12.07.2015 Views

36 Drying of Wood

36 Drying of Wood

36 Drying of Wood

SHOW MORE
SHOW LESS
  • No tags were found...

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>36</strong><strong>Drying</strong> <strong>of</strong> <strong>Wood</strong>: Principlesand PracticesPatrick Perré and Roger B. KeeyCONTENTS<strong>36</strong>.1 Structure <strong>of</strong> <strong>Wood</strong> ............................................................................................................................... 822<strong>36</strong>.1.1 Formation <strong>of</strong> <strong>Wood</strong> in Trees .................................................................................................. 822<strong>36</strong>.1.1.1 Knots ...................................................................................................................... 822<strong>36</strong>.1.1.2 Tissue and Cellular Structure <strong>of</strong> <strong>Wood</strong> .................................................................. 822<strong>36</strong>.1.1.3 Heartwood Formation............................................................................................ 825<strong>36</strong>.1.2 Chemical Composition ............................................................................................................ 826<strong>36</strong>.1.3 Reaction <strong>Wood</strong> and Juvenile <strong>Wood</strong>........................................................................................ 827<strong>36</strong>.1.3.1 Compression <strong>Wood</strong> ................................................................................................ 828<strong>36</strong>.1.3.2 Tension <strong>Wood</strong>......................................................................................................... 828<strong>36</strong>.1.3.3 Juvenile <strong>Wood</strong> ........................................................................................................ 828<strong>36</strong>.1.4 Implications for the <strong>Drying</strong> Process........................................................................................ 829<strong>36</strong>.2 Board Scale .......................................................................................................................................... 830<strong>36</strong>.2.1 Water in <strong>Wood</strong>: Sorbed and Capillary Water, and Shrinkage................................................ 830<strong>36</strong>.2.1.1 Moisture Content <strong>of</strong> <strong>Wood</strong> .................................................................................... 830<strong>36</strong>.2.1.2 Free Water .............................................................................................................. 830<strong>36</strong>.2.1.3 Bound Water........................................................................................................... 831<strong>36</strong>.2.1.4 Differential Heat <strong>of</strong> Sorption.................................................................................. 831<strong>36</strong>.2.1.5 Shrinkage ................................................................................................................ 832<strong>36</strong>.2.2 Heat and Mass Transfer in <strong>Wood</strong>........................................................................................... 833<strong>36</strong>.2.2.1 Fluid Migration in <strong>Wood</strong>: Single-Phase Flow ........................................................ 833<strong>36</strong>.2.2.2 Generalized Darcy’s Law: Multiphase Flow........................................................... 835<strong>36</strong>.2.2.3 Capillary Pressure ................................................................................................... 8<strong>36</strong><strong>36</strong>.2.2.4 Bound-Water Diffusion .......................................................................................... 837<strong>36</strong>.2.2.5 Physical Formulation.............................................................................................. 839<strong>36</strong>.2.3 Process <strong>of</strong> <strong>Drying</strong> .................................................................................................................... 840<strong>36</strong>.2.3.1 Low-Temperature Convective <strong>Drying</strong> .................................................................... 840<strong>36</strong>.2.3.2 <strong>Drying</strong> at High Temperature: The Effect <strong>of</strong> Internal Pressure on Mass Transfer.. 841<strong>36</strong>.2.3.3 Typical <strong>Drying</strong> Behavior: Difference between Sapwood and Heartwood .............. 842<strong>36</strong>.2.4 Mechanical Aspects <strong>of</strong> <strong>Wood</strong> <strong>Drying</strong> ..................................................................................... 845<strong>36</strong>.2.4.1 Mechanical Behavior <strong>of</strong> <strong>Wood</strong> ............................................................................... 845<strong>36</strong>.2.4.2 <strong>Drying</strong> Stress Formulation ..................................................................................... 846<strong>36</strong>.2.4.3 Memory Effect........................................................................................................ 847<strong>36</strong>.2.4.4 Stress Development during <strong>Drying</strong>: Some Examples.............................................. 850<strong>36</strong>.2.5 <strong>Drying</strong> Quality ........................................................................................................................ 855<strong>36</strong>.2.5.1 Factors Affecting the <strong>Drying</strong> Duration .................................................................. 855<strong>36</strong>.2.5.2 Factors Affecting the <strong>Drying</strong> Quality ..................................................................... 856<strong>36</strong>.2.5.3 Criteria for Obtaining a Fast and Good <strong>Drying</strong> Process........................................ 856<strong>36</strong>.3 Kiln Scale ............................................................................................................................................. 857<strong>36</strong>.3.1 Lumber Quality....................................................................................................................... 857<strong>36</strong>.3.1.1 Gross Features <strong>of</strong> <strong>Wood</strong>......................................................................................... 858<strong>36</strong>.3.1.2 Intrinsic Features <strong>of</strong> <strong>Wood</strong>..................................................................................... 860<strong>36</strong>.3.1.3 Sawmilling Strategies .............................................................................................. 862ß 2006 by Taylor & Francis Group, LLC.


