed with the aim of
application as antistatic materials to dissipate static charges. The materials
characterized by a low electric permittivity are applied as insulators to cut
off the energy flow in alternating and direct electric current circuits.
Etaati et al. determined the mechanical properties (tensile, flexural, and
impact) of ground hemp fiber–PP composites [69]. Ground alkali-treated
hemp fiber and noil hemp fibers with various initial fiber lengths were uti-
lized to reinforce PP matrix. SEM, FTIR analysis, and dynamic mechan-
ical analyzer were employed to delineate the microstructural and tensile
characteristics of the two types of fibers. The fibers were comminuted into
different lengths of 0.2, 0.5, 1, and 2 mm; composites containing 40 wt%
short hemp fiber and 5 wt% MAH-grafted polypropylene (MAPP) were
fabricated by means of a twin-screw extruder and an injection molding
machine. Finally, the influence of hemp fiber type and initial hemp fiber
Polymer Matrix: Polypropylene 457

length on the tensile property of the composites was investigated. The


results revealed that addition of either noil hemp fiber or normal treated
hemp fiber into the pure PP matrix increased the tensile strength almost
twice and stiffness of the composites more than three times.
Zhao et al. showed that in situ inorganic modification of hemp with a
mixture of NaOH and CaCl2 (F1) is a promising method to improve both
the flame-retarding and mechanical properties of the biocomposite [70].
Hemp-reinforced PP composite with excellent flame-retarding and mechan-
ical properties was prepared with the melt blending method. The thermal
stability, mechanical, and flame-retarding properties of the composites were
studied by TGA, mechanical tests, and LOI measurements, respectively. The
LOI of the composites with 50% of F1 and 15% poly(phosphoric acid amine)
reached 25.5%, which was much higher than that of PP (18.5%).
Green composites from PP and lignin-based natural material were
made using a melt extrusion process. The lignin-based material used was
the so-called “liquid wood”. Liquid Wood Arboform LV100 (LV) was sup-
plied by Tecnaro GmbH, Germany (density = 1.30 g/cm3). The PP/“Liquid
Wood” blends were extruded with “liquid wood” content varying from
20 wt% to 80 wt%. MAPP was used as a coupling agent. The flexural, impact,
and dynamic mechanical properties were determined [71]. The addition of
the liquid wood resulted in a great improvement in terms of both the flexural
modulus and strength but a reduction of the impact strength.
Kechaou et al. prepared composites with isotactic polypropylene (iPP)
matrix and hemp fibers via a patented FIBROLINE process based on the
principle of the dry impregnation of a fiber assembly with a thermoplastic
powder (iPP), using an alternating electric field [72]. The authors carried
out an investigation to determine the influence of fiber/matrix interfaces
on dielectric properties coupled with mechanical behaviors. Fibers or
more probably the fiber/matrix interfaces allow the diffusion of electric
charges and delocalize the polarization energy. In this way, damages are
limited during mechanical loading and the mechanical properties of the
composites increase. The structure of composite samples was investigated
by X-ray and FTIR analysis. The mechanical properties were determined
by quasi static and dynamic tests. The dielectric investigations were car-
ried out using the SEMME (scanning electron microscope mirror effect)
method coupled with the measurement of the induced current.
Badji et al. studied the emission of volatile organic compounds (VOCs)
from hemp-fiber-reinforced PP biocomposites and compared them with
neat PP [73]. Such a study has been carried out for the first time since
natural-fiber-reinforced composite materials can be used in closed envi-
ronments such as car cabins. It has hence become a necessity to study the
458 Adhesives for Wood and Lignocellulosic Materials

emissions of VOCs to check their impact on indoor air quality. The influ-
ence of under windshield glass weathering on VOC release was investi-
gated. The exposure lasted for 1 year and the influence of fiber loading
on emissions was studied throughout the weathering period. The VOC
concentration at the material/air interface was determined using a pas-
sive sampling method involving an emission cell coupled with solid phase
microextraction. VOC analysis was then carried out by gas chromatogra-
phy coupled to mass spectrometry and flame ionization detections. One
of the most significant results is the drastic increase of concentration of
oxygenated compounds during the exposure, especially for biocomposites.
Among these oxidation by-products, formaldehyde, acetaldehyde, fur-
fural, and 2-furanmethanol, recognized as cancerogenic, mutagenic and
toxic for reproduction, were detected. A broad range of alkanes, specific
for PP matrix degradation, was also identified. Finally, measured concen-
trations of substances found in this work and listed in vehicle indoor air
quality standards were gathered in order to discuss impact on indoor air
quality due to the biocomposite emissions.
Zhang used hemp fibers as reinforcement in the composites [74]. These
composites were made by compression molding method under differ-
ent parameters, temperature, pressure, time, etc. The influences of these
parameters on the quality of the manufacturing process are analyzed. The
results showed that the temperature had a significant effect on the mechan-
ical properties of the agro-composites. The composite manufactured under
190°C had better mechanical properties than other composites. Moreover,
the accelerated aging test was carried out under two conditions: relative
humidity and ultraviolet. Incremental hole-drilling method was adopted
to analyze the residual stress generated during the manufacturing process
as well as during the aging process.
Badji et al. studied the degradation behavior of hemp-fiber-reinforced
PP  biocomposites  under outdoor and artificial weathering condi-
tions [75]. The artificial weathering effectively accelerated the degradation
mechanisms. Oxidation pathways and surface aspect alteration of
both polymer and biocomposites occurred faster. Whereas biocompos-
ites were mainly subjected to outdoor conditions due to high sensitiv-
ity of hemp fibers, neat PP was mostly affected by laboratory chamber
conditions. Its oxidation rate largely outstripped reinforced material
ones. Principal component analysis was used for verifying the differ-
ences between artificial and natural aging dataset in order to compare
the degradation mechanisms. Through statistical analysis, some attempts
were made to find equivalence between artificial and outdoor weathering
times.
Polymer Matrix: Polypropylene 459

Panthapulakkal and Sain studied the performance of injection-molded


short hemp fiber and hemp/glass fiber hybrid PP composites [76]. Results
showed that hybridization with glass fiber enhanced the performance
properties. Notched Izod impact strength of the hybrid composites
exhibited great enhancement (34%). Analysis of fiber length distribution
in the composite and fracture surface was performed to study the fiber
breakage and fracture mechanism. Thermal properties and resistance to
water absorption properties of the hemp fiber composites were improved
by hybridization with glass fibers. Overall studies indicated that the
short hemp/glass fiber hybrid PP composites are promising candidates
for structural applications where high stiffness and thermal resistance is
required.
Sullins et al. studied the effect of treatment of hemp fiber on the
mechanical behavior of fiber–reinforced PP composites [77], chemically
treating the hemp fiber with different concentrations of NaOH and/or
adding MAPP to the PP matrix. The purpose of the material treatment is
to enhance the bonding between the hemp fiber and the PP matrix, which
otherwise has low surface energy and limited bonding capability. The effect
of treatment on the mechanical properties was investigated under differ-
ent combinations of treatment(s) such as 5 wt% MAPP, 5% NaOH-treated
hemp fiber, 10% NaOH-treated hemp fiber, and 5% NaOH + 5 wt% MAPP.
Composites contained 15 wt% and 30 wt% of treated hemp fiber.
It was found that the composites made with treated fibers were found
to have better mechanical properties compared to the composites without
any treatment. The composites with 5  wt% MAPP addition showed the
best mechanical properties.
Nonwoven mats from hemp fibers and PP in various proportions were
produced and hot pressed to make composite material [78]. The effect of
hemp fiber content and anisotropy in the nonwoven mats resulting from
the carding technology were examined on the basis of the three-point
bending, tensile, and impact properties of the resultant composite mate-
rials. Because of the hydrophilic nature and poor dimensional stability of
cellulosic fibers due to swelling, the effect of water sorption on mechanical
performances was also investigated. Optimal mechanical properties were
achieved in composites made from 40% to 50% of hemp fiber by weight.
As expected, better mechanical properties are found in the specimens cut
from the composite sheets parallel to the direction of carding. A decrease in
three-point bending properties was noticed after immersing the composite
samples in distilled water for 19 days, while the impact strength increased.
Double carding of raw materials resulted in a decreased anisotropy in the
composite material.
460 Adhesives for Wood and Lignocellulosic Materials

Puech et al. analyzed the impact behavior of short hemp-fiber-


reinforced biocomposites through mechanical measurements, high-speed
imaging, and finite element modeling [79]. A drop-weight impact machine
was instrumented with a high-speed camera to measure the propagation of
macro-cracks and correlate it to the force–displacement dynamic response
at several impact energy levels. PP–hemp composites were found to exhibit
higher absorbed energies (up to 40%) than PP–glass composites due to
higher strain at break. The video tracking analysis highlighted that for a
given cumulated crack length, PP–hemp composite absorbed much more
energy, related to differences in failure mechanisms. The developed finite
element model was in good agreement with the experimental measure-
ments and the fracture growth pattern, thus constituting a useful tool to
predict the impact response of biocomposite parts.

