treatise on adhesion and adhesivesdeteix/bois/documents_references/river_wood... · of ancient...

238
WOOD AS AN ADHEREND chapter 1 from volume 7 by Bryan H. River Charles B. Vick Robert H. Gillespie Treatise on ADHESION and ADHESIVES edited by Marcel Dekker, Inc. J. Dean Minford New York 1991

Upload: others

Post on 30-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • WOOD AS AN ADHEREND

    chapter 1fromvolume 7

    byBryan H. RiverCharles B. VickRobert H. Gillespie

    Treatise onADHESION and ADHESIVES

    edited by Marcel Dekker, Inc.J. Dean Minford New York 1991

  • TABLE OF CONTENTS CHAPTER I CHAPTER II

    CHAPTER III

    CHAPTER IV

    CHAPTER V

    CHAPTER VI

    INTRODUCTION WOOD CHARACTERISTICS INFLUENCING THE BONDING PROCESS AND BOND QUALITY

    Physical Structure Chemical Composition Physical Properties Mechanical Properties Thermal and Dielectric Properties Wood Supply

    Surface Preparation Machining Processes Chemistry of Wood and Fiber Surfaces Roughness Interactions Between Liquids and Wood Surfaces Revitalization and Modification of Wood and Fiber

    Surfaces WOOD ELEMENTS AND BONDED WOOD PRODUCTS

    WOOD AND FIBER SURFACES

    Unique Processiblity of Wood Usefulness of Wood Anisotropy Table of Wood Elements Combinations of Wood Elements and Other

    Materials and Products FUNDAMENTALS OF WOOD BONDING

    Mechanisms of Adhesion Setting of Adhesives Bonding Process

    Performance Criteria Strength and Related Criteria Stability and Related Criteria Appearance Performance Evaluation

    PERFORMANCE OF BONDED JOINTS AND MATERIALS

    Measures for Improving Bond Performance REFERENCES APPENDIX Summary of Representative Durability Specifications and

    Quality Assurance Standards from Various Countries

    Page 1 6

    6 13 19 28 50 53 54 55 55 75 85 87 98

    102 1 02 103 103

    107 112 112 114 115 134 1 34 134 162 178 180 188 190

    21 1

  • Figure 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58

    List of Figures A longitudinal radial view of southern pine Gross structure of a typical softwood Gross structure of a typical hardwood Schematic construction of the cell wall Chemical components of wood Relationship of the moisture content of wood at equilibrium with the humidity and temperature of air Pseudo-orthotropic structure of wood in relation to the tree and grain direction Adhesive bond areas of different types of joints Shear and tension modes of loading wood in relation to its orthotropic structure Temporary effect of wood moisture content on shear strength Temporary effect of temperature on the compression strength of wood parallel to the grain Effect of grain angle on mechanical properties of wood according to the Hankinson formula Schematic diagrams of typical loci of fracture in wood Scanning electron micrographs of typical transwall fractures in wood Scanning electron micrographs of typical lntrawall fracture in wood Crack-stopping mechanism in wood stressed parallel to the grain or in bending Schematic diagram of the maximum potential bond strength and factors that reduce maximum effective bond Dependence of maximum swelling pressure under uniaxial restraint upon the density of wood Computed thermal conductivity of wood perpendicular to the grain determined by moisture content End-grain view of earlywood cells just beneath the surface of wood after surfacing Angles of a cutting tool Codes for direction of travel of a cutting edge with respect to the orthogonal directions of wood Actions of cutting tools in forming various types of chips in orthogonal cutting of wood Effect of grain angle upon the usrface quality resulting from Type I chip formation Tensile zones in veneer cutting that cause lathe checks and rough surfaces as a result of tearout Flexure of veneer and formation of a lathe check as a result of tension zones in veneer cutting Influence of knife marks per inch on wood surface quality (smoothness) Effect of jointing the cutting edge upon the shape of the cutting edge and compression of the wood Downward forces created by sandpaper grit in abrasive planing Characteristic machined wood surfaces for bonding Roughness of wood surfaces due to locus of cut Compressed and smeared surface formed by planing with a dull knife excessively jointed Relationship between the shear strength of bonded joints and the porosity or roughness of wood Contact angle of a liquid with a solid surface formed during wetting Relationship between wood wettability and specific gravity Model for the work of wetting as functions of surface roughness and surface tension Improvement in wettability by light abrasion by water drop test Several types of wood elements used to manufacture reconstituted panel and lumber materials Average wood moisture content for interior use in the United States Various types of warp Relative humidity control required to maintain equilibrium moisture content of wood and products Bondlines between sound and crushed wood surfaces Schematic plot of strength and percentage wood failure as functions of wood specific gravity Relative strength of bonded shear test specimens as a function of percentage of wood failure Examples of conditions leading to differential shrinkage, warping and delamination stresses Predicted stresses in bondlines as a function of angle between growth rings of two lamina Effects of swelling and shrinking on mitered butt joints Common adhesively bonded end joints used with wood members. Effect of tip thickness on the tensile strength of finger joints Durable, nondurable, permanent, and nonpermanent behavior of adhesive compared to wood Recoverable and nonrecoverable effects of internal stress on adhesively bonded joints Effect of moisture content on wood and joints bonded with various adhesives Comparison of the permanence of common wood adhesives in plywood specimens outdoors Resistance of adhesive bondlines in oak to shrinkage stresses Five-layer panel showing symmetry and balanced construction Method of alternating growth ring orientation in narrow boards to minimize warping of panels Plywood panel durability as a function of outdoor exposure time Standard adhesive joint strength test specimens

    Page 7 9

    11 12 13 21 23 29 30 33 35 39 41 43 45 46 47 48 52 59 60 62 63 64 66 67 69 70 71 76 86 87 88 89 93 97

    100 106 116 121 122 138 141 142 148 150 153 1 54 158 163 1 66 169 171 1 74 1 76 177 182 185

  • Table 1.1Table 1.2 Table 1.3 Table 1.4 Table 1.5 Table 1.6 Table 1.7 Table 1.8 Table 1.9 Table 1.10 Table 1.11 Table 1.12 Table 1.13 Table 1.14 Table 1.15 Table 1.16 Table 1.17 Table 1.18

    Table 1.19 Table 1.20 Table 1.21

    LIST OF TABLES

    Wood Bonding Variables Major Groups of Extractable Materials Tangential and Radial Shrinkage from Fiber Saturation to Oven-Dry Condition for Selected Species Strength of Selected Woods at 12% Moisture Content under Various Types of Loading Representative Mechanical Properties of Wood in the Three Principal Directions Differences in Strength Between the Tangential-Longitudinal and Radial-Longitudinal Planes Thermal Conductivity of Wood and Other Materials Accessiblity and Contribution of Cell Wall Constituents to Hydrogen Bonding Specific Strength and Stiffness of Wood and Other Common Materials Comparative Processing Costs of Construction Materials Common Wood Elements in a Series of Diminishing Dimensions Single Element Products and Their Primary Properties Products Based on Two or More Elements Recommended Moisture Content Values for Various Wood Applications at Time of Installation Typical Size Changes in a 3-in.-Wide Piece of Wood Temperature Elevation of a Storage Area to Maintain Equilibrium Moisture Content (EMC) Drying (Delamination) Stresses Calculated from a Model of Elastic Wood Behavior Linear Expansion and Contraction of Selected Woods and Wood-Based Materials Between 30 and 90% Relative Humidity Moisture Excluding Effectiveness (MEE) of Selected Wood Coatings Durability Specifications for Wood Adhesives Quality Assurance Standards for Wood Joints and Wood-Based Materials

    Page 2

    17 26 31 37 38 50 77

    102 104 105 108 111 118 120 123 145 147

    189 212 220

  • 1

    Wood as an Adherend

    BRYAN H. RIVER and CHARLES B. VICK Forest ProductsLaboratory, USDA Forest Service, Madison, Wisconsin

    ROBERT H. GILLESPIE Consultant, Madison, Wisconsin

    I . INTRODUCTION

    Wood is a porous, permeable, hygroscopic, orthotropic, biologicalcomposite material of extreme chemical diversity and physical in-tricacy. Table 1.1 provides an overview of the may variables, in-cluding wood variables, that bear on the bonding and performanceof wood in wood joints and wood-based materials. Of particularnote is the fact that wood properties vary between species, be-tween trees within a species, and even within a tree. Variabilitywithin a single species alone is enough to significantly challengean adhesive to perform consistently and satisfactorily. In thischapter, we have attempted to describe wood and to explore howthis complex biological material interacts with adhesives to affect thebonding process and the quality of the bonded joint or material.First, we will present a short review of the history connectingwood and adhesives.

    The gluing of wood is ancient. No one knows where or whenit began. We do know that early civilizations used mud, plant

    The Forest Products Laboratory is maintained in cooperation with theUniversity of Wisconsin. This article was written and prepared byU.S. Government employees on official time, and it is therefore inthe public domain and not subject to copyright.

  • 2

    Table 1.1 Wood Bonding Variables

    Resin Wood Process

    River et al.

    Service

    TypeViscosity

    Tack

    Mole ratio ofreactants

    Filler

    Total solids

    Molecular weightdistribution

    Solvent system

    Age

    pHBuffering

    Cure rate

    Catalyst

    Mixing

    Species

    Species gravity

    Moisture con-tent

    Plane of cut:radial, tan-gential, trans-verse, mix

    Heartwood ver-sus sapwood

    Juvenile versusmature wood

    Earlywood ver-sus reactionwood

    Grain angle

    Porosity

    Surface rough-ness

    Drying damage

    Machining dam-age

    Dirt

    Contaminants

    Chemical sur-face

    Extractives

    pH

    Buffering ca-pacity

    Errata (April 2004): This table was originally developed by Norm Kutscha for the Forest Products Research Society Gluing Technical Committee.