<strong>36</strong>.3.2 Kiln Design ............................................................................................................................. 862<strong>36</strong>.3.2.1 Airflow Considerations........................................................................................... 863<strong>36</strong>.3.2.2 Moisture-Evaporation Considerations.................................................................... 864<strong>36</strong>.3.3 Kiln Operation........................................................................................................................ 864<strong>36</strong>.3.4 Practical Considerations.......................................................................................................... 866<strong>36</strong>.3.4.1 Schedule Development............................................................................................ 867<strong>36</strong>.3.4.2 Kiln Control ........................................................................................................... 867<strong>36</strong>.3.4.3 Volatile Emissions................................................................................................... 868<strong>36</strong>.3.4.4 Equalization and Stress Relief................................................................................ 869<strong>36</strong>.3.5 Less-Common <strong>Drying</strong> Methods.............................................................................................. 869<strong>36</strong>.3.5.1 Vacuum <strong>Drying</strong> ...................................................................................................... 869<strong>36</strong>.3.5.2 Dehumidifier Kilns.................................................................................................. 870<strong>36</strong>.3.5.3 High-Frequency Electrical Heating......................................................................... 870<strong>36</strong>.3.5.4 Solar <strong>Drying</strong> ........................................................................................................... 872References ...................................................................................................................................................... 872<strong>36</strong>.1 STRUCTURE OF WOOD<strong>36</strong>.1.1 FORMATION OF WOOD IN T REES<strong>Wood</strong> is produced in the hard stems and branches <strong>of</strong>trees and shrubs. In these plants, the primary growthis responsible for the stem and branch elongation,whereas the secondary growth, achieved through thecambium activity, is responsible for the thickening <strong>of</strong>elements (Figure <strong>36</strong>.1). <strong>Wood</strong> has evolved to fulfill thebasic needs <strong>of</strong> these plants during their life: watertransport<strong>of</strong>waterbornenutrients;mechanicalstrengthtosupportthephotoactivecanopy<strong>of</strong>leaves;andresistanceto biological attacks. Appearance and properties<strong>of</strong> this material depend strongly not only on the speciesbut also on the biological diversity and growthconditions <strong>of</strong> site and climate. Indeed, even for thesame species, wood properties depend on the tree andon the position within the tree. One part <strong>of</strong> this variabilityis genetically controlled, whereas the otherpart comes from the varying growth conditions (standcharacteristics and silviculture practices). Consequently,wood is a variable material that is extremelydifficult to characterize precisely. Hence wood processing,including drying, is very difficult to optimize.A cross section <strong>of</strong> a tree (Figure <strong>36</strong>.2), from thecore to the outer region, shows the following features:. Pith, a small core <strong>of</strong> tissue located near themiddle <strong>of</strong> a tree’s stem or branches, which originatesfrom the primary growth <strong>of</strong> the plant. <strong>Wood</strong>y material, the most important part <strong>of</strong> maturetrees, which is differentiated into sapwood(outer region), where the sap migrates from rootsto leaves and heartwood (inner region) that is nolonger used for sap transport, which exists onlywhen the stem, at that height, is old enough. Bark, differentiated into an outer corky deadpart (external part <strong>of</strong> the stem), whose thicknessvaries greatly with species and age <strong>of</strong> trees, andan inner thin living part (just near the cambiumzone), which carries food from the leaves to thegrowing elements<strong>36</strong>.1.1 .1 Knot sAs the tree grows in height (primary growth), branchingis initiated by lateral bud development. Knots arethe bases <strong>of</strong> branches, which have been covered as thetree grows laterally. After a branch dies, the trunkcontinues to increase in diameter and surrounds thatportion <strong>of</strong> the branch while the dead branch is stillpresent. This branch has to drop from the tree beforeclearwood can form. If the knot was alive when thetrunk grew around it, the xylem <strong>of</strong> the trunk and thebranch are continuous and the knot fits tightly intothe wood. If the branch was dead when the trunkgrew around it, no anatomical connection exists betweenthe xylem <strong>of</strong> the knot and the trunk. The knotis nonadhesive; it may fall out <strong>of</strong> the wood, leaving aknothole (Figure <strong>36</strong>.3).<strong>36</strong>.1.1 .2 Tiss ue and Cellu lar Structure <strong>of</strong> Woo dGrowth in thickness <strong>of</strong> the bark and wood is causedby cell division in the cambium. New wood cells areformed on the inside <strong>of</strong> the cambium and new barkcells on the outside. In the cambium region, immaturecells differentiate into various kinds <strong>of</strong> mature xylem(wood) and phloem (inner bark) cells characteristic <strong>of</strong>the species (Panshin and de Zeeuw, 1980). Then enlargement,elongation, and maturation allow thewoody material to be developed (Figure <strong>36</strong>.4). Most<strong>of</strong> the wood cells stay alive for not more than fewß 2006 by Taylor & Francis Group, LLC.


Primary growthSecondannual ringFirstannual ringPith1 mmThickening <strong>of</strong> the structure by secondary growthFIGURE <strong>36</strong>.1 Formation <strong>of</strong> wood in trees: the pith originates from the primary growth whereas the wood material is added,along the years, by the secondary growth. (Microphotograph: polished disc <strong>of</strong> Douglas-fir (Pseudotsuga menziesii), LER-MAB–ENGREF.)Age <strong>of</strong> wood cellsPithAge <strong>of</strong> treeBarkHeartwoodformationCambial activitydivision,elongation,differentiation,lignificationLTRFIGURE <strong>36</strong>.2 Cross section <strong>of</strong> a tree showing the internal structure <strong>of</strong> the stem. Growth rings can also be observed: light partsare earlywood and dark parts are latewood. Due to this stem geometry, three material directions: longitudinal (L), radial (R),and tangential (T), can be defined at each location. (Photograph: Yew (Taxus baccata L.), LERMAB–ENGREF.)ß 2006 by Taylor & Francis Group, LLC.