References
1. Sumitomo Chemical Co, Review on Development of Polypropylene
Manufacturing Process, 2018. file:///C:/Users/hp/Downloads/polypropylen-
emanufacturing%20(12).pdf.
2. Bleeke, J.R. and Frey, R.F., Designer Plastics: How Can Chemists Synthesize
Polymers of Their Own Design, 2006. http://www.chemistry.wustl.edu/
~edudev/Designer/session7.html.
3. Lotz, B., Wittmann, J.C., Lovinger, A.J., A structure and morphology of
poly(propylenes): A molecular analysis. Polymer, 37, 4979–4992, 1996.
4. Cermák, R., Obadal, M., Ponížil, P., Polášková, M., Stoklasa, K., Hečková, J.,
Injection-moulded α- and β-polypropylenes: II. Tensile properties vs. pro-
cessing parameters. Eur. Polym. J., 42, 2185–2191, 2006.
5. Wu, C.-M., Lai, W.-Y., Wang, C.-Y., Effects of surface modification on the
mechanical properties of flax/β-polypropylene composites. Materials, 9,
6–11, 2016.
6. Yu, T., Wu, C.-M., Wang, C.-J., Rwei, S.P., Effects of surface modifications
on the interfacial bonding of flax/β-polypropylene composites. Compos.
Interface, 20, 483–496, 2013.
7. Shiga, A., Matsuyama, K., Kakugo, K., Hashimoto, H., Sumitomo Kagaku,
1980-II, p. 52, Sumitomo Kagaku Kabushiki-gaisha, Chūō-ku, Tokyo, Japan,
1980.
8. Choi, K.Y., Gas-phase olefin polymerization, in: Polymeric Materials
Encyclopaedia, J.C. Salamone (Ed.), CRC Press, Boca Raton, 2707.
9. J.W. Shepard, J.L. Jezl, E.F. Peters, R.D. Hall, Divided Horizontal Reactor
for the Vapour Phase Polymerization of Monomers at Different Hydrogen
Levels. US Patent 3,957,488, 1976.
Polymer Matrix: Polypropylene 461

10. J.L. Jezl and E.F. Peters, Horizontal Reactor for the Vapour Phase
Polymerization of Monomers. US Patent 4,129,701, 1978.
11. J.L. Jezl, E.F. Peters, R.D. Hall, J.W. Shepard, Process for the Vapour Phase
Polymerization of Monomers in a Horizontal, Quench-Cooled, Stirred-Bed
Reactor Using Essentially Total Off Gas Recycle and Melt Finishing. US
Patent 3,965,083, 1976.
12. E.F. Peters, M.J. Spangler, G.O. Michaels, J.L. Jezl, Vapour Phase Reactor Off-
Gas Recycle System for Use in the Vapour State Polymerization of Monomers.
US Patent 3,971,768, 1976.
13. Seppala, J.V., Harkonen, M., Luciani, L., Effect of the structure of external
alkoxysilane donors on the polymerization of propene with high activity
Ziegler–Natta catalysts. Makromol. Chem., 190, 2535–2550, 1989.
14. Aji, I.S., Sapuan, S.M., Zainudin, E.S., Abdan, K., Kenaf fibres as reinforce-
ment for polymeric composites: A review. Int. J. Mech. Mater. Eng. (IJMME),
4, 3, 239–248, 2009.
15. Kamal, I., Thirmizir, M.Z., Beyer, G., Jani Saad, M., Azrieda, N., Rashid, A.,
Kadir, Y.A., Kenaf for biocomposites: An overview, Universiti Ten Hussein
Onn Malaysia. J. Sci. Technol., 6, 2, 41–65, 2014.
16. Mohd, H.A.B., Abdu, A., Junejo, N., Abdul Hamid, H., Ahmed, K., Journey
of kenaf in Malaysia: A review. Sci. Res. Essays, 9, 458–470, 2014.
17. Rowell, R.M. and Han, J.S., Changes in kenaf properties and chemistry as
a function of growing time, in: Kenaf Properties, Processing and Products,
pp.  33–41, Mississippi State University, Hood Road, Mississippi, 1999,
Chapter 3.
18. Sanadi, A.R., Caulfield, D.F., Rowell, R.M., Reinforcing polypropylene with
natural fibers. Plast. Eng., 50, 4, 27–28, 1994.
19. Sanadi, A.R., Caulfield, D.F., Jacobson, R.E., Rowell, R.M., Renewable
agricultural fibres as reinforcing fillers of plastics: Mechanical properties
of kenaf fibres–polypropylene composites. Ind. Eng. Chem. Res., 34, 1889–
1896, 1995.
20. Chen, L., Colombus, E.P., Pote, J.W., Fuller, M.J., Black, J.G., Proceedings Int.
Kenaf Assoc. Conf., Irving, pp. 15–23, 1995.
21. Karnani, R., Krishnan, M., Narayan, R., Biofiber-reinforced polypropylene
composites. Polym. Eng. Sci., 37, 476–483, 1997.
22. Feng, D., Caulfield, D.F., Sanadi, A.R., Effect of compatibilizer on the
structure–property relationships of kenaf-fiber/polypropylene composites.
Polym. Compos., 22, 506–517, 2001.
23. Sanadi, A.R., Hunt, J.F., Caulfield, D.F., Kovacsvolgyi, G., Destree, B., High
Fibre–Low Matrix Composites: Kenaf Fibre/Polypropylene. Proceedings, the
Sixth International Conference on Woodfibre–Plastic Composites, Madison,
Wisconsin, May 2001, pp. 121–124, 2005.
24. Rashdi, A.A.A., Sapuan, S.M., Ahmad, M., Khalina, A., Review of kenaf fibre
reinforced polymer composites. Polimery, 54, 11–12, 2009.
462 Adhesives for Wood and Lignocellulosic Materials

25. Dansiri, N., Yanumet, N., Ellis, J.W., Ishida, H., Resin transfer molding of
natural fiber reinforced polybenzoxazine composites. Polym. Compos., 23,
352–360, 2002.
26. Anuar, N.I.S., Zakaria, S., Harun, J., Wang, C., Kenaf/PP and EFB/PP: Effect
of fibre loading on the mechanical properties of polypropylene composites.
Mater. Sci. Eng., 217, 1–10, 2017.
27. Shibata, S., Cao, Y., Fukumoto, I., Lightweight laminate composites made
from kenaf and polypropylene fibres. Polym. Test., 25, 142–148, 2005.
28. Rozman, H.D., Shannon-Ong, S.H., Azizah, A.B., Tay, G.S., Preliminary
study of non-woven composite: Effect of needle punching and kenaf fibre
loadings on non-woven thermoplastic composites prepared from kenaf and
polypropylene fibre. J. Polym. Environ., 21, 1032–1039, 2013.
29. David, N.V., Khairiyah, S., Anwar Majeed, P.P., Elastic and dynamic response
characteristics of kenaf/polypropylene composite, in: WIT Transactions on
the Built Environment, vol. 124, pp. 395–405, High Performance Structure
and Materials VI, WIT Press, Wessex, 2012.
30. Ismail, A., Hassan, A., Bakar, A.A., Jawaid, M., Flame retardancy and
mechanical properties of kenaf filled polypropylene (PP) containing ammo-
nium polyphosphate (APP). Sains Malays., 42, 4, 429–434, 2013.
31. Zhang, S. and Horrocks, A.R., A review of flame retardant polypropylene
fibres. Progr. Polym. Sci., 28, 1517–1538, 2003.
32. Sain, M., Park, S.H., Suhara, F., Law, S., Flame retardant and mechanical
properties of natural fibre–PP composites containing magnesium hydroxide.
Polym. Degrad. Stabil., 83, 363–367, 2004.
33. Lu, S. and Hamerton, I., Recent developments in the chemistry of halo-
gen-free flame retardant polymers. Progr. Polym. Sci., 27, 1661–1712, 2002.
34. Dimeski, D., Manov, Z., Srebrenkoska, V., Grozdanov, A., Bogoeva-Gaceva,
G., Avella, M., Zucchini, V., Kenaf fibre/polypropylene composites as poten-
tial material for partitioning panels in buildings. 12th European Conference
on Composite Materials, Biarritz, France, 29 Aug–1 Sept 2006, 2006.
35. Dhakal, H.N.Z., Zhang, Y., Richardson, M.O.W., Effect of water absorption
on the mechanical properties of hemp fibre reinforced unsaturated polyester
composites. Compos. Sci. Technol., 67, 16747–1683, 2007.
36. Nobrega, M.M.S., Cavalcanti, W.S., Carvalho, L.H., de Lima, A.G.B., Water
absorption in unsaturated polyester composites reinforced with caroà fibre
fabrics: Modelling and simulation. Mater. Sci. Eng. Technol., 41, 300–305,
2010.
37. Shenoy, M.A. and D’Melo, D.J., Effect of water on mechanical properties
of unsaturated polyester-acetylated hydroxypropyl guar gum composites.
J. Reinf. Plast. Compos., 28, 2561–2576, 2009.
38. Rashdi, A.A.A., Sapuan, S.M., Ahmad, M.M.H., Khalina, A., Combined
effects of water absorption due to water immersion, soil buried and natural
weather on mechanical properties of kenaf fibre unsaturated polyester com-
posites (KFUPC). Int. J. Mech. Mater. Eng., 5, 1, 11–17, 2010.
Polymer Matrix: Polypropylene 463