    Adhesive spread

    Adhesive dis-tribution

    Relative humid-ity

    Temperature

    Open assemblytime

    Closed assem-bly time

    Pressure

    Gas-through

    Press time

    Pretreatments

    Posttreatments

    Adherend tem-perature

    Strength

    Shear modulus

    Modulus ofelasticity

    Creep

    Percentage ofwood failure

    Failure type

    Adhesive penetration

    Dry versus wet

    Temperature

    Finishing

    Heat resistance

    Hydrolysis re-sistance

    Swell-shrinkresistance

    Ultraviolet re-sistance

    Biologicalresistance :fungi, bac-teria, insects,marine organ-isms

  • Wood as an Adherend 3

    resins, beeswax, bitumen, and other naturally occurring substancesfor glue. Pottery and weapons bonded with resin have been dis-covered in grave sites almost 6000 years old (Stumbo 1965). Evi-dence has been found that lime-based plaster was used for bondingstone blades as long as 12,000 B.C. (Bower 1988). The use of firefor cooking undoubtedly led to the discovery that plant and animalproteins make sticky materials. Many of these naturally occurringmaterials were available to ancient peoples, including lime, eggwhite, and flour paste. With these primitive materials, wood bond-ing developed into a sophisticated art form. A love of luxury anda delight in beautiful surroundings evolved in the ruling familiesof ancient Egypt. They found wood to be an ideal material fromwhich to fashion decorative furniture. But decorative woods werenot abundant in that land of little rain. Ebony, teak, and rose-wood were imported, no doubt at great expense, prompting crafts-men to learn to cut thin veneers to stretch supplies. Wall carvingsin Thebes, dated about 1500 B.C., show thin wood veneers beingglued to a plank of wood. A glue pot and brush, with the potwarming over a fire, suggest the glue was a hot animal glue (Knightand Wulpi 1927, referencing Wilkinson 1878). These veneers wereglued to a core of ordinary material, which added strength anddurability while conserving the supply of decorative but scarce andcostly woods. Plywood, made by gluing together thin sheets orplies, was also well known. The people made table tops, chests,beds, and other furniture by gluing thin veneers to suitable corematerial. The decoration of furniture with inlays of precious stones,gold, and ivory, as well as the veneers of rare woods, was devel-oped to a high artistic level in Egypt. We can appreciate the resultof the Egyptians’ gluing skill, for many beautifully veneered and in-laid wooden artifacts recovered from the tombs of the Pharaohs, andfrom other archeological excavations, survive to this day.

    Undoubtedly the first articles of furniture were fashioned in onepiece. Later, various parts were held together by leather thongsor wooden pegs. But the marriage of wooden parts with gluedstructural joints occurred at least as early as the eighth centuryB.C. An ornately decorated and intricately formed table recentlydiscovered in a tomb, possibly that of the renowned King Midas,had been assembled with dowels and mortise and tenons joints(Simpson 1983). Although the table had failed under a heavy loadof bronze pots, it demonstrates that artisans had discovered thestructural advantages of adhesive-bonded joints for furniture (Dar-row 1930).

    After the Egyptian period, veneering continued as an art formin Greece and later in Rome. The ancient Greeks described arecipe for casein glue not unlike the recipes of the early twentiethcentury. The art suffered a setback with the fall of Rome and the

  • 4 River et al.

    loss of interest in decorative arts. Only traces of the veneeringart from this period can be found today.

    Interest in gluing rekindled during the Renaissance. From thefifteenth to the middle of the nineteenth century, the art of ve-neering and development of intricate glued structural joints led tothe creation of magnificent furniture and architectural woodwork inItaly, France, England, Holland, and Flanders, and finally in theUnited States. Distinctive periods of styling were known by thenames of ruling sovereigns or patrons of the arts-de Medici,Elizabethan, Victorian, Queen Anne, and Louis XIV, XV, and XVI.Great master cabinetmakers rose to prominence one after another,and styles were distinguished by their names-Adam, Hepplewhite,Chippendale, and Sheraton. All the great masters worked withanimal glue and, to lesser extent, with fish and casein glues, tocreate the forms of their imaginations in wood. The lasting graceand beauty of their creations are, in no small measure, a tributeto the performance of these glues (Pollen).

    Other glues from natural sources were neither highly developednor widely used until after the beginning of the twentieth century.Then, F. G. Perkins developed a vegetable glue (actually tapiocastarch) from the roots of the cassava plant (Perkins 1912). Thisglue became so successful for cold pressing furniture parts, ply-wood, veneering, and other wood applications that the Perkins Com-pany sold 230 × 106 lb of it in 1930 (Darrow 1930). The use ofanimal blood as a glue goes back many centuries-Aztec Indiansmixed blood with mortar for building construction. Two develop-ments spurred its widespread use in modern times. One was thediscovery in about 1910 that blood could be dried in a solubleform, making it easier to preserve and handle. The other wasthe urgent need for a water-resistant glue for plywood for aircraftconstruction during World War I (Lambuth 1977). The war alsospurred the development of water-resistant casein glues for themanufacture of laminated wood aircraft propellers (Truax 1930).A water-resistant, soybean-protein glue was possibly the last ma-jor development in glues of natural origin (as compared to syn-thetic resin adhesives) (Johnson 1923, Laucks and Davidson 1928).It became the major glue for interior softwood plywood with produc-tion of 34 × 106 lb annually in the 1930s. Combinations of bloodalbumin and soybean protein took advantage of the best propertiesof each material (Lambuth 1977); these combinations were used ex-tensively for plywood until displaced by synthetic resins in the1950s.

    Until about 1930, the term glue accurately described the materi-als used to bond wood because all materials were derived from nat-urally occurring substances such as casein and collagen. Thesematerials, their processing and use, and their performance with

  • Wood as an Adherend 5

    different woods were carefully described by T. R. Truax (1929).At that time, furniture was the principal product manufactured bywood bonding. Softwood plywood was in its infancy, and it wassuitable for only interior use because of the poor water-resistanceof the vegetable-starch glues used in its manufacture.

    With the development of synthetic resins, the term adhesive be-came a more appropriate word for the broad range of bondingagents that included synthetic resins as well as glues. The firstimportant wood product made with the new synthetic resin adhe-sives was water-resistant plywood for airplane construction. Butthe impact of synthetic resin adhesives was not really felt until theearly 1930s, when urea- and phenol-formaldehyde-bonded plywoodbegan to be used in furniture and housing. World War II intensi-fied the demand for water-resistant or waterproof bonded-woodproducts. During this time, adhesives based on the synthetic ther-mosetting resins-urea-formaldehyde, melamine-formaldehyde, phenol-formaldehyde, and resorcinol-formaldehyde-began to replace adhe-sives from natural resources. The emergence of commercial syn-thetic resin adhesives greatly expanded the variety of useful woodproducts that could be manufactured. In the early 1950s, adhe-sives based on thermoplastic vinyl acetate resin began to replaceanimal glue in furniture assembly. Two-polymer adhesives com-bining thermosetting and thermoplastic resins, such as polyvinyl-acetal/phenol-formaldehyde and nitrile rubber/phenol-formaldehyderesins, were developed that had the capability of bonding wood tometal. By the 1950s, the variety and number of synthetic resinadhesives and bonded-wood products mushroomed.

    Practically all branches of the wood-using industry now use ad-hesives (White 1979). Adhesives and the industries they havespawned are responsible for new or improved products, dramaticimprovements in the utilization of forest and mill residues, and con-servation of timber supplies. New industries such as plywood, par-ticleboard, flakeboard, and laminated-veneer lumber owe their veryexistence to synthetic resin adhesives. In the United States alone,these industries, which are entirely dependent on adhesives, an-nually produce :

    21 × 109 ft2 of structural panels (3/8-in.-thick basis).10 × 109 ft2 of nonstructural panels (3/8-in.-thick basis), excluding

    hardwood plywood.104 × 106 ft2 laminated-veneer lumber (1-1/2-in.-thick basis).

    The United States also imports 3.1 × 109 ft2 (surface measure) ofhardwood and softwood plywood.

    The development of these and other bonded-wood products andthe growth of related industries over the years have resulted in

  • 6 River et al.

    the consumption of huge amounts of synthetic resins. In 1986, thewood products industries consumed 1.4 kt of melamine-formaldehyde,1.4 kt of isocyanate, 1.1 kt of poly(vinyl acetate), 3.6 kt of resor-cinol-formaldehyde and phenol-resorcinol formaldehyde, 793 kt ofphenol-formaldehyde, 745 kt of urea-formaldehyde, and 12 kt ofmastic construction adhesive (Myers 1988). The total consumptionof adhesives used in the forest products and construction indus-tries was 2.37 billion pounds in 1988, and it is estimated that con-sumption will total more than 3 billion pounds by the year 2000(Anonymous 1989).

    I I . WOOD CHARACTERISTICS INFLUENCING THEBONDING PROCESS AND BOND QUALITY

    A. Physical Structure

    Wood

    There are two major types of wood: softwood (from needle-bearingtrees) and hardwood (from broad-leaved trees). The terms softwoodand hardwood are misnomers and have little to do with hardness ordensity. Many softwoods are actually much harder than many hard-woods.

    When a tree is cut, often the most noticeable features on the endof the log are the light outer ring and the dark inner core. Theouter ring, called the sapwood, consists of the cells that were ac-tively growing or carrying on the life processes of the tree when itwas cut. As the tree grew larger in diameter, cells closer to thecenter of the tree, which were no longer required for these activi-ties, died and were converted to heartwood. This conversion entailsboth anatomical and chemical changes. The anatomical changes haveminimal effect on strength, but they may reduce permeability, thusaffecting bonding behavior. Chemically, stored food disappears andnew chemicals are created. When these chemicals oxidize, the heart-wood darkens, and the border between sapwood and heartwood be-comes more evident. Heartwood formation is highly individualisticbetween species, even within trees of the same species. Thesechanges, especially the chemical changes, account for much of thedifficulty and unpredictability in the bonding of heartwood as weshall explain later.