Clearwood formedafter pruningDeath <strong>of</strong> thebranchNonadhesive part<strong>of</strong> the knotAdhesive part <strong>of</strong>the knotPithFIGURE <strong>36</strong>.3 The existence <strong>of</strong> knots in wood comes from the interaction between primary growth, responsible for thebranch formation, and secondary growth, responsible for thickening <strong>of</strong> stem and branches. (Photograph: Scots pine (Pinussylvestris), LERMAB–ENGREF.)weeks; the last development stage, namely lignification,induces the death <strong>of</strong> these cells. Only parenchymacells can live for years and they are responsible for thedevelopment <strong>of</strong> heartwood.In most species in temperate climates, the differencebetween woods that are formed early in a growing season(earlywood) and that formed later (latewood) issufficient to produce concentric contours in a crosssection (Figure <strong>36</strong>.2). These rings are known as growthrings. Each increment in size in the branch or trunkdiameter can be observed in these growth rings thatremain unchanged once formed. Provided no falserings exist (due to an interruption <strong>of</strong> the growth indiameter by drought or defoliation by insects), the ageat any cross section <strong>of</strong> the trunk may be determined bycounting these rings. Obviously, this simple rule doesnot apply for tropical species for which growth may bepractically continuous throughout the year and no welldefinedgrowth rings are formed, or for which growthringsare the result <strong>of</strong> the individual rhythm<strong>of</strong>eachtree.<strong>Wood</strong> cells are <strong>of</strong> various sizes and shapes. Theyare cemented together to form the structural woodmaterial. The majority <strong>of</strong> wood cells are elongatedand pointed at the ends. The types and dimensions<strong>of</strong> wood cells depend strongly on the species.In s<strong>of</strong>twoods, woods formed by cone-bearing trees(e.g., fir, pine, and spruce) with naked seeds, the xylemcontains mainly tracheids (90%). Tracheids are considerablyelongated cells (around 40 mm in diameterand between 2 and 8 mm in length), which ensure bothsap flow, by means <strong>of</strong> numerous bordered pits situatedon the radial cell walls, and mechanical strength.In s<strong>of</strong>twoods, the earlywood is characterized by cellswith large radial diameters and thin walls, and hencerelatively large cavities. Latewood cells have a muchsmaller radial diameter and thicker walls, which resultin much smaller cavities (Figure <strong>36</strong>.4). In addition,some s<strong>of</strong>twoods have resin canals. Parenchyma cellssurround these canals and actively secret resin into thecanals, and ultimately into the heartwood.Hardwood is the common name for the wood <strong>of</strong>species whose seeds are enclosed in ovaries. Thesespecies are more advanced than s<strong>of</strong>twoods in terms<strong>of</strong> biological evolution; consequently, they produce amore sophisticated anatomical pattern, with cellsmuch more adapted to meet specific requirements inrelation to water transport, food storage, and mechanicalsupport:. Fibersareusuallyrelativelythick-walled,sparselypitted, and about 1 mm in length. Different fibercells exist, but the tracheid fibers having borderedpits are generally the most abundant. Althoughfibers may have a certain role in sap conduction,theybasicallyfunctionasthemechanicalsupport,making the wood usually stronger, denser, andmore durable than s<strong>of</strong>twoods.. Vessels <strong>of</strong> relatively large diameter are alsoknown as pores. These cells form the main conduitsfor sap flow. A vessel is built up by severalvessel elements, with more or less open endplates, aligned in the longitudinal direction.ß 2006 by Taylor & Francis Group, LLC.


FIGURE <strong>36</strong>.4 <strong>Wood</strong> is formed by cell division in the cambium zone, hence the radial cell lines appear clearly in this figure, inspite <strong>of</strong> the large variation <strong>of</strong> radial diameter between earlywood and latewood. Some bordered pits allowing sap flow fromone tracheid to the other can also be observed on radial cell walls. (ESEM Photograph: Norway spruce (Picea abies),LERMAB–ENGREF.). Longitudinal or axial parenchyma cells functionmainly in the storage <strong>of</strong> food.Several thousands <strong>of</strong> hardwood species exist, andeach one has its own anatomical pattern. The density,for example, ranges from less than 100kg m 3 (i.e.,balsa wood) up to more than 1200kg m 3 (i.e., ebonywood). They are usually divided into ring-porous anddiffuse-porous types, though all intermediate typescan be found:. A ring-porous species produces very large vessels(up to 500 mm in diameter) in earlywood.Simple perforation plates allow the vessel cellsto communicate easily.. Diffuse-porous species have smaller vessels (50to 100 mm in diameter) <strong>of</strong> almost uniform sizeand distribution throughout the growth ring. Indiffuse-porous species, vessel cells are usuallyconnected by scalariform (‘‘ladderlike’’) perforationplates.Figure <strong>36</strong>.5 depicts the typical anatomical patternsencountered in temperate species.Both hardwoods and s<strong>of</strong>twoods have cells (usuallygrouped into structures or tissues) that areoriented horizontally in the radial direction andwhich are called rays. The rays, composed <strong>of</strong> parenchymawith lignified cell walls or sclerenchyma, connectvarious layers from pith to bark for storage andtransfer <strong>of</strong> food. In s<strong>of</strong>twoods, rays are one-cell thick.In hardwoods, they vary in size from one-cell wideand a few cells high to more than 15-cell wide andseveral centimeters high. Rays represent planes <strong>of</strong>weakness along which drying checks develop easily.<strong>36</strong>.1.1 .3 Hear twoo d F ormationJust few weeks after the formation <strong>of</strong> xylem, it containsmostlydeadcells,butcontinuestoplayarole<strong>of</strong>utmostimportance for the plant: the transport <strong>of</strong> sap. Hencethevascularsystemsoproducediscalledsapwood.Thisactive zone may vary in thickness and number <strong>of</strong>growth rings, commonly up to about 15 years, whichrepresents several centimeters in radial thickness. As arule, the more vigorously growing trees have more extensivesapwood. In sapwood, parenchyma cells stayalive and function primarily in the storage <strong>of</strong> food.ß 2006 by Taylor & Francis Group, LLC.