39. Haniffah, W.H., Sapuan, S.M., Abdan, K., Khalid, M., Hasan, M., Enamul
Hoque, M., Kenaf fibre reinforced polypropylene composites: Effect of cyclic
immersion on tensile properties. Int. J. Polym. Sci., 6 pages, 2015. Article ID
872387.
40. Wambere, P., Wens, J., Verpoest, I., Some Mechanical Properties of Kenaf/
Polypropylene Composites Prepared Using a Film Stacking Technique.
Proceedings, ICCM-16 Sixth International Conference on Composite Materials,
Kyoto, Japan, July 8–13, 2007, 2007.
41. Ridzuan, R., Beg, M.D.H., Rosli, M.Y., Rohaya, M.H., Astimar, A.A.,
Samahani, S., Zawawi, I., Effects of coupling agent and flame retardant on the
performances of oil palm empty fruit bunch fibre reinforced polypropylene
composites. Int. J. Mater. Metal. Eng., 7, 6, 1–9, 2013.
42. Kittikorn, T., Strömberg, E., Ek, M., Karlsson, S., The effect of surface modi-
fications on the mechanical and thermal properties of empty fruit bunch oil
palm fibre PP biocomposites. Polym. Renewable Resour., 3, 3, 79–100, 2012.
43. Kittikorn, T., Strömberg, E., Ek, M., Karlsson, S., Comparison of water uptake
as a function of surface modification of EFB oil palm fibres in polypropylene
biocomposites. Bioresources, 8, 2, 2998–3016, 2013.
44. Rozman, H.D., Tay, G.S., Kumar, R.N., Abusamah, A., Ismail, H., Mohd
Ishak, Z.A., Polypropylene–oil palm empty fruit bunch–glass fibre hybrid
composites: A preliminary study on the flexural and tensile properties. Eur.
Polym. J., 37, 1283–1291, 2001.
45. Rozman, H.D., Tay, G.S., Kumar, R.N., Abusamah, A., Ismail, H., Mohd
Ishak, Z.A., The effect of oil extraction of the oil palm empty fruit bunch
on the mechanical properties of polypropylene-oil palm empty fruit bunch–
glass fibre hybrid composites. Polym. Plast. Techn. Eng., 40, 2, 103–115, 2001.
46. Rozman, H.D., Kumar, R.N., Abdul Khalil, H.P.S., Abusamah, A., Lim, P.P.,
Ismail, H., Preparation and properties of oil palm from composite based on
methacrylic silane and glycidyl methacrylate. Eur. Polym. J., 33, 225–230,
1997.
47. Fatra, W., Rouhillahi, H., Helwani, Z., Muchtar, Z., Asmura, J., Effect of
alkaline treatment on the properties of oil palm empty fruit bunch fibre-
reinforced polypropylene composite. Int. J. Technol., 6, 1026–1034, 2016.
48. Czvikovszky, T., Electron-beam processing of wood fibre reinforced polypro-
pylene. J. Mech. Eng., 38, 209–224, 1994.
49. Czvikovszky, T., Electron-beam processing of wood fibre reinforced polypro-
pylene. Radiat. Phys. Chem., 35, 425–430, 1996.
50. Manarpaac, G.A., Zarnan, K., Harun, J., Oil Palm Empty Fruit Bunch Fibres–
Polypropylene Composites. Proceedings Pacific Rim Biobased Composites
Symposium, Canberra, Australia, 10–13 December 2000, 2000.
51. Khairul Zaman, H.M.D., Manarpaac, G.A., Jalaluddin, H., Electron
Beam Processing of Oil Palm Empty Fruit Bunch Fibers–Polypropylene
Composites. Proceedings, Takasaki Symposium on Radiation Application of
Natural Polymers in Asia, Jaeri, Takasaki, Japan, October 1 and 2, 2001, 2001.
464 Adhesives for Wood and Lignocellulosic Materials

52. Abdul Khalil, H.P.S., Poh, B.T., Issam, A.M., Jawaid, M., Ridzuan, R.,
Recycled polypropylene–oil palm biomass: The effect on mechanical and
physical properties. J. Reinf. Plast. Compos., 29, 1117–1130, 2010a.
53. Ahmed, A.S., Alam, M.A., Piee, A., Rahman, A.M.R., Hamdam, S., Study of
physical and mechanical properties of oil palm empty fruit bunch fibre rein-
forced polypropylene composites. J. Energy Environ., 2, 1, 16–21, 2010.
54. Abdul Khalil, H.P.S., Poh, B.T., Jawaid, M., Ridzuan, R., Suriana, R., Said,
M.R., Ahmad, F., Nik Fuad, N.A., The effect of soil burial degradation of
oil palm trunk fibre-filled recycled polypropylene composites. J. Reinf. Plast.
Compos., 29, 1653–1663, 2010b.
55. Khalid, M., Ratnam, C.T., Chuah, T.G., Ali, S., Choong, T.S.Y., Comparative
study of polypropylene composites reinforced with oil palm empty fruit
bunch fiber and oil palm derived cellulose. Mater. Des., 29, 1, 173–178, 2008.
56. ASTM D1104-56, Method of Test for Holocellulose in Wood, American
Society for Testing Materials, 1978.
57. Wirjosentono, B., Guritno, P., Ismail, H., Oil palm empty fruit bunch filled
polypropylene composites. Int. J. Polym. Mater. Po., 53, 295–306, 2004.
58. Jani, S.M., Rozman, H.D., Abusamah, A., Mohd Ishak, Z.A., Rahim, S., Oil
palm empty fruit bunch–polypropylene composites: The effect maleated
polypropylene on the mechanical properties. J. Oil Palm Res., 18, 260–271,
2006.
59. Rozman, H.D., Lai, C.Y., Ismail, H., Mohd Ishak, Z.A., The effect of coupling
agents on the mechanical and physical properties of oil palm empty fruit
bunch–polypropylene composites. Polym. Int., 49, 1273–1278, 2000.
60. Karuppuchamy, S., Andou, Y., Nishida, H., Nordin, N.I.A.A., Ariffin, H.,
Hassan, M.A., Shirai, Y., Superheated steam treated oil palm frond fibres and
their application in plastic composites. Adv. Sci. Eng. Medic., 7, 2, 120–125,
2015.
61. Abdan, K., Rheological Behaviours and Properties of Oil Palm (Elaeis
Guineensis, Jacq.) Empty Fruit Bunch Fibres/Polypropylene Composites, LAP
Lambert Academic Publishing, Saarbrucken, Germany, 2009.
62. Ramli, R., Yunus, R.M., Beg, M.D.H., Oil palm fiber reinforced polypropyl-
ene composites: Effects of fiber loading and coupling agents on mechanical,
thermal, and interfacial properties. J. Compos. Mater., 46, 1275–1284, 2012.
63. Zhu, J., Zhu, H., Njuguna, J., Abhyankar, H., Recent development of flax
fibres and their reinforced composites based on different polymeric matrices.
Materials, 6, 5171–5198, 2013.
64. Joseph, P.V., Mathew, G., Josepha, K., Thomas, S., Pradeep, P., Dynamic
mechanical properties of short sisal fibre reinforced polypropylene compos-
ites. Composites Part A, 34, 275–290, 2003.
65. Joseph, P.V., Mathew, G., Joseph, K., Thomas, S., Pradeep, P., Mechanical prop-
erties of short sisal fiber-reinforced polypropylene composites: Comparison
of experimental data with theoretical predictions. J. Appl. Polym. Sci., 88,
602–611, 2003.
Polymer Matrix: Polypropylene 465

66. Jayaraman, K., Manufacturing sisal–polypropylene composites with mini-


mum fibre degradation. Compos. Sci. Technol., 63, 367–374, 2003.
67. Pimenta, M.T.B., Carvalho, A.J.F., Vilaseca, F., Girones, J., López, J.P., Mutjé,
P., Curvelo, A.A.S., Soda-treated sisal/polypropylene composites. J. Polym.
Environ., 16, 35–39, 2008.
68. Markiewicz, E., Paukszta, D., Borysiak, S., Dielectric properties of lignocellu-
losic materials–polypropylene composites. Mater. Sci.-Poland, 27, 581–594,
2009.
69. Etaati, A., Pather, S., Rahman, M., Wang, H., Lejcek, P., Ground hemp fibres
as filler/reinforcement for thermoplastic biocomposites. Adv. Mater. Sci.
Eng., 2015, 1–11, 2015. Article ID 513590.
70. Zhao, W.J., Hu, Q.X., Zhang Y.C. Wei, N.N., Zhao, Q., Zhang, Y.M., Dong,
J.B., Sun, Z.Y., Liu, B.J., Li, L., Hu, W., In situ inorganic flame retardant mod-
ified hemp and its polypropylene composites. RSC Adv., 7, 32236–32245,
2017.
71. Cicala, G., Tosto, C., Latteri, A., La Rosa, A.D., Blanco, I., Elsabbagh, A.,
Russo, P., Ziegmann, G., Green composites based on blends of polypropylene
with liquid wood reinforced with hemp fibres: Thermo-mechanical proper-
ties and the effect of recycling cycles. Materials, 10, 9, 998, 2017.
72. Kechaou, B., Salvia, M., Beaugiraud, B., Juvé, D., Fakhfakh, Z., Treheux, D.,
Mechanical and dielectric characterization of hemp fibre reinforced polypro-
pylene (HFRPP) by dry impregnation process. eXPRESS Polym. Lett., 4, 3,
171–182, 2010.
73. Badji, C., Beigbeder, J., Garay, H., Bergeret, A., Bénézet, J.-C., Desauziers, V.,
Weathering of natural fibres reinforced composites: Study of the impact on
volatile organic compounds’ emissions in indoor air. Polym. Degrad. Stabil.,
149, 85–95, 2018a.
74. Zhang, X.H., Manufacturing of Hemp/PP Composites and Study of its Residual
Stress and Aging Behaviour, PhD Dissertation, Universite de Technologie de
Troyes, Troyes, France, 2016.
75. Badji, C., Beigbeder, J., Garay, H., Bergeret, A., Bénézet, J.-C., Desauziers, V.,
Correlation between artificial and natural weathering of hemp fibres rein-
forced polypropylene biocomposites. Polym. Degrad. Stabil., 148, 117–131,
2018b.
76. Panthapulakkal, S. and Sain, M., Injection-molded short hemp fiber/glass
fiber-reinforced polypropylene hybrid composites—Mechanical, water
absorption and thermal properties. J. Appl. Polym. Sci., 103, 2432–2441,
2007.
77. Sullins, T., Pillay, S., Komus, A., Ning, H., Hemp fibre reinforced polypropyl-
ene composites: The effects of material treatments. Composites Part B, 114,
15–22, 2017.
78. Hargitai, H., Rácz, I., Anandjiwala, R.D., Development of HEMP fibre rein-
forced polypropylene composites. J. Thermoplast. Compos. Mater., 21, 2,
165–174, 2008.
466 Adhesives for Wood and Lignocellulosic Materials