    All wood is made of fibrous cells organized to perform the supportand living functions of the plant. The cells are organized in annualgrowth increments or growth rings. Each ring is the result of 1year of growth. The rings are usually prominent because of cycli-cal variation in color or porosity. These variations in turn are dueto the formation of different types of cells and wood structure during

  • Wood as an Adherend 7

    different parts of the growing season. Lighter colored (less dense)and more porous cell tissue, called the earlywood, forms early in thegrowing season. The porous earlywood cells are largely responsiblefor the movement of liquid and nutrients about the tree. Darker(more dense) and less porous cell tissue, called the latewood, formslater in the growing season. The latewood cells are largely respon-sible for supporting the tree (Figure 1.1).

    The growth patterns of some species result in a large contrastbetween the earlywood and latewood densities. These woods are

    Figure 1.1 A longitudinal radial view of southern pine showing thedark, dense latewood and the light, less dense earlywood.

  • 8 River et al.

    called coarse textured. Southern pines are an example of coarse-textured softwoods. Coarse-textured hardwoods, such as red oak,have a special name; they are called ring-porous woods. In otherspecies, the growth pattern is less variable in density and poros-ity, and the wood is called uniform textured. White pine is a uni-form-textured softwood. Uniform-textured hardwoods, such asbasswood, also have a special name; they are called diffuse-porouswoods.

    Large differences between the earlywood and latewood porosityand density in some species like oak and southern pine often causedifficulty in bonding, as we will later discuss in detail.

    Cells

    Wood cells are microscopic, long, thin, hollow tubes, like sodastraws with their ends pinched shut. The long axis of the majorityof the cells is parallel to the longitudinal axis or grain direction ofthe tree trunk. Most longitudinal cells are either for support orfor the movement of fluids in the living tree. However, small num-bers of special cells either produce or store nutrients and chemicals.Some special cells are organized into tissue called rays that lie per-pendicular to the longitudinal axis of the tree trunk and along itsradii. The very broad rays of oak are plainly visible to the nakedeye on every surface. Ray cells are responsible for the productionand storage of amorphous materials of complex chemical nature.The rays are also the pathway for lateral movement of fluids inthe tree.

    There are two basic types of cells-prosenchyma and parenchyma.Softwoods and hardwoods have different types of prosenchyma andparenchyma cells. Prosenchyma cells are generally the strong woodycells responsible for mechanical support and the movement of fluidsin the living tree. Parenchyma cells are responsible for the pro-duction of chemicals and for the movement and storage of food.The real differences between softwoods and hardwoods are in thesize, shape, and diversity of these two types of cells.

    The structure of softwoods is characterized by relatively fewtypes of prosenchyma and parenchyma cells compared to hardwoods-a result of their lower position on the evolutionary scale (Figure1.2). 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 thetree. Earlywood tracheids are generally of large diameter and thinwalled. Earlywood cells are specifically adapted to moving fluidsthrough large openings (bordered pits) that connect adjoining Cells.Latewood tracheids, which are generally smaller in diameter, arethicker walled, have smaller pits, and are specifically adapted for

  • Figure 1.2 Gross structure of a typical softwood, showing relativelyfew types of cells and the preponderance of longitudinally (vertically)oriented cells called tracheids.

  • 10 River et al.

    strength. The remaining 10% of a softwood consists of longitudinalparenchyma cells, ray tracheids, and ray parenchyma cells. Gen-erally, parenchyma cells play a secondary strength role, but theyare important for adhesive bonding as paths for adhesive penetra-tion. Moreover, the chemicals contained by the cells affect adhe-sion and adhesive cure.

    In comparison to softwoods, the structure of hardwoods is char-acterized by a greater diversity of cell types and functions (Figure1.3). One notable difference is that specialized prosenchyma cellsare responsible for mechanical support, and other specialized pros-enchyma cells are responsible for fluid movement. Support is pro-vided by two types of small-diameter thick-walled prosenchyma cellscalled libriform fibers and fiber tracheids. Fluid movement is pro-vided by medium- to large-diameter, thin-walled, and open-endedcells called vessel elements. Normally, a number of vessel elementslink end-to-end along the grain to form long tube-like structuresknown as vessels. The cavities of large vessels in oak and otherspecies are large enough to see with the naked eye. Such large cav-ities obviously affect wood strength and adhesive flow when pressureis applied during bonding. Together the longitudinally oriented fi-bers and vessels constitute the major volume of cells (roughly 70-90%) in hardwoods. A number of other specialized longitudinal pros-enchyma and parenchyma cells and ray prosenchyma and parenchy-ma cells constitute the remaining volume. As in the softwoods, someof these minor hardwood cell types have important chemical roles andsecondary, though often minor, mechanical roles. Panshin anddeZeeuw (1980) provide further information on wood anatomy.

    Cell Wall

    Under the microscope, the end of a piece of wood looks rather likea honeycomb (Figures 1.2 and 1.3). The walls of the honeycomb,the wood cell walls, are a framework of oriented long-chain cellu-lose molecules called elementary fibrils. These are grouped inbundles called microfibrils. In certain regions of the microfibrils,the elementary fibrils are highly aligned, tightly packed, and crys-talline. In other regions they are less aligned, not packed, andnoncrystalline. The cellulose chains in the noncrystalline regions ofthe microfibrils are interpenetrated and thereby stiffened by a het-erogeneous and amorphous matrix of lignin that also binds the cellstogether. Spaces between the microfibrils are thought to be filledwith a heterogeneous matrix of short-chain cellulose-like materialscalled hemicellulose, as well as lignin and other amorphous materi-als, air, and water. The exact relationships between cellulose, hemi-cellulose, and lignin in these regions are not entirely understood.Water is able to enter and leave these noncrystalline regions quite

  • Figure 1.3 Gross structure of a typical hardwood, showing thegreater variety of cell types when compared to the softwood in Fig-ure 1.2. Also evident are the large-diameter vessels (verticallyoriented) and the larger amount of horizontally oriented ray cellscompared to the softwood.

    11

  • 12 River et al.

    freely; this accounts for the swelling and shrinking of wood in re-sponse to changes in relative humidity.

    Successive lamellae of microfibrils are laid down in waves fromthe middle to the end of the cell, each lamella with a slightly dif-ferent orientation of the microfibrils. These lamellae surround. thecentral cavity or lumen of the cell. Distinct groups of lamellae orwall layers are distinguished by differences in the orientation ofthe microfibrils. The first layer, the primary wall, is very thinand consists of randomly oriented microfibrils. The primary wallprovides the framework for the subsequent formation of the sec-ondary wall. The secondary wall is formed of three distinct layers,the S-1, S-2, and S-3 layers; each layer is much thicker than theprimary wall (Figure 1.4). The secondary wall is the principal

    Figure 1.4 Schematic construction of the cell wall showing themiddle lamella (ML) and primary wall (P), the S1 layer of the sec-ondary wall, the dominant S2 of the secondary wall, the S3 layer,and the warty layer (W).

  • Wood as an Adherend 13

    structural element of the wood cell. In the secondary wall, the mi-crofibrils are aligned helically around the lumen. The angle anddirection of the helices vary from layer to layer within the wall,and the thickness of the individual layers and the overall thicknessof the cell wall vary with the type of cell. The properties of thecell are strongly influenced by the degree of orientation of themicrofibrils and the proportions of the various layers in the cellwall. For example, the S-2 layer, whose microfibrils are orientednearly parallel to the long axis of the cell, is responsible for re-sisting principal stresses in the living tree and for the longitudinalstrength and stiffness of lumber cut from the tree. Latewood cellsnormally have very thick S-2 layers and thus are very resistant tostresses parallel to the long axis of the cell. Cell wall thicknessvariation and swelling and shrinking are principally, but not exclu-sively, due to thickness variation of the S-2 layer. The S-1 layeris important to resisting stresses perpendicular to the grain direc-tion; the S-1 and S-3 layers, whose microfibrils lie at a large angleto the long axis, also restrain the swelling and shrinking of theS-2 layer, and thus of the wood as a whole. The secondary wallby virtue of its bulk largely determines the mechanical and physi-cal properties of the wood.

    B. Chemical Composition

    Wood is made up of cell wall constituents and extraneous materials(Figure 1.5). The cell wall constituents that form the structural

    Figure 1.5 Chemical components of wood.

  • 14 River et al.

    components of the wood cell wall are collectively called wood sub-stance. Wood substance typically accounts for 95-98% of theweight of the wood, the remainder being extraneous organic andinorganic materials. The ratio of polysaccharides to lignin in thewood substance is roughly 3:1. The most abundant polysaccharide,cellulose, in the form of microfibrils, provides the framework forall plant tissues. Pettersen (1984) provides an extensive compila-tion of the chemical composition of woods from the entire world.

    Cell Wall Constituents

    Cellulose is an unbranched and highly oriented homopolymer formedof ß-D-glucose units linked by ß-1,4-glycosidic bonds to form longlinear macromolecules. In nature, the cellulose polymer consists of5000 to 10,000 repeating units and may consist of as many as 30,000repeating glucose units. The physiochemical relationship betweenadjacent cellulose chains is not completely understood. X-ray dif-fraction evidence indicates large portions of cellulose exist as well-ordered parallel arrays of molecules held together by intermolecularhydrogen bonding. This portion of the cellulose is highly crystal-ine, whereas other parts are not well ordered and are amorphous.Each glucose unit has three hydroxyl units that are available forhydrogen bonding. Hydroxyl units in the amorphous regions areresponsible for the great attraction of wood for water, and theyprovide the primary sites for adhesive bonding.