(a) Norway spruce (Picea abies) (b) Pedunculate oak (Quercus rubra) (c) European beech (Fagus sylvatica)FIGURE <strong>36</strong>.5 Typical anatomical patterns encountered in temperate species: (a) s<strong>of</strong>twood, (b) ring-porous species, and (c)diffuse-porous species. The height <strong>of</strong> these images represents about 2 mm. (Microphotographs: J.C. Mosnier, LERMAB–ENGREF.)After some years, an intense biological activity <strong>of</strong>these parenchyma cells gives rise to heartwood formation.Metabolites are deposited in the heartwoodand the tree uses the heartwood as a place to storewaste products that are collectively known as extractives.These include resins, gums, oils, and tanninsthat stop up the vessels (or the tracheids in s<strong>of</strong>twood)and clog the wood. In heartwood, all the cells aredead and inactive; they do not function in eitherwater conduction or food storage. The heartwood is<strong>of</strong>ten darker, slightly denser, and more durable (resistanceto fungi or insect attack) than the sapwoodand plays an important role in supporting the tree.However, numerous species do not have dark heartwood(e.g., spruce, fir, and beech) and no correlationexists between heartwood color and durability.In some species, such as certain oaks, the vesselsbecome plugged with the development <strong>of</strong> cellularmembranes known as tyloses, which enter the vesselsfrom adjacent parenchyma cells (Figure <strong>36</strong>.6).<strong>36</strong>.1.2 CHEMICAL C OMPOSITION<strong>Wood</strong> is a typical organic material made up <strong>of</strong> threemain elements: carbon, oxygen, and hydrogen. Becausevery small variations between different wood speciesare observed, the numbers depicted in Table <strong>36</strong>.1 arebroadly general but can be higher for some tropicalspecies. Nitrogen as well as some additional inorganicelements (sodium, potassium, calcium, magnesium, andsilicon) are also essential compounds, which are mostlyinvolved in the metabolism <strong>of</strong> living cells during woodFibersRay cellsTyloses invesselsRTTLFIGURE <strong>36</strong>.6 Example <strong>of</strong> tyloses development in the heartwood <strong>of</strong> Pedunculate oak (Quercus rubra L.). (ESEM photographs:Patrick Perré and Riad Bakour, LERMAB–ENGREF.)ß 2006 by Taylor & Francis Group, LLC.


TABLE <strong>36</strong>.1Elementary Composition <strong>of</strong> <strong>Wood</strong>Element Content (%)Carbon 49–50Hydrogen 6Oxygen 43–44Nitrogen


s<strong>of</strong>twoods, compression wood is found in the lowerpart. However, the compression wood can also befound in some primitive groups <strong>of</strong> hardwoods (Carlquist,2001).<strong>36</strong>.1.3 .1 Com pression Wo odCompared with normal wood, this tissue is characterizedby shorter tracheids, higher lignin and hemicellulosecontent, and lower cellulose content. Thisreaction wood is easily identified on smooth surfaces,in particular in a transverse view. When compressionwood is formed, the growth rings appear darker,reddish brown, and <strong>of</strong>ten wider than on the oppositeside. Therefore, when compression wood develops inthe same side for several years, the cross section <strong>of</strong> thestem tends to be oval with an eccentric pith in thecore; this is typical <strong>of</strong> branches or stem <strong>of</strong> bent trees.At the anatomical level the cells observed in crosssection are more rounded than rectangular, showinglarge intercellular spaces (Figure <strong>36</strong>.7). The cell wallconsists only <strong>of</strong> ML, P, S 1 , and S 2 layers. Once dry,the cell wall shows deep, helically arranged checksfrom the lumen. The latter is rather thick and itsmicr<strong>of</strong>ibril angle is much larger than in normalwood (about 45 8 from the axial direction). In consequence,the density is higher, the longitudinal shrinkageis increased to some percentage (compared with0.1 to 0.2% for normal wood), and, in spite <strong>of</strong> thehigher density, the longitudinal mechanical propertiesare less than in normal wood. Due to the large micr<strong>of</strong>ibrilangle in the S 2 layer, the effect is opposite in thetransverse plane: lower shrinkage and higher stiffness.<strong>36</strong>.1.3 .2 Te nsion Wo odTension wood is characterized by increased cellulosecontent and increased density. Sawn surfaces appearwoolly and rough. Strength properties are reduced.Longitudinal shrinkage can be more than 1%. Nosignificant coloration marks out tension-wood zones.In addition, the regulation <strong>of</strong> hardwood seems to besubtler than in s<strong>of</strong>twoods, sometimes leading to verylocalized presence <strong>of</strong> reaction wood (thin layers in theradial direction and variable angular position fromone year to the other). Therefore, the macroscopicfeatures are not very reliable.At the microscopic level, tension wood is mucheasier to identify when it is fully developed. Fiber cellwalls are much thicker than normal, enclosing verysmall lumens. Secondary walls are loosely attached tothe primary wall and thus are responsible for some <strong>of</strong>the differing mechanical properties. The thick secondarywall <strong>of</strong> tension-wood fibers is significantly lesserlignified; it consists <strong>of</strong> almost pure cellulose. Due tothis consistency, this layer is termed gelatinous orG layer (Figure <strong>36</strong>.8). The micr<strong>of</strong>ibrils are almostparallel to the grain. The reason for tension-woodshrinkage is not well understood. The loose contactbetween the G layer and the remaining cell wall,which does not restrain the outer cell region (i.e.,primary layer) and from contraction during drying,is one possible explanation. However, a recent worktends to prove that the G layer itself has a highlongitudinal shrinkage, which is not consistent withthe small micr<strong>of</strong>ibril angle (Clair, 2001). It may benoted that intermediate indications <strong>of</strong> tension woodexist without the G layer.<strong>36</strong>.1.3.3 Juvenile <strong>Wood</strong>All trees during their growth produce juvenile wood,i.e., the inner core <strong>of</strong> xylem surrounding the pith.The time during which juvenile wood is formed variesamong individuals, with species, and with environmentalconditions. It is <strong>of</strong>ten recognized that juvenileRTNormal woodCompression woodFIGURE <strong>36</strong>.7 Compression wood compared with normal wood. (ESEM photographs: White fir (Abies alba), LERMAB–ENGREF.)ß 2006 by Taylor & Francis Group, LLC.