79. Puech, L., Ramakrishnan, K.R., Le Moigneau, N., Corn, S., Slangen, P.R.,
LeDuc, A., Boudhani, H., Bergeret, A., Investigating the impact behaviour
of short hemp fibres reinforced polypropylene biocomposites through high
speed imaging and finite element modelling. Composites Part A, 109, 428–
439, 2018.
21
Biodegradable Polymers
as Matrix for Biocomposites

21.1 Introduction
Biopolymers are mainly classified as agro-polymers (starch, cellulose), bio-
degradable polyesters, namely, polyhydroxyalkanoates (PHAs), obtained
by microbiological production, and poly(lactic acid) synthesized chemi-
cally using monomers obtained from agro-resources [1].
The most emerging renewable resource-based bioplastics are polylac-
tides (PLA) and PHAs. Under the PHA family of bioplastics, there are two
important polymers, namely, polyhydroxybutyrate-co-valerate (PHBV)
and polyhydroxybutyrate (PHB) [2].
Yu and Chen [3] and Nagarajan [4] have reported that the production of
1 kg of PHA resin contributed to only 0.49 kg of CO2 emission as opposed
to 2–3 kg of CO2 emissions by the production of the same amount of
petroleum-based plastics. They have also mentioned that this is about 80%
reduction in the global warming potential. The fossil energy required for
bioplastics production (44 MJ/kg) is also nearly half of that required for the
petro-based plastic (78–88 MJ/kg) [3, 4].

21.2 Polyhydroxyalkanoates
Since the mechanical, physical, and thermal characteristics of PHA biopoly-
mers are similar to many fossil-fuel-based plastics, such as polypropylene
(PP) and polyethylene (PE), research on these biopolymers has intensified in
recent times. They have the potential to replace the petroleum-derived bulk
polymers in a number of applications [5, 6].
PHAs are natural polyesters that are produced by bacteria and extensive
variety of microorganisms as intracellular inclusions meant for carbon and
energy storage [7, 8]. Thus, PHA polymers are true “biopolymers”, as they
are synthesized within the bodies of bacteria grown in special digesters

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (467–481) © 2019
Scrivener Publishing LLC

467
468 Adhesives for Wood and Lignocellulosic Materials

and subsequently extracted from the biomass. The polymer can make up
to 80% of the biomass, stored by the bacteria as an energy reserve [9]. Due
to their biocompatibility and biodegradability, PHAs have a wide range of
applications in various industries.
PHAs are green plastics and they have positive social and environmental
impact when compared with conventional plastics in terms of production
and recycling. They do not possess acute and chronic health effects when
used in vivo. They are a renewable and sustainable resource. Progress in
the recovery techniques has improved the extraction efficacy from biomass
with high purity [10].
Although production costs have decreased substantially over the last
several decades, PHA prices are still significantly higher than traditional
plastics [11].
A simplified structure of PHA is given in Figure 21.1 [12].
PHA synthesis is promoted by unbalanced growth during the fermen-
tation and accumulation of PHA granules as part of a survival mechanism
of the microbes [13]. The general molecular structure of PHAs is presented
in Figure 21.1b. Depending on the carbon numbers in the monomeric con-
stituents, PHAs can be classified as short-chain-length PHAs that consist
of 3–5 carbon monomers, and medium-chain-length PHAs (MCL-PHA)
that consists of 6–14 carbon monomers in the 3-hydroxyalkanoate units
(Figure 21.1b) [8, 14].
More than 150 different PHA monomers have been identified, which
renders them the largest group of natural polyesters [15].

(a) R O
H
O x OH
n

(b)

O O O O O O

O O O O O O
3HB 3HV 3HHx 3HO 3HD 3HDD

SCL-HA MCL-HA

Figure 21.1 Simplified structures of PHAs.


Biodegradable Polymers for Biocomposites 469

21.2.1 Poly(3-hydroxybutyrate) PHB


Among these PHBs, poly(3-hydroxybutyrate) [P(3HB)] was the first and
most common type of PHA to be identified by the French microbiologist
Maurice Lemoigne. P(3HB) is an optically active biological linear polyes-
ter that is insoluble in water and exhibits a high degree of polymerization
that ranges from 105 to approximately 107. The biosynthesized P(3HB) is
thus perfectly isotactic and, upon extraction from the microorganisms,
has a crystallinity of about 55–80% with a melting point at around 180°C
[16–18].
PHB has attracted much attention as a biodegradable thermoplastic
polyester and has high potential as an environmentally degradable plas-
tic as it can degrade through various types of bacteria and fungi into car-
bon dioxide and water through secreting enzymes [19–22]. It can also be
degraded through nonenzymatic hydrolysis.
PHB can also be synthesized by ring-opening polymerization (ROP) of
β-butyrolactone using distannoxane derivatives as catalysts, such as zinc
and aluminum [23].
PHB belongs to a PHA class of shorter pendant groups that confers a
high degree of crystallinity [18].
PHB can be processed into pellets that can be handled on plastics
machines in the same way as traditional plastics produced from oil. Poor
thermal stability and relatively low impact resistance reduce its process-
ability [24].
The glass transition temperature (Tg) is around 55°C [25]. The strength
and stiffness of PHB is closer to that of PP [26]. Doi et al. found that PHB
had a perfectly isotactic structure with R-configuration, which was respon-
sible for the high crystallinity of PHB [27]. It is hard, brittle, stiff, and highly
crystalline [25, 28].
Further, the high melting point [29] and a narrow processing window
and susceptibility to thermal degradation at temperatures close to its melt-
ing point [25] are disadvantages. In contrast, PHBV has better flexibility
and toughness and has lower melting and glass transition temperatures,
depending on PHV content [30].
Despite high prices, there are many areas where PHB is used. The US
Navy opted to use PHB cups, which can be easily thrown overboard after
use and degrade in the sea. In Japan, PHB is being used for manufacture of
women’s disposable razor [19].
470 Adhesives for Wood and Lignocellulosic Materials

O CH3 O H3C CH3

C CH2 CH O C CH2 CH CH2 O


x y
HB HV

Figure 21.2 Polyhydroxybutyrate-co-valerate.

21.2.2 Poly(3-hydroxybutyrate-co-3-hydroxyvalerate)
To overcome the shortcomings of PHB, copolymers of PHB were
developed. One of the copolymers of PHB is poly(3-hydroxybutyrate-
co-3-hydroxyvalerate) (Figure 21.2), PHBV with randomly arranged 3-
hydroxybutyrate (HB) and 3-hydroxyvalarate (HV) groups. It was first
synthesized by ICI in 1983 [31]. PHBV is a biopolymer with high crystal-
linity having a melting point of 180°C. The glass transition temperature
(Tg) of the polymer is in the range −5°C to 20°C [32]. The flexibility and
processability of PHBV is better than PHB. Elongation at break for PHBV
is lower than 15% and elastic modulus is 1.2 GPa. The valerate content in
the PHBV strongly influences the crystallinity, melting point, and rate of
crystallization of the copolymer [33]. The copolymer melting point, glass
transition temperature, and crystallinity decrease as the hydroxyvalerate
unit content increases. With an increase of the HV units in the copo-
lymer, the impact strength increases and the tensile strength decreases
[34, 35]. The rate of degradation of PHBV is faster than that of PHB. The
degradation kinetics was found to depend on the structure (copolymer
or homopolymer), crystallinity, and processing conditions [36].

21.3 Polylactic Acid


Polylactic acid, known as PLA, is a unique polymer that in many ways
behaves like poly(ethylene terephthalate) (PET), but also performs a lot
like PP.
It is one of the fully biodegradable polymers that is commercially avail-
able. It is a highly versatile, biodegradable, aliphatic polyester derived from
100% renewable resources [37].
It is produced from lactic acid, a naturally occurring organic acid that
can be produced by fermentation.
Polylactide or polylactic acid (PLA) is a highly versatile biodegrad-
able polymer that has been derived from renewable resources like corn.
Biodegradable Polymers for Biocomposites 471

O
O
C OH C
OH
OH OH

C C

CH3 H
H CH3

L-Lactic acid D-Lactic acid

Figure 21.3 D- and L-lactic acid.

The hydrolysis of cornstarch yields sugars that can be fermented to lactic


acid followed by polymerization to yield PLA. PLA is now produced com-
mercially in different parts of world. NatureWorks LLC in the United States
is the lead producer of PLA now globally. Other companies are Teijin in
Japan, HISUN in China, PURAC in the Netherlands, and Inventa Fischer
in Germany [2, 38, 39].
Lactic acid exists as two isomers, namely, L-lactic acid and D-lactic acid,
as shown in Figure 21.3.
Pure L-lactic acid or D-lactic acid, or mixtures of both the compounds
are needed for the synthesis of PLA. The homopolymer of LA is a white
powder at room temperature with Tg and Tm values of about 55°C and
175°C, respectively. High-molecular-weight PLA is a colorless, glossy, rigid
thermoplastic material with properties similar to polystyrene. The two iso-
mers of LA can produce four distinct materials.