    Cellulose is the principal structural component of wood. It con-stitutes roughly 42% of the wood in both softwoods and hardwoods.Anatomically, cellulose is most abundant in the S-2 layer and leastabundant in the middle lamella. Pure wood cellulose is strong, verystiff, and fibrous, but it is unable to function alone in supportingthe tree because the microfibrils in the noncrystalline regions buckleeasily under compression because of their small diameter. The mi-crofibrils are intimately associated with lignin and hemicellulose, whichbond and support the cellulose microfibrils (Winnandy and Rowell 1984).

    Lignin interpenetrates and rigidifies the cellulose microfibrillarframework, making certain plant tissues woody and thus able toresist compression forces. Lignin is a phenolic, highly branchedthree-dimensional heteropolymer formed by enzymatic polymerizationof three elementary monomers-coumaryl, coniferyl, and sinapyl al-cohols. These alcohols are linked by ether and carbon-carbonbonds to form structural units (Pettersen 1984). The structuralcomponents in the lignin polymer are referred to as guaiacyl,syringyl, and p-hydroxyphenyl units, respectively, from the threemonomeric alcohols. Softwood lignins are distinguishable by thepredominance of guaiacyl units in their composition, whereas hard-wood lignins contain both syringyl and guaiacyl units. The ratio

  • Wood as an Adhered 15

    of guaiacyl and syringyl units is an important measure of lignin char-acterization. The ratio varies with the type of cell and the locationwithin the cell wall (Fergus and Goring 1970).

    Lignin also has hydroxyl units available for the adsorption of wa-ter and adhesive bonding, although there are few compared to cellu-lose. In its native state, lignin is thermoplastic and softens at about100°C. The temperature at which it softens, however, is stronglyaffected by moisture. The ability of the binding agent lignin toflow under heat and pressure accounts for many of the unique proc-essing characteristics of wood, such as thermomechanical pulping,steam bending, and bonding of certain types of reconstituted panelproducts such as masonite.

    Lignin constitutes 24-33% of the wood substance in softwoods and16-24% in hardwoods. Anatomically, lignin content is most dense inthe middle lamella, where it constitutes 60-90% of the wood sub-stance. However, because of the great thickness of the secondarywall, most lignin is actually located in the secondary wall betweenthe cellulose microfibrils. Lignin contributes to compression strengthas a rigidifying and bulking agent. It also contributes to tensileand shear strength indirectly by protecting hydrophilic polysac-charidic materials that act as bonding agents between the cellulosemicrofibrils and between adjacent cells. Chemical removal of ligningreatly increases dry tensile strength but also greatly lowers wetstrength of wood and fibers (Klauditz 1952). Lignin is consideredto provide some measure of protection to the hydrophilic polysac-charides of wood substance from water.

    The last major component of wood substance is a heterogeneousgroup of polysaccharides known as hemicellulose. Hemicellulosescomplement the lignin fraction, so they constitute about 20-29% ofthe cell wall substance of softwoods and 29-37% of hardwoods. Col-lectively, the hemicelluloses are hydrophilic, thermoplastic, alkali-soluble, and heat-labile polysaccharides. Their function is less wellunderstood than that of lignin. Unlike cellulose, which consists ofone basic repeating unit, hemicelluloses are comprised of five dif-ferent sugar monomers (glucose, mannose, galactose, xylose, andarabinose). For example, glucomannans are formed by polymeriza-tion of glucose and mannose. Xylans are formed by polymerizationof xylose. The degrees of polymerization of the hemicellulose mole-cules are tens or hundreds of repeating units, instead of thousandsof units as in cellulose. Branching may occur, as well as the addi-tion of acetyl ester and uronic acid ester groups. Glucomannansare the predominant hemicellulose found in softwoods; xylans pre-dominate in hardwoods. Both softwoods and hardwoods also containsmall amounts of water-soluble pectic substances such as uronans,galactan, and arabinan.

  • 16 River et al.

    Glucomannans have short branches and lie parallel to and in closeassociation with the cellulose chains in the micrfibrils. Xylans arewell distributed throughout the cell wall and appear to be locatedwithin the interstices and upon the surfaces of the cellulose micro-fibrils. They are well branched and apparently form a complex in-terpenetrating matrix material with lignin (Kerr and Goring 1975;Fengel and Wegener 1984; Bach-Tuyet, Hyama, and Nakano 1985).Pectic substances are found mainly in the middle lamella and pri-mary wall. Pectic substances are thought to provide bonding be-tween adjacent cells and to control the properties of the primarywall.

    The thermal softening temperature of hemicellulose (about 60°C)is much lower than that of cellulose or lignin, and it is lowered fur-ther by the presence of water (Goring 1965). However, in thepresence of lignin, hemicellulose flow is inhibited until the ligninsoftening temperature (about 100°C) is reached (Byrd 1979). Hemi-celluloses are believed to act as bonding agents in paper formationfrom high-yield pulps, which contain large amounts of hemicelluloseand lignin (Horn 1979). In a process called press drying, a wetpulp sheet is dried under heat and pressure to form paper with un-usually high wet strength. The unusual properties of the press-dried paper are attributed to flow of both the hemicellulose and thelignin (Horn 1979, Byrd 1979). Hemicellulose flow is thought to beresponsible for superior adhesion between fibers in press-driedpaper. However, lignin is thought to flow and surround the hemi-cellulose bonds, protecting them from water (Horn 1979). Hemicellu-lose and lignan flow may also play important roles in solid-woodbonding (Young and others 1985). The relationships betweencellulose, lignin, and hemicelluloses are discussed in depth byMark (1967).

    Wood is mildly acidic. The acidity arises from acetyl groups at-tached to the xylan but also from the absorption of cations of ex-traneous mineral substances (ash) and from the organic extractives(Rowell 1982a). Gray (1958) measured the pH of damp sawdust of109 hardwood and 20 softwood species. Most species fell in the pHrange of 3.0-6.0. Only one species, Parana pine, was alkaline,with a pH of 8.8.

    Extraneous Materials

    Extraneous materials are organic or mineral substances found inthe cell wall and cell lumen. These materials usually account forup to 5% of the dry weight of unextracted wood. However, in somespecies or in certain locations within the tree, they may constituteas much as or more than 30% of the weight of the wood. The or-ganic substances are a mixture of compounds with diverse chemical

  • Wood as an Adherend 17

    properties. They are called extractives because they can be re-moved (extracted) from wood by fairly gentle procedures, such asbathing with hot water, alcohol, benzene, or ether. The extrac-tion procedure is a convenient way to group these materials, asoutlined in Table 1.2. Each group includes several classes of ma-terials; each class may include many different compounds, and somematerials overlap in terms of their solubility. More can be learnedof specific extractives in various textbooks on the subject (Hillis1962, Pettersen 1984, Fengel and Wegener 1984).

    Many unique properties of various species of wood are due totheir different extractives. Even in very small quantities, someextractives impart strong resistance to decay and insects. Thechemical uses of various species of wood, as in pulp and papermaking, syrup production, and naval stores, are all totally basedon, or strongly influenced by, extractive materials. Extractivesare almost totally responsible for the color, odor, or smell of agiven species of wood. The abrasiveness of inorganic extraneousmaterials, such as silica, dulls cutting tools and adversely affectsthe machineability of the wood, even though the extraneous materialis present in small quantities.

    Table 1.2 Major Groups of Extractable Materials

    Group by methodof extraction

    Steam distillable

    Alcohol-benzene or Fatty acids, including unsaturated and sat-ether extractable urated fatty acids

    Fats and oils, waxes, resins, resin acids,and sterols

    Alcohol extractable

    Water extractable

    Individual or classes of compounds

    Terpenes, including sesquiterpenes, diter-penes, triterpenes, tetraterpenes, andpolyterpenes

    Phenols, hydrocarbons, and lignans

    Coloring matter, including flavonoids andanthocyanins

    Phlobaphenes, tannins, and stilbenes

    Carbohydrates, including monosaccharides,starch and pectic materials

    Proteins, alkaloids, and inorganic materials

  • 18 River et al.

    With regard to wood as an adherend, extractives are extremelyimportant because of their often undesirable and unpredictable ef-fect upon adhesive bonding. As indicated in Table 1.2, innumer-able opportunities exist for chemical reactions between extraneousmaterials and the atmosphere, and between these materials and ad-hesives or other chemicals that may contact the materials at thewood surface. The pH, buffering capacity, and acid content ofthe wood can be strongly affected by the type and amount of ex-tractives. The setting or curing reactions of some adhesives havebeen reported to be sensitive to these factors.

    Woods that are very acidic, such as the oaks, Douglas fir, andkapur, are sometimes difficult to bond with adhesives that are sen-sitive to extractives. Mizumachi (1973) studied the effects of 18species of wood with varying amounts and types of extractive con-tents upon the activation energy of the urea-formaldehyde curingreaction. The resin cured with an activation energy of 29 kcal/mol. When wood flour of the various species was added to theresin, the activation energies of the filled resin ranged from 26 to63 kcal/mol. Wood flours, including red and white lauan, apitong,and sugi, had virtually no effect on the reaction. However, woodflours such as septir (38 kcal/mol), kapur (39 kcal/mol), and del-lania (63 kcal/mol) had strong effects. Similarly, extractives ob-tained from pressure refining a group of five hardwoods and lob-lolly pine decreased the gel time of a urea-formaldehyde resin whenadded in small amounts. The addition of about 6-9% of alcohol-soluble fractions shortened gel time by as much as 70%. Water-soluble fractions had a lesser, although still strong, effect (Slay,Short, and Wright 1980).

    Extractives that are insoluble in the adhesive-solvent systemmay cause more adhesion problems than extractives that are sol-uble. For example, Narayanamurti, Gupta, and Verna (1962)found that the extractives of teak (Tectona grandis) that are in-soluble in water, although soluble in alcohol/benzene, adverselyaffected the setting of water-based animal and urea-formaldehydeadhesives. The extractives of acacia that are soluble in hot waterdid not interfere with either animal glue or urea-formaldehyde.