FIGURE <strong>36</strong>.8 Tension wood in Pedunculate oak (Quercus rubra); note the existence <strong>of</strong> the G layer. (ESEM photograph:LERMAB–ENGREF.)wood formation lasts as long as living branches existat the corresponding height <strong>of</strong> the tree. The transitionfrom juvenile to mature wood takes place gradually.No clear demarcation exists between juvenile andadult wood.In juvenile wood the cells are smaller than those <strong>of</strong>the mature xylem. Particular differences exist in thelength <strong>of</strong> the cells as well as in the structure <strong>of</strong> thelayered cell wall. The micr<strong>of</strong>ibril angle in the S 2 layeris greater than in cells <strong>of</strong> the mature tissue. As forcompression wood, this causes a higher value <strong>of</strong> longitudinalshrinkage and a reduced tensile strength. Inaddition, the spiral grain (angle between the stem axisand the fiber orientation) is <strong>of</strong>ten large in the juvenilewood. Together with the high longitudinal shrinkage,this explains why the warping <strong>of</strong> timber containingjuvenile wood may be dramatic after drying. Juvenilewood is a major problem in processing wood fromplantations <strong>of</strong> fast-growing species (eucalyptus,radiata pine, etc.) that produce logs made up mostly<strong>of</strong> juvenile wood.<strong>36</strong>.1.4 IMPLICATIONS FOR THE DRYING P ROCESSToinducetheascent<strong>of</strong>sapintree,themeniscipresentinthe leaf stomata pull up water (Zimmerman 1983).Because most trees are more than 10 m high, one candeduce that the absolute liquid pressure in the sapcolumn is negative. No gaseous phase can exist in suchconditions. The vascular system developed in trees hasmany other implications for the drying process:. Because the system is designed for longitudinalsap flow from the roots to the canopy, the woodmaterial is strongly anisotropic.. Because <strong>of</strong> negative pressure, the vascular systemmust be able to support a gas invasion dueto injury or cavitation. This is the role <strong>of</strong> borderedpits or vessel-to-vessel pits. These anatomicalfeatures may dramatically inhibit thefluid migration in the wood.. In heartwood, due to metabolite deposition, aspirationor closure <strong>of</strong> bordered pits, or tylosedevelopment, the permeability is <strong>of</strong>ten reducedby one or several orders <strong>of</strong> magnitude.. The wood is fully saturated in the sapwood part<strong>of</strong> logs (an air-free sap column is required toobtain negative pressures), whereas the heartwoodzone is generally only partly saturated.Table <strong>36</strong>.3 indicates some orders <strong>of</strong> magnitudegenerally observed for the moisture content <strong>of</strong> greenwood.Indeed, because the sapwood part is fully saturated,the maximum moisture content in this zonecan be calculated by assuming that the entire porevolume is filled with water:X ¼ fr ‘(1 f)r swith f ¼ 1r 0r s(<strong>36</strong>:1)where f is the porosity, r 0 is the basic density (ovendrymass/green volume), r s is the density <strong>of</strong> the cellß 2006 by Taylor & Francis Group, LLC.