(1) Poly(D-lactic acid) (PDLA), a crystalline material with a


regular chain structure;
(2) Poly(L-lactic acid) (PLLA), which is hemicrystalline, and
likewise with a regular chain structure;
(3) Poly(D,L-lactic acid) (PDLLA), which is amorphous; and
(4) meso-PLA, obtained by the polymerization of meso-lactide.

21.3.1 Synthesis of PLA


The synthesis of PLA has been exhaustively reviewed by Xiao et al. [40].
Two main synthetic methods are used to obtain PLA:

(1) Direct polycondensation (including solution polyconden-


sation and melt polycondensation), and
(2) Ring-opening polymerization.
472 Adhesives for Wood and Lignocellulosic Materials

21.3.2 Direct Polymerization


Since the lactic acid monomer has both –OH and –COOH groups, nec-
essary for polymerization, the reaction can take place directly by self-
condensation as shown in Figure 21.4.

21.3.2.1 Solution Polycondensation


In this case, an organic solvent capable of dissolving the PLA without inter-
fering with the reaction is added, and the mixture is refluxed with removal
of the water generated in the polycondensation process, which is beneficial
to achieve a high molecular weight. PLA has a weight-average molecular
weight (Mw) of over 200,000 via this method [41, 42].

21.3.2.2 Melt Polycondensation


In contrast to solution polycondensation, the melt polycondensation of
monomers can proceed without any organic solvent [43]. Moon et al.
found that high Mw PLLA [Mw ≥ 100,000] could be produced in this way in
a relatively short reaction time (≤15 h) [44].
Water produced through this process should be continuously removed
to prevent the reversible reaction to occur and allow chain extension to
take place to obtain high molecular weight. Thus, the chain-extension
reaction is an effective method to obtain PLA with high molecular weight
by polycondensation [44].
This method can lower the cost of the synthesis significantly due to the
simplified procedure, but major problems still need to be solved before
it can be applied industrially because of its sensitivity to reaction condi-
tions. Thus, Moon et al. [43] worked to develop a melt/solid polyconden-
sation technique using a binary catalyst system (tin dichloride hydrate
and p-toluenesulfonic acid) [40].

CH3 CH3
Polymerization
HO C COOH H O C COO H + (n–1) H2O
Catalyst

H H n

Lactic acid Poly(lactic acid)

Figure 21.4 Polyesterification of lactic acid.


Biodegradable Polymers for Biocomposites 473

21.3.2.3 Ring-Opening Polymerization


PLA can be synthesized by ROP, an important and effective method to
manufacture high-molecular-weight PLA (Figure 21.5) [40, 45]. This reac-
tion requires high purity of the lactide monomer, obtained by dimerization
of the lactic acid monomer. PLA is obtained by using a catalyst along with
the monomer under vacuum or an inert atmosphere. By controlling the
residence time and the temperatures in combination with the catalyst type
and concentration, it is possible to control the ratio and sequence of D- and
L-lactic acid (LA) units in the final polymer [46].
Many kinds of initiators are known, but stannous octoate is used pre-
dominantly, as it provides high reaction velocity, high rate of transforma-
tion, high molecular weight, and relatively mild reaction condition [47].
Many nontoxic catalysts derived from magnesium, calcium, zinc, alkali
metals, and aluminum have been developed for the ROP of lactides to
solve pollution problems caused by heavy metal catalysts [48].
PLA has received particular attention for biocomposites manufacturing.
PLA is probably the most promising matrix because its properties are close
to the most widespread thermoplastic matrix of PP and PET.
The homopolymer of LA is a white powder at room temperature with
Tg and Tm values of about 55°C and 175°C, respectively. High-molecular-
weight PLA is a colorless, glossy, rigid thermoplastic material with proper-
ties similar to polystyrene.
The attractive price and commercial availability of lactic acid are import-
ant reasons for PLA development [49].

CH3 O
H3C O
Cyclization
HO C COOH
–H2O
O O CH3
H
Lactic acid Lactide

Ring-opening polymerization Catalyst

CH3

O C COO

H n

Figure 21.5 ROP of lactic acid.


474 Adhesives for Wood and Lignocellulosic Materials

O
O O
O O
O m O n

Figure 21.6 Polybutylene adipate terephthalate.

21.4 Polybutylene Adipate Terephthalate


PBAT (short for polybutylene adipate terephthalate) is a biodegradable
random copolymer, specifically a copolyester of adipic acid, 1,4-butanediol
and terephthalic acid  (from  dimethyl terephthalate). PBAT is produced
by many different manufacturers and may be known by the brand names
Ecoflex , Wango, Ecoworld, Eastar Bio, and Origo-Bi. It is also called
poly(butylene adipate-co-terephthalate). It is generally marketed as a fully
biodegradable alternative to low-density PE, having many similar proper-
ties including flexibility and resilience, allowing it to be used for many sim-
ilar uses such as plastic bags and wraps. The structure of the PBAT polymer
is shown in Figure 21.6.
PBAT is an aliphatic–aromatic copolyester commercialized by BASF
under the trade name Ecoflex .

21.5 All Green Composites


Development of all green composites by combining biodegradable poly-
mers derived from renewable resources and natural fibers has greatly
attracted researchers in recent years. The most important feature of green
composites is their complete biodegradability.
Biopolymers have been filled with natural fillers such as wood flour,
kenaf, flax, jute, hemp, sisal, and kapok, which could bring an improve-
ment in some properties such as toughness, stiffness, and also thermal
stability. Natural fibers are biodegradable and renewable in perpetuity
and the biocomposites based on biopolymers and natural fibers consti-
tute “All Green Composites” and can be employed for specific end uses.
It is believed that biocomposites can supplement and eventually replace
petroleum-based composite materials in several applications such as
packaging, automotive industry, construction materials, furniture, and
consumer goods [19].
Biodegradable Polymers for Biocomposites 475

As a matrix, PHB Biomer P226 was used, a biodegradable polyester


produced by the German company Biomer.
Kuciel and Liber-Knec have explored the possibilities of processing
biocomposites based on PHB filled with 25 and 40 wt% of wood flour
or kenaf fibers [19]. They determined the mechanical properties (tensile
strength, modulus of elasticity) of the composites under three testing tem-
peratures. Water absorption and its influence on mechanical properties
were determined. The process of biodegradation of the composites was
evaluated in the garden compost. Scanning electron microscope (SEM)
images of fractured composites confirmed good adhesion between natural
fibers and PHB. Modulus of elasticity and tensile strength increased with
rising content of wood flour and kenaf fibers. The results show that PHB
filled with kenaf fibers have good mechanical properties. Biocomposites
absorb water and their mechanical properties decrease after exposure to
water. It may be caused by the onset of the biodegradation process. This
property can be interesting for packaging especially for fresh produce like
fruits or vegetables and for industrial products with short life cycles.
The incorporation of bioresources, e.g., crop-derived green plastics and
plant-derived natural fibers, into composite structures is gaining prime
importance in designing and engineering of green composites. Garkhail
et al. employed renewable resources for the development of biodegrad-
able composite materials [50]. Two types of biocomposites were made
using flax fibers as a reinforcement and poly[(R)-3-hydroxyalkanoates]
(PHA) as a biodegradable polymer: First, natural-fiber-mat-reinforced
thermoplastics was prepared by compression molding and needle-
punched method to produce nonwoven flax fiber mats, and second,
injection molding compounds based on short flax fibers. The influence
of fiber content and processing method on the tensile and impact proper-
ties of these composites was studied. Results indicated that the addition
of inexpensive flax fiber to PHB could be advantageous as far as cost per-
formance of these materials is concerned. For example, the addition of
flax fibers to the relatively brittle PHB matrix offers the chance to obtain
cheaper products together with improved toughness, while retaining
biodegradability of the resulting polymer composites. In the case of rel-
atively long fibers, the improvement in impact resistance is reported to
be significant.
The poor toughness of PHB originates from the relatively high glass
transition temperature and its tendency to form large spherulites. As a
result, PHBV copolymers (Figure 21.7) have been developed to increase
toughness.
476 Adhesives for Wood and Lignocellulosic Materials

O=
CH
C
CH2 O
n
R = CH3 Hydroxy Butyrate
= C2H5 Hydroxy Valerate

Figure 21.7 PHB/valerate.