    Gardner (1965) describes an interesting test ascribed to Sander-man, Dietrichs, and Puth (1960) for compatibility between a finishand various extractives. A paper chromatogram of a solution ofthe extractives is made and then coated with finish. The spots ofthe chromatogram are observed for signs of failure. Specific typesof interference, such as interference with drying, discoloration,or cracking, can then be associated with specific types of extrac-tives. The same technique might be used to detect effects of ex-tractives upon the curing behavior of adhesives.

  • Wood as an Adherend 19

    From the physical standpoint, heavy concentrations of extrac-tives can physically block an adhesive from the intimate molecularcontact with the wood substance that is necessary to form a strong,durable bond. These interactions will be discussed in more detaillater in the chapter.

    C. Physical Properties

    Density

    Wood substance normally accounts for most of the weight of wood inservice, followed by water and extraneous materials. Wood sub-stance, that is, cellulose, hemicellulose, and lignin, has a densityof about 93.6 lb/ft3, regardless of the species. But wood is a por-ous material, so the void volume, and consequently the amount ofwood substance, varies with the anatomy of species, the growthrate of the tree, and even the position of the wood within a tree.Wood is also hygroscopic (see next section), and so its density alsovaries with the environment. Because water is less dense (62.4 lb/ft3) than wood substance, wood density decreases as the moisturecontent increases. For example, increasing the moisture contentfrom 8% to 28% decreases specific gravity from 0.54 to 0.44. Wooddensity is the weight of wood substance, extractives, and water perunit volume. It is usually expressed in pounds per cubic foot (gramsper cubic centimeter). Specific gravity is the ratio of the density ofthe wood to the density of a standard substance, usually water. Spe-cific gravity is usually determined by measuring the volume at agiven moisture content (e.g., green, 12%) and the oven-dry weight.The extractive content is not usually a major contributor to density.However, in some cases, extractives do constitute up to 30% or moreof the dry weight of wood.

    When dry, the least and most dense woods weigh about 2.3 and89 lb/ft3, respectively, which translate to specific gravities of 0.04and 1.42. Among the familiar woods, balsa weighs 10 lb/ft3 (spe-cific gravity 0.16) and oak about 44 lb/ft3 (specific gravity 0.75).Most commercial species fall within the range of 19-50 lb/ft3 (spe-cific gravity 0.30-0.80). Not surprisingly, density or specificgravity is the best single indicator of the mechanical properties ofwood. This relationship is discussed in detail later in this sec-tion. As applied to adhesive bonding, the strength of adhesive-bonded joints is strongly dependent upon the wood density, as isdiscussed further in Section VI.

    The challenge presented in bonding extremely low-density wood,like balsa, is to prevent overpenetration of the wood by the adhe-sive. Overpenetration produces an adhesive-starved, and thus

  • 20 River et al.

    weak, joint. Fortunately, very strong joints are not required toexceed the strength of very low-density wood. At the oppositeend of the scale, very dense wood may be stronger than the adhe-sive. However, low porosity and permeability and increased amountsof extractives in the very heavy woods may be more important.These characteristics limit mechanical adhesion and the ability ofthe adhesive to penetrate the wood surface and to make intimatemolecular contact over a large surface area. In addition, high ex-tractives content in dense woods increases the opportunities forinterference with wetting and cure.

    Even if the average specific gravity of a wood is well within therange of gravity for easy bonding, the disparity between earlywoodand latewood densities in a coarse-textured wood can make bondingdifficult. The southern pines provide a good example. The aver-age specific gravities of the southern pines range from 0.51 to 0.59,which is well within the range of density that can be readily bond-ed. However, southern pines are often difficult to bond well.Their wood presents two problems. First, the wood contains highlevels of oleoresinous extractive materials. Second, the earlywoodis very low in specific gravity (about 0.3), and thus it is veryporous and conducive to overpenetration by the adhesive. Thelatewood, on the other hand, is very high in specific gravity (about0.8), and thus it is nonporous and nonconducive to penetration andadhesion, Similar difficulties are often experienced in bondingcoarse-textured hardwoods such as ash and oak, which are ring-porous.

    In spite of these interacting factors, density is the best singleindicator of the mechanical properties, swelling and shrinking be-havior, and difficulty of bonding that can be expected of a wood,and of the probable durability of the bonded joints and wood prod-ucts.

    Hygroscopicity

    In the living tree, wood holds water as free water in the cell lumenand as bound water within the cell wall. Free water, which occursin the lumens of sapwood and to some extent in the heartwood cellsof the living tree, may range from 30% of the oven-dry weight ofthe wood in the heartwood and up to 250% of wood weight in thesapwood. When a tree is cut and converted to wood products, thefree water is removed by air or kiln drying; once removed, freewater will never return unless the wood is soaked in water. Thebound water in the cell wall is attracted to the free hydroxyl groupsof the cellulose, hemicellulose, and lignin, where it is adsorbed inmono- or polymolecular layers between the microfibrils and other sub-microscopic spaces in the cell wall. At normal service temperatures

  • Wood as an Adherend 21

    Figure 1.6 The relationship of the moisture content (MC) of woodat equilibrium with the humidity (RH) and temperature of the sur-rounding air.

    and humidities, the bound or hygroscopic water in wood is always inbalance, or at least tending toward a balance, with the environment.This balance is called the equilibrium moisture content (EMC) (Fig-ure 1.6). In theory, the attraction of the hydroxyl units for thewater is balanced by the force required to separate or push apartthe cellulose microfibrils. Increasing the temperature decreases theamount of moisture adsorbed, and consequently the EMC. at a givenrelative humidity. Bound water has the most influence and is of thegreatest concern with regard to wood as an adherend.

    When the relative humidity is zero, the EMC is zero. This isoften referred to as the ovendry moisture content. At the otherextreme, when the relative vapor pressure is 1, the EMC reachesthe fiber saturation point. At this point, the cell wall is fullysaturated. The actual moisture content at the fiber saturationpoint varies with species, tree, temperature, and pressure, in therange of about 26-34% of the oven-dry weight of the wood. Thisis a critical point. Below the fiber saturation point, wood swellsand shrinks ; significant changes in mechanical properties occur asthe moisture content changes. The addition of water above the fi-ber saturation point (by soaking in water), on the other hand, doesnot affect the mechanical properties of wood.

  • 22 River et al.

    In typical service environments, wood naturally maintains a mois-ture content between 5 and 20% of the oven-dry weight of the wood.The actual value is determined by the surrounding temperature andrelative humidity. If the environment remains constant, the woodwill reach EMC. However, most service environments vary contin-ually, so there are always slight changes in wood moisture content.Short-term changes, such as daily fluctuation of humidity, only af-fect the wood surface. Seasonal changes may affect the moisturecontent in the core of a wood member, but the greater the thick-ness, the slower the moisture content change in the core of themember. The EMC in the core of a large member may only changeslightly in response to long-term seasonal changes in the environ-ment. Coatings such as paint, varnish, and lacquer can alsodampen the hygroscopic response of wood. Changes of the boundwater level in the cell wall (moisture content) affect the density,dimensional stability, and mechanical properties of wood, and notsurprisingly, the bonding process and bond performance. Mois-ture content and density may exert an interactive effect on adhe-sive bonding; however, this effect is probably not of practicalsignificance compared to their separate effects. The effects ofboth density and hygroscopicity on bonding and bonded productsare discussed in Sections V and VI.

    Anisotropy

    Wood is an anisotropic material. Its physical and mechanical prop-erties differ in the three principal directions relative to the trunkof the tree (Figure 1.7a):

    Longitudinal: Parallel to tree trunk and parallel to long axis oflongitudinally oriented cells (tracheids and fiber tracheids).

    Radial : Perpendicular to longitudinal direction and parallel to ra-dius of trunk and wood rays.

    Tangential: Perpendicular to longitudinal direction and parallel togrowth rings.

    The properties of wood are also often referred to an orthotropicplane such as the tangential/radial (TR), longitudinal/radial (LR),and longitudinal/tangential (LT) planes shown in Figure 1.7b.

    Part of the explanation for the anisotropy of wood is the elon-gated shape of the majority of wood cells, and their orientation inthe longitudinal or grain direction of the wood. But anisotropy ismanifest at all levels of wood structure. The cell wall is also an-isotropic. In the dominant layer of the cell wall, the S-2 layer,the cellulose chains and the microfibrils are predominately orientedin the longitudinal direction of the cell itself.

  • Wood as an Adherend 23

    Figure 1.7 The pseudo-orthotropic structure of wood in relation(a) to the tree and (b) to the grain direction and growth rings.

    The properties of wood differ in each principal direction, muchas with fiber-reinforced plastic composite materials. The propertiesalso vary as a function of the angle between the principal directions.With regard to adhesive bonding and performance, differences andvariations in permeability, swelling and shrinking, and strength areof the greatest significance.

    Porosity and Permeability. In the context of anistropy, theseterms refer to the macro- and microscopic pathways by which a

  • 24 River et al.

    liquid or vapor passes through a piece of wood. Stamm (1964b)says that wood is highly porous, but not very permeable. This isbecause the cell lumens, which largely account for high porosity,are discrete. The void volume or porosity of commercial woodsranges from 45 up to 80% of the total wood volume, but the pitsand smaller cell wall voids provide poor communication betweenthe larger voids (lumens).

    The formation of high-quality joints is dependent on the poros-ity or macroscopic pathways that allow a liquid adhesive to pene-trate several cells below the surface. Penetration, in many in-stances, seems to be a requirement for high-performance joints.First, penetration allows the adhesive to repair damaged cells;second, it also diffuses the stress concentration between the woodand the adhesive at the interface; third, it increases mechanicalinterlocking and surface area for bonding.