TABLE <strong>36</strong>.3Typical Values <strong>of</strong> Moisture Content Found inGreenwood<strong>Wood</strong> Type Moisture Content Dry Basis (%)wall substance (r s ffi 1530 kg m 3 ), and r ‘ is the sapdensity (r ‘ ¼ 1000 kg m 3 ).To highlight that the development <strong>of</strong> wood intrees leads to a very anisotropic material, Table <strong>36</strong>.4indicates some order <strong>of</strong> magnitude for dimensionlessanisotropy ratios found in wood for the most importantproperties involved in drying. The ease <strong>of</strong> fluidmigration in wood (i.e., the permeability) is by far theproperty that presents the highest anisotropy ratio.The reduction in wood permeability from sapwood toheartwood affects particularly the longitudinal directionin hardwoods (especially for ring-porous speciesdeveloping tyloses) and all directions in s<strong>of</strong>twoods.<strong>36</strong>.2 BOARD SCALE<strong>36</strong>.2.1 WATER IN WOOD: SORBED AND CAPILLARYWATER, AND SHRINKAGE<strong>36</strong>.2.1.1 Moisture Content <strong>of</strong> <strong>Wood</strong>The moisture content <strong>of</strong> wood (dry basis) is definedby the following mass ratio:X ¼Sapwoodmass <strong>of</strong> wateroven-drymassHeartwoodS<strong>of</strong>twoods 150–200 40–80Hardwoods 80–120 60–100(<strong>36</strong>:2)Moisture in wood is distinguished between the liquidsap present unattached in the cell lumens (free water)and the water molecules held in the cell walls (boundwater), with an activity less than 1.<strong>36</strong>.2.1.2 Free WaterThe sap flows in wood through the vascular systemproduced by the cambium. This liquid, present in thecell cavities, is usually referred to as ‘‘free water’’ justbecause its properties are very close to those <strong>of</strong> liquidwater: density, viscosity, saturated vapor pressure, etc.However,onehastokeepinmindthatthiswateristied to the solid matrix through capillary forces. Due tothe surface energy <strong>of</strong> the interface between liquid andgas (the surface tension), together with the contact anglebetween this interface and the woody matrix, a pressuredifference exists between liquid and gaseous phases.This pressure difference, which obeys Laplace’s law,increases with decreasing pore diameter. Because liquidis the wetting phase in the case <strong>of</strong> water and wood, theliquid pressure is less than the gaseous one.In addition, due to the curvature <strong>of</strong> the interface,a deviation <strong>of</strong> the saturated vapor pressure exists.This deviation can be calculated from the definitionand properties <strong>of</strong> the Gibbs free energy and leads toKelvin equation (Dullien, 1992):ln w ¼ lnP v¼ s 1 þ 1 Mv(<strong>36</strong>:3)P vs r 1 r 2 r ‘ RTIn Equation <strong>36</strong>.3, P vs is the saturated vapor pressure, P vis the equilibrium vapor pressure at the curved interface,s is the surface tension, r 1 and r 2 are the twoprincipal radii <strong>of</strong> the surface, and M v is the molarmass <strong>of</strong> water. The quantity w is known as the relativehumidity. However, a very small pore radius is requiredfor the deviation to become significant (Table <strong>36</strong>.5).TABLE <strong>36</strong>.5TABLE <strong>36</strong>.4Order <strong>of</strong> Magnitude <strong>of</strong> the DimensionlessAnisotropy Ratios Encountered in<strong>Wood</strong> Relative to Tangential ValuePropertyDirectionRadius <strong>of</strong> theCapillaryT R LStiffness 1 2 20Shrinkage 1 0.5 0Thermal conductivity 1 1.5 2Mass diffusivity 1 1–2 20Pressure Difference DP and Relative Humidity wat the Surface for Different Radii Values (ValuesCalculated at 208C, for a Perfectly Wetting Liquidand Cylindrical Tubes)DP (Equivalent WaterColumn)w ¼ P v /P vs1 mm 0.146 kPa (14.9 mm) 0.999999100 mm 1.46 kPa (0.149 m) 0.99998910 mm 14.6 kPa (1.49 m) 0.999891 mm 0.146 MPa (14.9 m) 0.99890.1 mm 1.46 MPa (149 m) 0.9890.01 mm 14.6 MPa (1.49 km) 0.898ß 2006 by Taylor & Francis Group, LLC.


<strong>36</strong>.2.1.3 Bound WaterBound moisture is associated with the hygroscopicnature <strong>of</strong> the woody components. There are someuncertainties about the limits <strong>of</strong> hygroscopic behavior,particularly with woods <strong>of</strong> high extractives content;but it is useful to define a maximum sorptivemoisture content, called the fiber saturation point(FSP). If the capillary condensation effects in poresgreater than 0.1 mm in equivalent cylindrical diameterare ignored, FSP <strong>of</strong> the wood may be defined as theequilibrium moisture content (EMC) in an environment<strong>of</strong> 99% relative humidity. This yields a value <strong>of</strong>30 to 32% for most commercial species (Keey et al.,2000) at room temperature. FSP falls with increasingtemperature. For a s<strong>of</strong>twood such as Sitka spruce(Picea sitchensis), FSP falls from about 31% at 258Cto 23% at 1008C (Stamm, 1964).As the relative proportions <strong>of</strong> the woody componentsvary only within narrow ranges for commoncommercial species, the EMCs at a given relativehumidity and temperature are closely similar forthese woods. However, at high relative humiditiesdeviations from mean values can appear. Shubin’sdata (1990) show, for instance, that at 95% relativehumidity the EMC at 42.48C ranges from 22% for apine to 33% for an oak. Hoadley (1980) notes that inspecies with a high extractives content, such as redwood(Sequoia sempervirens) and mahogany (Swieteniamahogani), the fibers remain saturated at 22 to24% moisture content, whereas birch (Betula spp.)may have a moisture content up to 35% at fibersaturation.Because <strong>of</strong> sorption hysteresis, the EMC at agiven relative humidity is higher in drying (desorption)from greenwood than in wetting (adsorption)from perfectly dried wood. The ratio <strong>of</strong> values normallyranges between 0.75 and 0.88 (Schniewind,1989). However, in previously dried wood subject toenvironmental swings in relative humidity, the isothermfor adsorption and desorption becomes moresimilar. Thus, for practical proposes, the hygroscopicity<strong>of</strong> wood is sufficiently similar for most commercialspecies for generalized sorption data to be useful.The curves depicted in Figure <strong>36</strong>.9 represent an averagebetween desorption and adsorption. They wereplotted using the mathematical correlation used in thenumerical code ‘‘Transpore.’’ The parameters <strong>of</strong> thisexpression have been fitted from various data availablein the literature: those published by Rasmussenin 1961 (Siau, 1984) and those <strong>of</strong> Loughborough onSitka spruce published by Hawley in 1931 and arrangedby Keylwerth in 1949 (Kollmann and Côté,1968; Joly and More-Chevalier, 1980).<strong>36</strong>.2.1.4 Differential Heat <strong>of</strong> SorptionSorbed water in the cell wall has a lower enthalpythan liquid water. However, contrary to other forms<strong>of</strong> water, such as solid, the enthalpy <strong>of</strong> bound waterincreases with increasing moisture content up to FSP.Above this value, the enthalpy <strong>of</strong> water in wood isessentially the same as that <strong>of</strong> liquid water.The value <strong>of</strong> the differential heat <strong>of</strong> sorption canbe calculated from the sorption isotherms using theClausius–Clapeyron equation (Skaar, 1988):35Equilibrium moisture content (%)3025201510520C60C100C00 20 40 60 80 100Relative humidity (%)FIGURE <strong>36</strong>.9 Sorption isotherms calculated by a mathematical expression fitted from published data.ß 2006 by Taylor & Francis Group, LLC.