Maleic anhydride (MAH) was used as the grafting monomer. The


graft copolymer was prepared by melt grafting reaction in the twin-screw
extruder with dicumyl peroxide as the initiator. Polylactic acid grafted with
maleic anhydride (MAH-g-PLA) was successfully prepared as the interface
compatibilizer [51]. The PLA/wood fiber/MAH-g-PLA composites were
prepared by melt blending and injection molding with different propor-
tions of compatibilizer added, within which PLA was for the matrix phase
and wood fiber was for the reinforcing phase. The crystallinity, micro-
structure, thermal stability, and dynamic thermomechanical property of
the composites were studied by X-ray diffraction, SEM, thermogravimetric
analyzer, and dynamic mechanical thermal analysis. The results of all the
tests showed that the addition of MAH-g-PLA could improve the interfa-
cial compatibility of PLA/wood fiber composites and improve the mechan-
ical properties of the composites.
In a recent work, two different macromolecular coupling agents were
synthesized [52], namely,

(a) MPS-g-PLA (polylactide  graft  γ-methacryloxypropyltrime-


thoxysilane)
(b) PLA-co-PGMA (polylactide-co-glycidyl methacrylate)

The novel copolymer of polylactide and glycidyl methacrylate (PLA-co-


PGMA) was prepared by free radical polymerization [53]. MPS grafting
PLA was carried out using benzoyl peroxide as the free radical initiator
[54].
Surface modification of sisal fiber by the MPS-g-PLA treatment was car-
ried out by immersing fibers in the dioxane solution of MPS-g-PLA at the
concentration of 1 wt% for 48 h using acetic acid to adjust pH. Sisal fibers
were taken out of the solution and dried for 2 days at room temperature.
Sisal fibers are then reacted with MPS-g-PLA at 120°C for 2 h.
Biodegradable Polymers for Biocomposites 477

Surface modification of sisal fiber surface by PLA-co-PGMA treatment


was carried out by immersing sisal fiber in the dioxane solution of PLA-
co-PGMA at the concentration of 1 wt% for 48 h. After drying at room
temperature, the products were Soxhlet extracted with THF for 24 h and
then dried.
The effect of surface treatment of sisal fiber with MPS-g-PLA and
PLA-co-PGMA on the mechanical properties was also evaluated, and
the results suggested that they were efficient in modifying natural fiber
surface and in improving the compatibility of PLA/cellulose composites
[53, 54].
Asokan et al. summarized the work carried out carried out on the bio-
composites made from sisal, flax, and hemp fibers [55]. The results included
the tensile and flexural strength, flexural modulus elongation at break, and
impact strength [56–61].

References
1. Platt, D.K., Biodegradable Polymers, Market Report, Smithers Rapra Limited,
2006.
2. Misra, M., Nagarajan, V., Reddy, J., Mohanti, A.K., Bioplastics and Green
Composites From Renewable Resources: Where We Are and Future
Directions. 18th International Conference on Composites Materials (ICCM),
Jeju Island, Korea, 21st to 26th August 2011, 2011.
3. Yu, J. and Chen, L.X.L., The greenhouse gas emissions and fossil energy
requirement of bioplastics from cradle to gate of a biomass refinery. Environ.
Sci. Technol., 42, 6961–6966, 2008.
4. Nagarajan, V., Sustainable Biocomposites from ‘Green’ Plastics and Natural
Fibres, PhD Dissertation, The University of Guelph, Ontario, Canada, 2012.
5. Ferreira, B.S. and Schlottbom, C., Production of Polyhydroxybutyrate
from Lignocellulosic Hydrolysates—Optimization of Bacillus sacchari
Fermentation and Scale Up from 2 L to 200 L, Eppendorf, Application
Note No. 302, 2016.
6. Choi, J. and Lee, S.Y., Factors affecting the economics of polyhydroxyal-
kanoate production by bacterial fermentation. Appl. Microbiol. Biotechnol.,
51, 13–21, 1998.
7. Li, Z., Yang, J., Loh, X.J., Polyhydroxyalkanoates: Opening doors for a sus-
tainable future. NPG Asia Mater., 8, 265, 2016.
8. Li, Z. and Loh, X.J., Water soluble polyhydroxyalkanoates: Future materials
for therapeutic applications. Chem. Soc. Rev., 44, 2865–2879, 2015.
9. Anonymous, https://www.smithersrapra.com/SmithersRapra/media/Sample-
Chapters/Mouldable-Particle-Foams.pdf.
478 Adhesives for Wood and Lignocellulosic Materials

10. AliRaza, Z., Abid, S., Banat, I.M., Polyhydroxyalkanoates: Characteristics,


production, recent developments and applications. Int. Biodeter. Biodegr.,
126, 45–56, 2018.
11. Ravenstijn, J., PHA. Is it here to stay? Presentation, 2014. http://www.kcpk.
nl/algemeen/bijeenkomsten/presentaties/20140508-jan-ravenstijn-pha-is-it-
here-to-stay.
12. Yu, L., Dean, K., Li, L., Polymer blends and composites from renewable
resources. Prog. Polym. Sci., 31, 576–602, 2006.
13. Chen, G.Q., A microbial polyhydroxyalkanoates (PHA) based bio- and
materials industry. Chem. Soc. Rev., 38, 2434–2446, 2009.
14. Kai, D., Low, Z.W., Liow, S.S., Development of lignin supramolecular hydro-
gels with mechanically responsive and self-healing properties. ACS Sustain.
Chem. Eng., 3, 2160–2169, 2015.
15. Kim, D.Y., Kim, H.W., Chung, M.G., Rhee, Y.H., Biosynthesis, modifica-
tion, and biodegradation of bacterial medium-chain-length polyhydroxyal-
kanoates. J. Microbiol., 45, 87–97, 2007.
16. Lemoigne, M., Products of dehydration and polymerization of b-hydroxybu-
tyric acid. Bull. Soc. Chim. Biol., 8, 770–782, 1926.
17. Rosa, D.S. and Lenz, D.M., Biocomposites: Influence of matrix nature and
additives on the properties and biodegradation behaviour, in: Biodegradation—
Engineering and Technology, R. Chamy and F. Rosenkranz (Eds.), pp. 432–475,
InTech Croatia, Rijeka, Croatia, 2013.
18. Quental, A.C., de Carvalho, F.P., Rezende, M.I., Rosa, D.S., Felisbert, M.I.,
Aromatic/aliphatic polyester blends. J. Polym. Environm., 18, 308–317, 2010.
19. Kuciel, S. and Liber-Knec, A., Biocomposites based on PHB filled with wood
or kenaf fibers. Polimery, 56, 3, 218–223, 2011.
20. Karlsson, J.O., Gatenholm, P., Peguy, A., Blachot, J.F., Improvement of adhe-
sion between polyethylene and regenerated cellulose fibre by surface fibrilla-
tion. Polym. Comp., 17, 300–304, 1996.
21. Zini, E., Scandola, M., Gatenholm, P., Heterogeneous acylation of flax fibers.
Reaction kinetics and surface properties. Biomacromolecules, 4, 821–827, 2003.
22. Felix, J.M. and Gatenholm, P., Formation of entanglements at brushlike
interfaces in cellulose–polymer composites. J. Appl. Polym. Sci., 50, 699–708,
1993.
23. Corrêa, M.C.S., Rezende, M.L., Rosa, D.S., Agnelli, J.A.M., Nascente, P.A.P.,
Surface composition and morphology of poly(3-hydroxybutyrate) exposed
to biodegradation. Polym. Test., 27, 447–452, 2008.
24. Frollini, E., Leao, A.L., Mattoso, L.H.C., Natural Polymers and Agrofibres
Based Composites, Embrapa, San Carlos, S. P. Brazil, 2000.
25. Vroman, I. and Tighzert, L., Biodegradable polymers. Materials, 2, 307–344,
2009.
26. van de Velde, K. and Beatens, E., Thermal and mechanical properties of flax
fibres as potential composite reinforcement. Macromol. Mater. Eng., 286, 6,
342–349, 2001.
Biodegradable Polymers for Biocomposites 479

27. Doi, Y., Kitamura, S., Abe, H., Microbial synthesis and characterization of
poly(3-hydroxybutyrate-co-3-hydroxyhexanoate). Macromolecules, 28, 4822–
4828, 1995.
28. Wang, Q., Liu, X., Qi, Q., Biosynthesis of poly(3-hydroxybutyrate-co-3-
hydroxyvalerate) from glucose with elevated 3-hydroxyvalerate fraction
via combined citramalate and threonine pathway in Escherichia coli. Appl.
Microbiol. Biotechnol., 98, 3923–3931, 2014.
29. Xing, P., Dong, L., An, Y., Feng, Z., Avella, M., Martuscelli, E., Miscibility and
crystallization of poly(hydroxybutyrate) and poly(p-vinylphenol) blends.
Macromolecules, 30, 2726–2733, 1999.
30. Marchessault, R.H. and Yu, G.-E., Crystallization and material properties of
polyhydroxyalkanoates. Biopol. Online, 7, 157–166, 2005.
31. Loo, C. and Sudesh, K., Polyhydroxyalkanoates: Bio-based microbial plastics
and their properties. Malays. Polym. J., 3, 1, 31–57, 2007.
32. Avella, M., Immirzi, B., Malinconico, M., Martuscelli, E., Volpe, M.G.,
Reactive blending methodologies for biopol. Polym. Int., 39, 191–204,
1996.
33. Kunioka, M., Tamaki, A., Doi, Y., Crystalline and thermal properties of
bacterial copolyesters: Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) and
poly(3-hydroxybutyrate-co-4-hydroxybutyrate). Macromolecules, 22, 694–
697, 1989.
34. Hsieh, W., Wada, Y., Mitobe, T., Mitomo, H., Seko, N., Tamada, M., Effect of
hydrophilic and hydrophobic monomers grafting on microbial poly(3-hy-
droxybutyrate). J. Taiwan Inst. Chem. Eng., 40, 413–417, 2009.
35. Amass, W., Amass, A., Tighe, B., A review of biodegradable polymers:
Uses, current developments in the synthesis and characterization of biode-
gradable polyesters, blends of biodegradable polymers and recent advances
in biodegradation studies. Polym. Int., 47, 89–144, 1998.
36. Parikh, M., Gross, R.A., MacCarthy, S.P., The influence of injection moulding
conditions on biodegradable polymers. J. Injection Moulding Technol., 2, 1,
32–36, 1998.
37. Drumwright, R.E., Gruber, P.R., Henton, D.E., Polylactic acid technology,
Adv. Materials, 12, 1841–1846, 2000.
38. Shen, L., Haufe, J., Patel, M.K., Product overview and market projec-
tion of emerging biobased plastics, Report PRO-BIP 2009, the European
Polysaccharide Network of Excellence and European Bioplastics, 2009.
39. Mohanty, A., Bioplastics, biobased plastics and bioplastic alloys: Current
trends and future prospective, in: Bioplastics and Green Composites Workshop,
Guelph, Ontario, Canada, March 31–April 1 2010.
40. Xiao, L., Wang, B., Yang, G., Gauthier, M., Poly(lactic acid)-based biomateri-
als: Synthesis, modification and applications. InTech Open Sci., 11, 247–282,
2011.
41. M. Ohta, S. Obuchi, Y. Yoshida, Preparation process of polyhydroxycarbox-
ylic acid. US Patent 5,444,143, 1995.
480 Adhesives for Wood and Lignocellulosic Materials