    Hardening of many adhesives is dependent on the permeabilityor microscopic pathways for removal of adhesive solvents or liquidcarriers. Most adhesives used with wood have water or other liq-uid carriers, and many adhesives release water of condensation asthey cure. The water must be removed from the bondline in atimely fashion for nonchemically curing adhesives such as poly-(vinyl acetate). The water must be removed at just the righttime with respect to the chemical curing reaction of adhesives, suchas urea-formaldehyde, to develop the best joint.

    Water and other liquids move through wood in three ways: (1)as vapor in the lumens under a vapor pressure gradient, (2) asadsorbed moisture in the cell walls under a moisture content gradi-ent (but in effect a relative vapor pressure gradient), and (3) ascapillary-entrained water in the cell-lumen/pit system under a liq-uid surface tension gradient.

    One can easily visualize that water and other liquids move mostrapidly in the grain direction either as liquid or vapor because themajority of long, hollow cells are oriented that way. In fact, thecell lumens are capillary tubes capable of drawing liquids into thewood interior. The force exerted by surface tension in tubes thesize of cell lumens and pits is theoretically strong enough to lift acolumn of water 390 ft against gravity (Tarkow 1981). The size ofthe tracheid in softwoods or vessel lumens in hardwoods and thepresence or absence of blockages in the heartwood of hardwoodshave a considerable effect on the movement of water and, in par-ticular, on the penetration of adhesives into the wood structure.

    In the tangential and radial directions, pathways into the woodstructure are few and indirect. The magnitude of anisotropy inporosity and permeability can be extreme. The ratios of longi-tudinal to tangential or radial permeability can be as great as 1

  • Wood as an Adherend 25

    million to 1 (Comstock 1970). When a liquid or vapor reaches a cellwall, it must diffuse through the wall or pass through the pits.Many large pits on the radial cell walls assist movement in the tan-gential direction. Even so, passage is interrupted 50 times moreoften in the tangential direction than in the longitudinal direction.The pits sometimes close during drying of the wood, and this makespenetration by water or adhesives even more difficult in the tan-gential direction. Movement in the radial direction is further re-stricted because pits on the tangential cell wall are smaller and in-frequent. This forces radial movement by diffusion or circuitousflow along a longitudinal-tangential path. Density and extractivesprimarily reduce permeability by reducing the void volume of thewood through which the liquid or vapor can pass.

    The anisotropic aspects of porosity have several effects on theflow of adhesives during bonding. First, they affect how andwhere the adhesive moves after spreading, and when pressure isapplied during bonding. Stamm (1973) found that water at atmos-pheric pressure penetrated white oak and loblolly pine more than25 times faster in the grain direction than laterally. We have pre-viously mentioned that woods of high porosity provide the potentialfor robbing or starving the joint of adhesive. The fact that woodon the end grain is most porous explains in part the weakness ofend-grain to end-grain adhesive bonds. On the radial and tan-gential surfaces, porosity is much reduced. Sufficient pressurecan be applied to achieve a thin, uniform bondline without forcingall the adhesive out of the bondline, although there may still belarge differences between earlywood and latewood. Very porouswoods or high-density woods with zones of high porosity, like theoaks and southern pines, are often difficult to bond even on theradial and tangential surfaces because capillarity and bonding pres-sure draw or force adhesive away from the bondline, leaving thejoint in an adhesive-starved condition in those regions.

    Anisotropic permeability also affects the rate of loss of water orother liquid carriers from an adhesive. Some adhesives are quitesensitive to the relationship between the liquid carrier content andthe cross-linking reaction. If the carrier is removed too rapidly,the adhesive molecules will not have the mobility necessary for op-timum cross-linking. If the carrier is removed too slowly, fullcure may not be achieved (Moult 1977, Pillar 1966). Thus, dif-ferences between the radial and tangential surface permeabilitiesand between earlywood and latewood permeabilities may account forsome variability in bond quality that occurs between radial and tan-gential surfaces and in coarse-textured woods that are likely to ex-hibit large differences in the permeability between earlywood andlatewood.

  • 26 River et al.

    Swelling and Shrinking. Between the oven-dry moisture contentand the fiber saturation point, wood swells or shrinks dependingon whether it is gaining or losing moisture. Longitudinal shrinkageamounts to only 0.1-0.3%, but it can have a strong effect becauseof the high longitudinal modulus. Longitudinal shrinkage is signifi-cantly greater in veneers with cross-grain, in reaction wood, andin juvenile wood than in normal straight-grain wood. Even smalldifferences in longitudinal shrinkage can have significant effectsupon the performance of certain types of bonded products. Lat-eral shrinkage (radial and tangential) is much greater. Lateralshrinkage varies greatly between species, with density, and be-tween the radial and tangential directions within a given species(Table 1.3). In common U.S. woods, the tangential shrinkageranges from about 4.5 to 12.5% over the range of moisture contentfrom the fiber saturation point to oven dry. The radial shrinkageranges from about 2.0 to 8.5% over the same range. Generally,tangential shrinkage is two times greater than radial shrinkage;

    Table 1.3 Tangential and Radial Shrinkage from Fiber Saturation toOven-Dry Condition for Selected Species

    Shrinkage (%) Tangential/radialSpecies Tangential Radial

    Southern magnolia 6.6 5.4 1.2

    Yellow birch 9.2 7.2 1.3

    Eastern redcedar 4.7 3.1 1.5

    Douglas fir 7.8 5.0 1.6

    Redwood 4.4 2.6 1.7

    Hard maple 9.5 4.9 1.9

    Red oak 8.9 4.2 2.1

    American beech 11.0 5.1 2.2

    Western hemlock 6.8 3.0 2.3

    American elm 9.5 4.2 2.3

    White pine 6.0 2.3 2.6

    Black willow 8.1 2.6 3.1

  • Wood as an Adherend 27

    however, the ratio varies from 1.2 to 3.3 depending on species(Forest Products Laboratory 1987, Noack and Schwab 1973). Thesedifferences are thought to be due to the restraint of ray cells onswelling and shrinking of the longitudinal cells in the radial direc-tion. The lower density earlywood also has a lower tendency toswell and shrink (swell-shrink coefficient) compared to the late-wood. Swelling in the tangential direction is dominated by thehigher swelling and shrinking latewood, which forces the earlywoodto move as the latewood moves. In the radial direction, however,the lower swelling and shrinking earlywood can act independently,thus minimizing radial movement.

    When wood swells and shrinks, stresses develop that can rupturethe adhesive bond or the wood, whichever is weaker. Stresses de-velop because the dimensional changes are anisotropic. Woods withlow tangential (T) and radial (R) shrinkage coefficients and low T/Rratios (Table 1.3) are more stable and less likely to warp, crack,or delaminate when moisture content changes after bonding or laterin service. The T/R anisotropy of swelling and shrinking is a crit-ical factor in the performance of most adhesively bonded joints infurniture construction. Anisotropy must also be considered in theselection and machining of lumber for edge-glued panels. The T/Rratios of 2.0 are considered favorable, normal,and unfavorable, respectively, from a technological view (Noack,Schwab, and Bartz 1973).

    Actually, adhesive bonding can be used to overcome some prob-lems caused by anisotropy, or even to take advantage of aniso-tropy, as illustrated by the following examples. First, lumbercan be dried free of strength-reducing checks; large, solid beamscannot. However, large beams can be made free of checks bybonding together pieces of carefully dried, check-free lumber.Second, T/R anisotropy of swelling and shrinking is very smallcompared to T/L and R/L anisotropy. These last ratios may beas high as 100:1. Wood products manufacturers take advantageof this fact to create dimensionally stable wood panels. Cross-banded furniture panels, plywood, and structural flakeboardsare dimensionally stable in the plane of the board because lat-eral swelling and shrinking of the components is restricted byadjoining components. The grain directions of adjoining lumber,veneer, or flakes are by design or chance at an angle, often aright angle, to each other. Thus, lateral movement of each pieceof lumber, veneer, or flake is restrained through the adhesivebond by the low longitudinal movement and high stiffness of itsneighbor. As a result, movement in the plane of the panel isnot much greater than movement of solid wood in the longitudinaldirection.

  • 28 River et al.

    D. Mechanical Properties

    Wood strength and other mechanical properties are extremely vari-able between species, within a species, within a given tree, and indifferent directions within the tree. In this section, we will outlinethe mechanical properties of greatest importance to wood as an ad-herend and discuss their variability.

    Strength

    Fibers and Clear Wood. Strength is usually of most concern.The strength of individual wood fibers varies widely. Fibers areextremely strong in the longitudinal direction, with tensile strengthsin the range of 40,000-140,000 lb/in. 2 based on the type of cell andwood species (Mark 1967). Wood, for many reasons, is not thatstrong, but in the longitudinal direction, the strength of clearstraight-grained lumber parallel to the grain ranges from 10,000lb/in.2 to as high as 20,000 lb/in.2.

    An adhesive bond would require similar strengths to join woodfibers or lumber pieces end to end (a tensile butt joint). Neitherpeople nor nature has devised an adhesive capable of such strength.In a simplistic way, the stress on the adhesive bond in a tensilebutt joint is equal to the force applied at the ends of a fiber orwood member divided by the area of the bond (Figure 1.8a). Ob-viously, a larger bond area would reduce the stress. Unfortunately,the bond area of the end of a fiber or piece of lumber is limited tothe diameter of the fiber or the transverse dimensions of the lumber.However, this problem can be overcome by overlapping and bondingfibers or members to form joints that are stressed in shear insteadof tension. The bond area on the side of the fiber or member canbe easily increased by increasing the amount of overlap. Thus, thesame force can be transmitted from one fiber or member to anotherat a much lower bond stress (Figure 1.8b.c). This explains whyboth people and nature rely on large, lateral, shear bond areas (lapjoints). (Notice the overlapping shear joints between fibers in Fig-ure 8c.)