Pure waterWater in woodVaporL vΔh sL v−L fFiber saturation pointIceLiquidBound waterMoisture content≈ −L v /2FIGURE <strong>36</strong>.10 Differential heat <strong>of</strong> sorption versus the moisture content. (Adapted from Skaar, C., <strong>Wood</strong>–Water Relations,Springer, Berlin, 1988.)Dh s ¼d ln P vR P vsM vd1 (<strong>36</strong>:4)TDimensionLThis heat <strong>of</strong> sorption is slightly higher than 1000 kJkg 1 for oven-dry wood and decreases rapidly withmoisture content to reach zero, with a horizontalslope at the FSP (Figure <strong>36</strong>.10).RT<strong>36</strong>.2.1 .5 Sh rinkageIn trees, the cell walls <strong>of</strong> wood are in a fully swollencondition. However, when sorbed water is removedfrom the cell wall, new hydrogen bonds form betweenthe hydroxyl groups <strong>of</strong> the molecular chains andreduce the distance between the chains. This variationaffects the dimension in a direction normal to themicr<strong>of</strong>ibril direction. This fact explains why the longitudinalshrinkage is usually negligible. In the transversedirections, the dimensional variations plottedagainst bound water content are very close to astraight line (Figure <strong>36</strong>.11). Shrinkage can then bedefined by the total shrinkage (variation <strong>of</strong> dimensionbetween the green state and the oven-dry condition)or in terms <strong>of</strong> a shrinkage coefficient (variation <strong>of</strong>dimension divided by the variation in moisture content).For most species, the tangential shrinkage isabout twice the radial shrinkage. Several featurescan explain this observation: the cell arrangement <strong>of</strong>tissues, the difference between earlywood and latewood,and the presence <strong>of</strong> ray cells. The extrapolation<strong>of</strong> length values against moisture content to the originalunshrunken length yields the so-called shrinkageintersection point (SIP) (Figure <strong>36</strong>.11). Normally, thismoisture content is a little higher than the FSP, but<strong>of</strong>ten the two moisture content values are assumed tobe same.FSP SIPMoisture contentFIGURE <strong>36</strong>.11 Dimensional variations <strong>of</strong> a typical woodsample vs. moisture content.The shrinkage values tend to increase with basicdensity, but values vary strongly between and withinspecies. Table <strong>36</strong>.6 indicates the order <strong>of</strong> magnitudevalues encountered in some species.In the case <strong>of</strong> compression wood, the micr<strong>of</strong>ibrilangle in the S 2 layer is large, which results in a lowertransverse shrinkage and a significant longitudinalshrinkage. In spite <strong>of</strong> the very low value <strong>of</strong> themicr<strong>of</strong>ibril angle in the G layer, tension wood <strong>of</strong>hardwood has also an abnormally high longitudinalshrinkage.As logs are sawn in green state, shrinkage occursafter sawing. The size <strong>of</strong> each section is reduced byshrinkage and also the shape <strong>of</strong> the section changescaused by the tangential or radial anisotropy inshrinkage (Figure <strong>36</strong>.12):a. Only the quartersawn sections without the pithkeep their rectangular shapeb. Flatsawn sections cup into a trough-like shapeß 2006 by Taylor & Francis Group, LLC.