42. F. Ichikawa, M. Kobayashi, M. Ohta, Y. Yoshida, S. Obuchi, H. Itoh, Process


for preparing polyhydroxycarboxylic acid. US Patent 5,440,008, 1995.
43. Moon, S.I., Lee, C.W., Miyamoto, M., Kimura, Y., Melt polycondensation of
L-lactic acid with Sn(II) catalysts activated by various proton acids: A direct
manufacturing route to high molecular weight poly(L-lactic acid). J. Polym.
Sci., Part A: Polym. Chem., 2000. https://doi.org/10.1002/(SICI)1099-0518
(20000501)38:9<1673::AID-POLA33>3.0.CO;2-T.
44. Gu, S.Y., Yang, M., Yu, T., Ren, T.B., Ren, J., Synthesis and characterization of
biodegradable lactic acid-based polymers by chain extension. Polym. Int., 57,
982–986, 2008.
45. Bardone, E., Bravi, M., Keshavarz, T., Synthesis and characterizations of
poly(lactic acid) by ring-opening polymerization for biomedical applica-
tions. Chem. Eng. Transact., 38, 331–336, 2014.
46. Lunt, J., Large-scale production, properties and commercial applications of
polylactic acid polymers. Polym. Degrad. Stabil., 59, 145–152, 1998.
47. Mehta, R., Kumar, V., Bhunia, H., Upadhyay, S., Synthesis of poly(lactic acid):
A review. J. Macromol. Sci. Part C, Polym. Rev., 45, 325–349, 2005.
48. Lee, C. and Hong, S., An overview of the synthesis and synthetic mechanism
of poly(lactic acid). Modern. Chem. Appl., 2, 4, 144, 2014.
49. Lopes, M.S., Jardini, A.L., Filho, R.M., Synthesis and characterizations of
poly(lactic acid) by ring-opening polymerization for biomedical applica-
tions. Chem. Eng. Transact., 38, 331–336, 2014.
50. Garkhail, S.K., Meurs, E., Van de Beld, T., Peijs, T., Thermoplastic composites
based on biopolymers and natural fibres. Compos. Mater., ICCM-12 Europe
12th International Conference on Composite Materials, 1–10, 1999.
51. Zhang, L., Shanshan, L., Sun, C., Wan, L., Tan, H., Zhang, Y., Effect of MAH-
g-PLA on the properties of wood fibre/polylactic acid composites. Polymers,
9, 11, 591, 2017.
52. Li, Z., Zhou, X., Pei, C., Effect of Sisal Fibre Surface Treatment on Properties
of Sisal Fibre Reinforced Polylactide Composites. Int. J. Polym. Sci., 7, 2011.
Article ID 803428.
53. Li, Z., Zhou, X., Pei, C., Synthesis and characterization of MPS-g-PLA
copolymer and its application in surface modification of bacterial cellulose.
Int. J. Polym. Anal. Character., 15, 4, 199–209, 2010.
54. Li, Z.Q., Zhou, X., Pei, C., Synthesis of PLA-co-PGMA copolymer and its
application in the surface modification of bacterial cellulose. Int. J. Polym.
Mater., 59, 725–737, 2010.
55. Asokan, P., Firdoous, M., Sonal, W., Properties and potential of bio fibres,
bio-binders, and bio composites. Rev. Adv. Mater. Sci., 30, 254–261, 2012.
56. Thomas, S. and Pothan, L.A., Natural Fibre Reinforced Polymer Composites:
From Macro to Nanoscale, Editions des archives contemporaines, Paris, and
Old City Publishing, Philadelphia, PA, 2009.
Biodegradable Polymers for Biocomposites 481

57. Oksman, K., Skrifvars, M., Selin, J.-F., Natural fibres as reinforcement in
polylactic acid (PLA) composites. Compos. Sci. Technol., 63, 1317–1324,
2003.
58. Bax, B. and Mussig, J., Impact and tensile properties of PLA/Cordenka and
PLA/flax composites. Compos. Sci. Technol., 68, 1601–1622, 2009.
59. Bodros, E., Pillin, I., Montrelay, N., Baley, C., Could biopolymers rein-
forced by scattered flax fibre be used in structural applications? Compos. Sci.
Technol., 67, 462–470, 2007.
60. Avella, M., Buzarovska, A., Errico, M.E., Gentile, G., Grozdanov, A., Eco-
challenges of bio-based polymer composites. Materials, 2, 911–925, 2009.
61. Bergeret, A., Environmental-friendly biodegradable polymers and compos-
ites, in: Integrated Waste Management, vol. 1, S. Kumar (Ed.), InTechopen,
Wien, Austria, 2011.
Index

Acetylacetone, 308 Lignin adhesives, 275–277


Adhesion Protein adhesives, 277
Bonding Tannin adhesives, 268–275
Chemical bonds, 38 Green or Sustainable Epoxy
Ionic bonding, 34, 37–38 adhesives, 417–419
Hydrogen bonds, 34, 37 New industrial technologies, 269,
Contact angle, 44, 45, 50, 51 270
Surface roughness effect, 50–51 Tannins-hexamine, 270–272
Critical surface tension, 50 Tannins autocondensation, 272–275
Interactions, 34, 35 Unsaturated oil adhesives, 279–283
Apolar, 34 Wood welding without adhesives,
Debye, 34 283–287
Keesom, 34 Adhesives rheology, 317–345
London, 34, 35 Dynamic mechanical analysis,
Van der Waal, 34 328–332
Dipole-Dipole, 35, 36 Experimental results, 338–342
Dipole-Ionic, 36 Newtonian and Non-Newtonian
Mechanism of, 33, 34 Fluids, 324, 325
Specific adhesion, 34, 35 TMA and DMA analysis, 339–342
Spreading coefficient, 49, 50 TTT and CHT diagrams, 332–338
Theories of, 38–43 Viscoelasticity, 325–328
Adsorption, 42–51 Viscometers
Covalent bonding, 56 Capillary, 318–320
Electronic/electrostatic, 41–42 Contentric cylinder, 322
Mechanical interlocking, 39–41 Cone and plate, 321, 322
Weak boundary layers, 51–55 Ford cup, 322, 323
Wetting Gardner-Holt tubes, 323, 324
Of substrate, 43, 45 Parallel plate, 322
Work of adhesion, 45–49 Rotational, 320, 321
Zisman’s rectilinear relationship, 50 Viscosity theory, 317, 318
Adhesives from bioresources, 267–292 Amines, 406–412
Biobased polyurethane adhesives, Anhydrides, 412–414
185–192 Chromatography, 84
Carbohydrate adhesives, 279, 280 Chromotropic acid, 304

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (483–487) © 2019
Scrivener Publishing LLC

483
484 Index

Dendrimers, 89, 90 Melamine-formaldehyde resins


Differential Scanning Calorimetry Acetoguanamine, 108–110
(DSC), 86, 87 Chemistry, 102–105
Cold-setting, 111
Formaldehyde emission, 87, 88, 293–315 Manufacture, 124–129
Emission standards, 298–301 Methylolmelamines, 102–104
Exposure and safe levels, 297 MUF resins, 105, 106
Factors affecting it, 296, 297 PMUF resins, 107, 108
Test and refrence methods, 301–312 Preparation, 106, 107
2-ureido-4,6-diaminotriazine-1,3,5,
Grafting, 231, 381 106
Glycidyl metacrilate, 380
NMR, 82, 83, 255, 256
Hagen-Pouisseille law, 320
Hot-melt adhesives, 347–365 Peroxides, 376, 377, 393, 394
Amorphous polyolephins, 355 Peroxyacids, 416, 417
Antioxidants, 361–363 Phenol-formaldehyde resins
Diluents, 364 Acid vs Base catalysis, 116–123
Ethylene-Vinyl acetate copolymers, NaOH catalyzed, 116–123
347–349 Ammonia and amine catalyzed,
Plasticizers, 363 123–124
Polyamides, 352–354 Chelate ring intermediate, 123
Polyurethane reactive, 351, 352 Curing acceleration, 135–141
Silane reactive, 352 Dry resin films, 133–134
Styrenic block copolymers, 350, 351 History, 115, 116
Tackifiers, 356 Molecular Weight distribution, 132
Aliphatic hydrocarbon resins, Metal catalysis, 130
356, 357 Process variables
Aromatic hydrocarbon resins, 356 Catalyst and pH, 129–131
Mixed aliphatic and aromatic Viscosity, 131–132
resins, 357 Spray-dried types, 133
Rosin and rosin derivatives, Types, 116
359–361 Polymeric matrices
Terpene resins, 358 Biogradable polymer matrices,
Terpene-Phenol resins, 359 467–481
All Green Composites, 474–477
Isocyanates, 171, 172, 177, 381, 382 Polybutylene Adipate
Terephthalate, 474
Kissinger equation, 87 Polyhydroxyalkanoates, 467–468
Poly(3-hydroxybutyrate), 469,
Laminates, High pressure, 108–110 470
Poly(3-hydroxybutyrate-co-3-
MALDI ToF, 84, 85, 256–259 hydroxyvalerate), 470
Mark-Houwink equation, 319 Polylactic acid, 470–474
Index 485