    This example explains why the shear strength of wood parallel tothe grain (Figure 1.9a,b) is particularly important. Tension strengthperpendicular to the grain (Figure 1.9c-f) is important because sig-nificant tensile stress perpendicular to the grain arises in most shear-type joints, and when wood shrinks after bonding. Shear strengthperpendicular to the grain, or “rolling shear” (Figure 1.9g,h), isimportant in plywood and certain types of assembly joints where theadherends are bonded with their grain directions at an angle to eachother. The term rolling shear arises from the tendency of the cellsto roll under this type of loading. Rolling-shear strength has not

  • Wood as an Adherend 29

    Figure 1.8 Adhesive bond areas of different types of joints: (a)the fixed maximum area of a tensile butt joint, (b) the easily in-creased or adjustable shear bond area of a lap joint, and (c) thelarge shear bond area formed by the overlap between the taperedends of the two wood tracheids. Note that the bond area of thebutt joint is limited by the cross-sectional area, but the bond areaof the lap joint is limited only by the length of the overlap.

    been measured for individual fibers, but rolling-shear strength val-ues are undoubtedly much lower than the longitudinal fiber strengthvalues. Representative shear parallel to the grain, tension perpen-dicular to the grain, and rolling-shear strength values of varioussolid woods are given in Table 1.4.

    Compression strength perpendicular to the grain is importantduring processing the wood surface for bonding and when apply-ing pressure during bonding. Pressure from dull or improperly

  • 30 River et al.

    Figure 1.9 Shear and tension modes of loading wood in relation toits orthotropic structure: (a) shear in the longitudinal direction onthe LR plane, (b) shear in the tangential direction on the LT plane,(c and e) tension perpendicular to the grain on the LR (or RL)plane, (d and f) tension perpendicular to the grain on the LT (orTL) plane, (g) shear in the radial direction on the RL plane (roll-ing shear), and (h) shear in the tangential direction on the TLplane (rolling shear).

    sharpened cutters, from abrasive planing, and from feed rolls orpressure bars can exceed the strength of the wood and cause per-manent and irreparable damage. Excessive pressure during bond-ing can also cause permanent damage that weakens the surface anddetracts from the strength and durability of the bonded joint.

  • Wood as an Adherend 31

    Table 1.4 Strength of Selected Woods at 12% Moisture Content underVarious Types of Loading

    Species

    Ash spp.Quaking aspenBeech spp.Yellow birchBlack cherrySugar mapleRed oakWhite oakBlack walnutYellow poplarPort orford cedarWestern redcedarCoast Douglas-firBalsam firSubalpine firWhite firWestern hemlockPine spp.Loblolly pinePonderosa pineSugar pineWestern white pineRedwoodBlack spruceEnglemann spruceWhite spruceRed spruceSitka spruce

    Strength values (lb/in.2)

    Shearparallelto grain

    2232-2815850

    2090-236018801700233017802000137011901370990

    1130944

    107011001290

    1265-15921390113011301040111012301200970

    12901150

    Tensionperpendicular

    to grain

    260782–1308a

    625–1038a

    920560–

    800800690540720220340180–

    300340

    470420500

    –250550350430350370

    Rolling shear

    325b

    194b

    216b

    208b

    256-569a

    293b

    297b

    273b

    259b

    259b

    252c

    aKollman and Cote (1968).bBendtsen (1976).cMunthe and Ethington (1968).

  • 32 River et al.

    Species Differences. Based on the data in Table 1.4, shearstrengths parallel to the grain range from about 1000 to 3000 lb/in.2,whereas tension strengths perpendicular to the grain range fromabout 200 to 1000 lb/in.2 The importance of the rolling-shearstrength of wood has only recently been recognized so little dataare available, but the rolling-shear strength for a given species isapparently in the same range as the tension strength perpendicularto the grain.

    Based on tests of some 50 species, the coefficient of variationfor wood strengths are 14% for shear parallel to the grain and 25%for tension perpendicular to the grain at a given moisture content(Forest Products Laboratory 1987). Of course, many factors affectthe strength of wood in service, so the variation of individual piecesof wood of a given species may be higher than these values. Amongthe most important factors are density, moisture, temperature, andgrowth characteristics, such as knots and other grain deviations.

    Density Effect. The strength of clear, straight-grain, defect-free wood can be approximated by the relationship

    S = KGn

    where S is a property such as shear strength, K and n are con-stants for that property, and G is the specific gravity. The expo-nential factor n ranges from 0.55 to 2.25, depending on the property,on whether the wood is a softwood or a hardwood, and on the woodmoisture content. As an example, the Wood Handbook (Forest Prod-ucts Laboratory 1987) provides the following general relationshipsfor the shear strength parallel to the grain of softwoods and hard-woods :

    S = 2430G0.86 for softwoods

    S = 3200G1.15 for hardwoods

    The fact that n exceeds 1.0 means that the strength increasesfaster than would be expected from the simple increase in the amountof wood substance. The explanation for this increase must lie in thechanging ratios of the three major cell wall constituents and thechanging ratios of the three secondary cell-wall layers. For a givenspecies, the specific gravity, shear plane (LR compared to LT),moisture content, and extractive content are major causes of varia-tion in the relationship of specific gravity to mechanical property.

    High levels of extractives distort the relationship between spe-cific gravity and mechanical properties. For example, in an unpub-lished experiment conducted by one of the authors, a specimen ofwestern larch wood was found to have a dry specific gravity of

  • Wood as an Adherend 33

    0.55. Based on this density, the relationship of density to specificgravity predicts a modulus of rupture of 13,665 lb/in.2; however,the measured value was only 6048 lb/in.2 Hot-water extraction re-vealed that 30% of the dry weight of the wood consisted of extrac-tives. The dry specific gravity of the wood after extraction or, inother words, the specific gravity based primarily on the amount ofwood substance was only 0.36. The predicted modulus of rupturebased on the extractive-free density was 8757 lb/in2, much closerto the observed value. Wood with high extractive content may ac-tually be stronger in compression and hardness perpendicular to thegrain than expected for its density, but lower than normal in bend-ing, tension, and other strength properties.

    Moisture Content Effect. Air- and kiln-dried wood are hygroscopicAdsorption of water and other polar liquids expands the intermolecu-lar spaces and reduces direct hydrogen bonding within the cell wall.These actions plasticize the wood and reduce its strength. Betweenthe oven-dry condition and the moisture content at which all theintermolecular spaces are fully expanded (fiber saturation point),the strength decreases by 40-60% depending on the property andthe species of wood. The sensitivity of the strength property tomoisture content varies with the species and property, as shown inFigure 1.10. The relationship can be described by the equation(Forest Products Laboratory 1987) :

    Figure 1.10 The temporary effect of wood moisture content on shearstrength parallel to the grain and tensile strength perpendicular tothe grain.

  • 34 River et al.

    where P is strength at the desired moisture content, P12 is strengthat 12% moisture content, Pg is strength of wood above the fiber sat-uration point, Mp is moisture content approximating the fiber satura-tion point below which strength begins to change, and M is moisturecontent for which the strength is to be determined.

    Values for P12 and P are tabulated for most commercial NorthAmerican woods in the Wood Handbook (Forest Products Laboratory1987) and in ASTM D 2555 (ASTM 1989a). The relationship betweenstrength and moisture content is temporary and largely reversiblewithin normal service temperatures and the expected service environ-ment of most wood products.

    Temperature Effect. Temperature has temporary and permanenteffects on strength. The temporary effects are reversible and linearin the range between at least -70 and 150°C as long as the moisturecontent is constant. Increasing moisture content increases the rateof change of strength with temperature (Figure 1.11). For example,increasing the service temperature from 20 to 50°C will cause thefollowing reductions in shear and tensile strengths (Forest ProductsLaboratory 1987) :

    Moisture content Strength lossStrength property (%) (%)

    Shear parallel to the grain Above fiber saturation 25point

    Tension perpendicular to 4-6 10the grain 11-16 20

    As mentioned previously, elevated temperatures have a measur-able and permanent effect on strength. The permanent effect fol-lows the Arrhenius time-temperature relationship (Stamm 1964b) :

    log10 t = A + B/T

    where t is aging time, T aging temperature in degrees Kelvin, Amaterial constant, and B temperature coefficient. Experimental re-sults suggest that if wood is kept dry, it will lose about 25% of itsoriginal strength in about 2500 years. This is borne out by thecondition of dry wood artifacts discovered in Egyption tombs. Thethermal/chemical changes wrought by short-term exposures to highertemperatures are cumulative and are not recoverable. The same 25%

  • Wood as an Adherend 35

    Figure 1.11 The temporary effect of temperature on the compres-sion strength parallel to the grain at two moisture contents relativeto the strength at 20°C.

    loss will occur within only 1000 to 2000 days of aging at 100°C andin only 1-2 days of aging at 170°C.

    Moisture accelerates cellulose hydrolysis at elevated temperatures.The moisture content of the wood, particularly if it is high, increasesthe temperature coefficient by at least 10 times. Wet wood will lose25% of its original strength in 80-400 years at 20°C. The same losswill occur in 100-2000 days at 60°C, and in only 2-4 days at 100°C(Millett and Gillespie 1978). Equations for the time to lose 25% ofthe original shear strength of two species are as follows (Millett andGillespie 1978) :

    Condition

    Species Oven-dry Soaked

    Hard maple log10 t = -17.308 + 7614/T log10 t = -15.185 + 5758/T

    White pine log10 t = -16.222 + 7268/T log10 t = -16.145 + 6246/T

  • 36 River et al.

    Effect of Grain and Growth–Ring Angle Effect. Aside from the ob-vious effect of a large knot or knothole, the direction of loading withrespect to the grain angle and the growth-ring direction probablyhas the most pronounced effect on mechanical properties. That thestrength of wood differs in various directions of the grain is com-mon knowledge. For example, we know that wood can be split quiteeasily along the grain, but we must laboriously chop or saw it acrossthe grain. As children we are taught to hold a baseball bat with thetrademark (tangential face) up as it strikes the ball because the batis less likely to break than if held the other way. These are famil-iar effects of the anisotropic mechanical properties of wood.

    The outstanding effects of wood anisotropy are due to large dif-ferences between the :

    1. Elastic modulus and strength in the longitudinal direction, andthe elastic moduli and strengths in the tangential and radial di-rections.

    2. Rolling-shear modulus and strength, and the shear moduli andstrengths in the longitudinal/tangential and longitudinal/radialdirections.

    3. Lateral contractions in the radial and tangential directions underlongitudinal load, and the longitudinal contraction under eithertangential or radial loading.

    The general relationships of the major mechanical properties ofwood in the different directions are shown in Table 1.5. In thetable, the tensile modulus along the grain is in the range of 13-20times higher than that of properties across the grain. Kollman andCote (1968) summarized the properties for a group of 7 softwoodsand 14 hardwoods from several sources. Their summary shows thatthe difference between the longitudinal and tangential directions ison the order of 10-25 times, whereas the difference between longi-tudinal and radial directions is on the order of 5-20 times. Insome species, the longitudinal properties may be less than 10 timesor greater than 40 times the lateral property. However, thesestrengths are for clear wood; the differences are considerablysmaller when comparing lumber with typical defects, such as knots,that greatly reduce the longitudinal strength. The longitudinalshear strength is about 30 times higher than the shear strengthin the tangential and radial directions (Table 1.5). As a generalrule, the directional differences in the mechanical properties ofhardwoods are somewhat smaller than these differences in softwoods.The smaller difference in hardwoods is thought to be due to the ef-fects of the greater volume of rays (Schniewind 1980); rays pre-sumably increase properties in the radial direction with some sacri-fice of properties in the tangential and longitudinal directions.

  • Wood as an Adherend 37

    Table 1.5 Representative Mechanical Properties of Wood in the ThreePrincipal Directions

    Elastic Contraction/elongationa

    Strength modulus Longi- Tangen-(lb/in.2) (lb/in.2) tudinal Radial tialb

    Tensile load directionLongitudinal 15,000 2,000,000 – 0.30 0.45Radial 470 150,000 0.04 – 0.40Tangential 420 100,000 0.03 0.40 –

    Plane of shear failureLongitudinal/radial 1,370 120,000Longitudinal/tangential 1,430 150,000Radial/tangential 350 15,000

    aPoisson’s ratio.bDirection of contraction.

    Both modulus and strength differences between the radial andtangential directions are much smaller, generally not exceeding 1.5times in North American woods. Strength of the longitudinal/tan-gential (LT) plane is often higher. Data from a Canadian source,which lists both LT and longitudinal/radial (LR) shear strengths,offer the opportunity to compare these differences (Forest Prod-ucts Laboratories of Canada 1956). Table 1.6 shows the averagepercentage of difference between the strengths in the LT and LRplanes for a group of 21 softwood and a group of 32 hardwood spe-cies. In most species, the strength in the LT plane is higher thanthat in the LR plane, and the difference between LT and LR strengthsis less than 10%. In contrast, the tensile modulus when load is ap-plied in the radial direction is usually higher than that when loadis applied in the tangential direction (Table 1.5). This is due totwo factors.rection by the ray cells, which are oriented radially instead of longi-

    The first factor is the reinforcement in the radial di-

    tudinally. The second factor is the additional bending of the cellwalls in the tangential direction stemming from the staggered posi-tion of cells in adjacent rows. The ratio of lateral contraction toelongation in the direction of an applied tensile load (Poisson’s ra-tio) is large (0.30 to 0.45) when load is applied parallel to thegrain, but quite small (0.03 to 0.04) when load is applied perpen-dicular to the grain.

  • Ta

    ble

    1.6

    Dif

    fere

    nce

    s in

    Str

    engt

    h B

    etw

    een

    th

    e T

    ange

    nti

    al-L

    ongi

    tudi

    nal

    an

    d R

    adia

    l-L

    ongi

    tudi

    nal

    Pla

    nes

    Dif

    fere

    nce

    (%)

    Ran

    geM

    ean

    Pla

    ne

    wit

    h h

    igh

    er s

    tren

    gth

    Dif

    fere

    nce

    s (n

    um

    ber)

    (nu

    mbe

    r)

    Ove

    rO

    ver

    Ove

    r T

    ange

    n-

    No

    10%

    20%

    30%

    tial

    Rad

    ial

    chan

    ge

    21 S

    oftw

    oods

    Sh

    ear

    para

    llel

    to

    grai

    nT

    ensi

    on p

    erpe

    ndi

    cula

    rto

    gra

    in

    0–12

    .14.

    12

    00

    710

    40–

    32.2

    12.4

    142

    117

    22

    32

    Har

    dwoo

    dsS

    hea

    r pa

    rall

    el t

    o gr

    ain

    0.6–

    35.4

    10.0

    162

    128

    40

    Ten

    sion

    per

    pen

    dicu

    lar

    0.9–

    36.9

    18.7

    2716

    531

    10

    to g

    rain

    38 River et al.

  • Wood as an Adherend 39

    The differences in strength and elastic properties such as thosesummarized in Tables 1.5 and 1.6 are responsible for much of thestress that develops in adhesively bonded joints and materials as thewood swells and shrinks. Changing the grain direction by 90° withrespect to the applied load direction reduces shear strength by 60-80% and tensile strength by 93-96%. These large differences in prop-erties between orientations parallel and perpendicular to the grainhave already been described (Table 1.5). However, loads are oftenapplied at angles between 0 and 90° to the grain direction becauseof the growth patterns, the way the board was cut, or the designof the bonded joint or product. Strengths at these intermediateangles follow a continuous function (Figure 1.12) often describedby a Hankinson-type relationship (Forest Products Laboratory 1987) :

    Figure 1.12 The effect of grain angle on mechanical properties ofwood according to the Hankinson-type formula. Q/P is the ratioof the mechanical property across the grain (Q) to that parallelto the grain (P) ; it is an empirically determined constant.

  • 40

    N= PQP sinn θ + Q sinn θ

    River et al.

    where N is strength, P is strength parallel to the grain, Q isstrength perpendicular to the grain, n is an empirically determinedconstant between 1.5 and 2, and θ is the angle between the loadand fiber directions. Generally, the ratios between strength val-ues parallel and perpendicular to the grain range from 0.04 to 0.07for tension and from 0.20 to 0.45 for shear, although shear behav-ior has not been thoroughly studied (see Table 1.4).

    Fracture

    Wood is a fibrous, laminar, anisotropic material. Cracks initiate andpropagate easily in planes parallel to the fibers, but with great dif-ficulty in the plane perpendicular to the fibers. In planes parallelto the fibers, the actual values of fracture toughness for wood pa-rallel to the fibers range from 50 to 1000 J/m2, and perpendicular,from about 10 to 30 kJ/m2. These values are comparable to artifi-cial fiber composites (Jeronimidis 1976).across the grain is 104 J/m2;

    The fracture toughnesscomparable on a weight basis with

    the energy consumed during crack propagation in ductile metals(Gordon and Jeronimidis 1974).

    Locus of Fracture. The modes of wood fracture most importantto adhesive bonding and bond performance are shear parallel to thegrain and tension perpendicular to the grain, or a combination ofthese. Transverse tensile fractures are common in earlywood cellsbut less common in latewood cells. In some instances, the damageresulting from excessive compression stress is important to per-formance. In most cases, however, fracture in service will actuallyoccur in shear or tension.

    At the molecular level, Porter (1964) found that fracture occursin the amorphous, water-accessible regions rather than in the crys-talline cellulose regions of the cell wall. At the microscopic level,wood fractures in different locations depend on the type of cell,direction of load, temperature, moisture content, speed of test,grain angle, pH of adhesive, pH of wood, aging of wood, and, inthe case of adhesively bonded wood, penetration of adhesive.

    At the microscopic level, there are three types of fracture: trans-wall, intrawall, and intercellular. A longitudinal transwall crackpasses through the cell wall and across the cell lumen (Figure 1.13a).A longitudinal intrawall crack travels within the cell wall and aroundthe lumen (Figure 1.13b). An intercellular crack occurs when hotand wet wood is fractured (Figure 1.13c). Transverse transwallcracks (Figure 1.13d) may also occur. These characteristic types

  • Wood as an Adherend 41

    Figure 1.13 Schematic diagrams of typical loci of fracture in wood,(a) longitudinal transwall, (b) intrawall, (c) intercellular, and (d)transverse transwall.

    of failure have been observed not only in solid wood but also in woodparticles bonded with droplets of adhesive and in bonds formed bycontinuous films of adhesive between solid wood surfaces (Wilsonand Krahmer 1976, Koran and Vasishth 1972).

    Transwall failures are characteristic of thin-walled cells, such asthe earlywood tracheids in softwoods and vessel and parenchyma

  • 42 River et al.

    cells in hardwoods (Figure 1.14a). In these cells, the transwallfracture breaks the relatively thin layer of fibrils and leaves asmooth surface or a surface with only short fibril ends exposed(Figure 1.14b). When transwall failure occurs in combined shearand tension, the crack path follows the helical winding of the S-2layer in a combined shear and tension failure (Figure 1.14c). Longi-tudinal transwall fracture of thick-walled cells is unusual but ex-tremely fibrous (Figure 1.14d). Transwall fracture in the LT planeoccurs preferentially in the first earlywood cells of a given growthring in ring-porous hardwoods like oak and in coarse-textured soft-woods like the southern pines. A mixture of transwall and intrawallfractures is more likely in the LR and intermediate planes