Epoxy resins, 403–423 Unsaturated polyester, 389–402


Characteristics, 404 Bisphenol A-fumarate resins, 391
Bisphenol A Glycidyl Ethers, 405 Cross-linking, 393–401
Epoxy equivalent, 404 Curing at Elevated
Properties enhancement, 405 Temperatures, 393, 394
Resin types, 405 Peroxides, 393, 394
Epoxidized Natural Rubber, 417 Curing at room temperature,
Epoxidized novolacs, 414–415 394, 395
Epoxidized Vegetable Oils, 416, Cyclopentadiene based resins,
417 390, 391
Green or Sustainable Epoxy Diols, 389, 390
Matrix, 417–419 Isophthalic-Acid-Based Resin,
Preparation, 403, 404, 406 391
Curing agents, 406–414 Polyesterification reaction, 392
Tetrabromo Bisphenol A Epoxy Sheet molding compounds, 395,
Resins, 415, 416 396
Polyethylene, 425–440 Unsaturated polymeric matrix
Properties biocomposites, 397–400
Biocomposites, 430 UV Curable UP/Vinyl Ester
Flax-Fiber-Based, 434 Resins, 396
Hemp-Fiber-Based, 435 Vinyl ester, 391, 392
Kenaf-Based, 430–432 Polymeric matrices for natural fibres,
Sisal-Fiber-Based, 432–434 369–388
Melt Flow Index, 430 Compatibilizing biofibers
Production treatments
Gas Phase Fluidized Bed Physical methods, 370–371
Reactor, 429 Steam explosion, 371
High pressure process, 426 Chemical methods, 371
Low pressure process; Ziegler- Acetylation, 373
Natta, 426, 427 Acrylation, 374, 375
Slurry Process, 428, 429 Additives, 378
Solution Process, 428 Chlorotriazines, 377, 378
Polypropylene, 441–466 Isocyanates, 375, 376
Biocomposites Maleated coupling agents
Flax-Fiber-Based, 454 (MAH), 377
Hemp-Based, 456–460 Mercerization, 372, 373
Kenaf-Based, 445–448 Permanganate, 377
Oil-Palm-Fiber-Based, Peroxide, 376, 377
448–454 Silane coupling agents, 373, 374
Sisal-Based, 455, 456 Compatibilizing by matrices
Manufacture, 441, 442 functionalization, 378–382
α- and β-Forms of PP, 442, 443 MAH grafted polyolefins as matrix
Catalysts, 442 additives, 382
Polymerization methods, 443–445 Reactive extrusion systems, 383
486 Index

Polyurethane adhesives Surface migration, 211, 212, 218


Biobased PU adhesives, 185–192 Reduction methods
Blocked isocyanates, 177 Mechanical, 218, 219
Catalysts, 176–177 Chemical, 219, 220
EMDI adhesives, 183, 184 Thermal, 200–210
EPI adhesives, 184, 185 Causes, 202, 203
History, 170 Factors affecting it
Hybrid, 182, 183 Wood species, 203, 204
Isocyanate reactions, 171, 172 Drying, high, temp, 204–206
Interaction with wood, 178–181 Mechanisms, 203
Mechanism and kinetics, 178–181 Of inactivation, 206–207
Non-isocyanate polyurethanes Physical, 206–207
(NIPU), 185–192 Chemical
Polyurethane reactive hot melts, Sources, 202, 203
351, 352 Wood stabilization, 200
Surface modification
Resins without formaldehyde, 89–92 Methods
Resorcinol Resins, 147–151 Chemical, 231–234
Cold-set adhesive without Sol-gel, 233, 234
resorcinol, 163, 164 Corona, 227, 228
Reduced Resorcinol content Enzymatic, 230, 231
Branched, 161–163 Plasma, 224–230
“Honeymoon” fast setting
laminating adhesives, 159–161 Tannins
Reduced Resorcinol content, 159 Adhesives, 268–275
Hydroxymethyl resorcinol, 151–158 Analysis methods, instrumental,
Applications, 157–159 256, 257
Mechanism, 156–157 Chemistry, 239–244
PF without resorcinol Reactions
Preparation, 156 Autocondensation, catalytic,
Reactions with formaldehyde 248–250, 273–275
Quinone methides, 150–151 Condensation, acid, alkaline,
Reactive sites, 151 244–246
Reduced Resorcinol content Hydolysis, 244–246
Branched, 161–163 New reactions, 253–256
“Honeymoon” fast setting Rearrengement, 247
laminating adhesives, 159–161 Reactivity and orientation, 251,
Reduced Resorcinol content, 252
159– Sulphitation, 246
Silane coupling agents, 231–234 Metals complexation, 250
Surface inactivation Colloidal behavior, 252, 253
By Extractives, 211– Structure, Tridimensional, 250,
Influence on bonding, 212–217 251
pH effect, 217, 218 Thermomechanical analysis, 339–342
Index 487

Urea–formaldehyde adhesives Reactions, 62–64


Characterization, analysis, 82–87 In storage, 72
Commercial resins, 70–72 Parameters, 74
Curing, 75–78 Resins without formaldehyde, 89–92
Ammonium salts, 75–76 Second urea, 69
By ionic liquids, 79–81 Urons, 72
Hardeners, 78 Water resistance improvement, 81
Triazinone, 77, 78
Cyclic prepolymers, 89 Wood
Dendrimers addition, 89, 90 Cells, 6
Formaldehyde Cell walls organisation, 8–10
Content, 83 Chemical composition, 11–18
Emission, 87, 88 Extractives, 5
Without formaldehyde, 89–92 Fibrils, 9
History, 62 Hardwoods, structure, 6, 7
Kinetics, of curing, 87 Hierarchy, structural, 4
Low formaldehyde emission, 74 Influence on adhesion, 18–24
Manufacturing processes, 65–67 Influence of resin type, 25, 26
Methylol ureas, 65, 66 Processing parameters, effects, 26, 27
pH effect, 68 Resin penetration, 24–26
Polyamines modification, 88 Softwoods, structure, 7
Also of Interest

Check out these published and forthcoming related titles


from Scrivener Publishing
Reviews of Adhesion and Adhesives
(including open access research articles)
Editor: K.L. Mittal and Anil Netravali
Quarterly publication. ISSN 2168-0965
www.scrivenerpublishng.com

Progress in Adhesion and Adhesives, Volume 4


Edited by K.L. Mittal
Published 2019. ISBN 978-1-119-62525-4

Advances in Contact Angle, Wettability and Adhesion Volume 3


Edited by K.L. Mittal
Published 2018. ISBN 978-1-119-45994-1

Laser Technology: Applications in Adhesion and Related Areas


K.L. Mittal and Wei-Sheng Lei
Published 2018. ISBN 978-1-119-18493-5

Progress in Adhesion and Adhesives, Volume 3


Edited by K.L. Mittal
Published 2018. ISBN 978-1-119-52629-2

Textile Finishing: Recent Developments and Future Trends


Edited by K.L. Mittal and Thomas Bahners
Published 2017. ISBN 978-1-119-42676-9

Progress in Adhesion and Adhesives, Volume 2


Edited by K.L. Mittal
Published 2017. ISBN 978-1-119-40638-9
R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (489–491) © 2019
Scrivener Publishing LLC
Adhesion in Pharmaceutical, Biomedical and Dental Fields
Edited by K.L. Mittal and F.M. Etzler
Published 2017. ISBN: 978-1-119-32350-1

Interface/Interphase in Polymer Nanocomposites


Edited by Anil Netravali and K.L. Mittal
Published. 2016. ISBN 978-1-119-18491-1

Progress in Adhesion and Adhesives


Edited by K.L. Mittal
Published 2015. ISBN 978-1-119-16219-3

Advances in Contact Angle, Wettability and Adhesion Volume 2


Edited by K.L. Mittal
Published 2015. ISBN 978-1-119-11698-1

Particle Adhesion and Removal


Edited by K.L. Mittal and Ravi Jaiswal
Published 2014. ISBN 978-1118-83153-3

Laser Surface Modification and Adhesion


Edited by K.L. Mittal and Thomas Bahners
Published 2014. ISBN 978-1-118-83163-2

Adhesion in Microelectronics
Edited by K.L. Mittal and Tanweer Ahsan
Published 2014. ISBN 978-1-118-83133-5

Advances in Contact Angle, Wettability and Adhesion Volume 1


Edited by K.L. Mittal
Published 2013. ISBN 978-1-118-47292-7

Advances in Modeling and Design of Adhesively Bonded Systems


Edited by S. Kumar and K.L. Mittal
Published 2013. ISBN 978-1-118-68637-9

Atmospheric Pressure Plasma Treatment of Polymers


Edited by Michael Thomas and K.L. Mittal
Published 2013. ISBN 978-1-118-59621-0
Atomic Layer Deposition
Principles, Characteristics, and Nanotechnology Applications
By Tommi Kääriäinen, David Cameron, Marja-Leena Kääriäinen and
Arthur Sherman
Published 2013. ISBN 978-1-118-06277-7

Encapsulation Nanotechnologies
Edited by Vikas Mittal
Published 2013. ISBN 978-1-118-34455-2

Atmospheric Pressure Plasma for Surface Modification


By Rory A. Wolf
Published 2012. ISBN 9781118016237

Introduction to Surface Engineering and Functionally Engineered


Materials
By Peter Martin
Published 2011. ISBN 978-0-470-63927-6

You might also like