TABLE <strong>36</strong>.6Order <strong>of</strong> Magnitude <strong>of</strong> Shrinkage Values in theTransverse Plane for Some Species. FSP is Assumedto be Equal to 30% for All SpeciesSpecies Radial Shrinkage Tangential ShrinkageTotal(%)Coefficient(%/%)Total(%)Coefficient(%/%)Spruce, pine 5.0 0.17 9.0 0.30Oak 6.0 0.20 11.0 0.37Beech 6.5 0.22 12.0 0.40Balsa 3.0 0.10 6.0 0.20Teak 3.0 0.10 5.0 0.17Okoumé 4.0 0.13 6.5 0.22Azobé 8.0 0.27 11.0 0.37Source: Kollmann, F.P. and Côté, W.A., Principles <strong>of</strong> <strong>Wood</strong>Science and Technology, Solid <strong>Wood</strong>, Vol. 1, Springer, Berlin,1968; Aléon, D., Chanrion, P., Négrié, G., and Perré, P.,FormaXylos 4—Le séchage (Training in <strong>Wood</strong> Science: <strong>Drying</strong>,Vol. 4), CD-Rom français/English, CTBA, Paris, 2003.c. Quartersawn sections, which include the pith,present a nonuniform thickness once dried;near the pith, the thickness shrinks along theradial direction whereas the tangential shrinkageis involved elsewhered. A squared section cut along the grain directionbecomes rectangular after dryinge. A diamond shape is obtained if the grain directionis along the diagonal<strong>36</strong>.2.2 HEAT AND MASS T RANSFER IN WOOD<strong>36</strong>.2.2 .1 Flui d Migration in <strong>Wood</strong>:Singl e-Phase FlowThe understanding <strong>of</strong> the fluid migration in wood is<strong>of</strong> utmost importance to understand the drying process.The differences in drying behavior between speciesand within the log primarily come from thepermeability value and from the initial moisture saturation.When only one fluid phase is present, theclassical Darcy’s law applies:v ¼ K r(P) (<strong>36</strong>:5)mwhere v is the apparent velocity <strong>of</strong> the fluid throughthe specimen (m s 1 ), K is the permeability (m 2 ), andm is the dynamic viscosity <strong>of</strong> the fluid (Pa s 1 ).Fluid migration in wood uses the vascular systemdeveloped by trees for their physiological requirements.For this reason, wood has several specificfeatures concerning permeability among which themost important are the anisotropy ratios. <strong>Wood</strong> hasdramatic anisotropy ratios: the longitudinal permeabilitycan be 1000 times greater than the transversepermeability for s<strong>of</strong>twoods and more than a millionfoldfor hardwoods (Banks, 1968).Table <strong>36</strong>.7 summarizes some values <strong>of</strong> directionalpermeability available in the literature for differentspecies. Depending on the experimental apparatusand the protocol used by the authors, some data aremissing in the papers. For example, it is not alwaysFlatsawn(d)(b)(e)Quartersawn(a)(c)FIGURE <strong>36</strong>.12 Section deformations depending on the sawing pattern. The shape after drying results from the anisotropyratio between radial and tangential shrinkage. These deformations exist even when the equilibrium is achieved and with auniform moisture content throughout the section. (Adapted from Aléon, D., Chanrion, P., Négrié, G., and Perré, P.,FormaXylos 4—Le séchage (Training in <strong>Wood</strong> Science: <strong>Drying</strong>, Vol. 4), CD-Rom français/English, CTBA, Paris, 2003.)ß 2006 by Taylor & Francis Group, LLC.


TABLE <strong>36</strong>.7Order <strong>of</strong> Magnitude <strong>of</strong> Permeability for Some Species along the Three Material Directions <strong>of</strong> <strong>Wood</strong>References Species Notes Permeability (m 2 ) Anisotropy RatioL · 10 12 R · 10 16 T · 10 16 K L /K R K L /K T K R /K TChoong and Kimbler, 1971 Populus sp. S. l. a 0.59 — 0.44 — 13,000 —Populus sp. H. l. a 0.61 0.15 0.18 40,000 33,000 0.835Alnus rubra S. l. a 1.9 0.15 0.12 125,000 166,000 1.327Liquidambar sp. S. l. a 2.7 0.19 0.69 145,000 39,000 0.273Liriodendron sp. H. l. a 5.4 — 0.80 — 67,000 —Sequoia sp. S. l. a 4.9 11.2 19.4 4,000 2,000 0.580Sequoia sp. H. l. a 0.25 0.112 0.88 23,000 3,000 0.127Pseudotsuga sp. H. l. a 0.026 0.000 0.000 — — —Pseudotsuga sp. S. l. a 0.049 — 0.37 — 1,300 —Pinus sp. H. l. a 0.0026 0.17 — 153 —Choong. et al., 1974 Acer rubrum S. g. o 10.3 — — 8,300 11,400 1.4Acer rubrum H. g. o 7.4 — —Liriodendron sp. S. g. o 28.9 — — 1,450 1,600 1.1Liriodendron sp. H. g. o 1.87 — —Liquidambar sp. S. g.o 13.9 — — 1,300 2,750 2.1Liquidambar sp. H. g. o 15.3 — —Quercus rubra S g 62.0 — — 13,000 18,000 1.4Quercus rubra H g 56.0 — —Quercus falcata S g 69.0 — — 110,000 143,000 1.3Quercus falcata H g 13.0 — —Chen et al., 1998 Liriodendron sp. S l 26.0 4.5 — 57,000 — —Liriodendron sp. H l 0.1 — — — — —Juglans nigra S l 27.0 0.08 — 3 10 6 — —Juglans nigra H l 0.0052 — — — — —Quercus rubra S l 61.0 0.68 — 900,000 — —Quercus rubra H l 45.0 — — — — —Perré, 1992 and 2002 Picea sp. S. g. a 0.2 — — 700 — —Fagus sylvatica S g. a 3.8 — — 3,000 65,000 21.0Fagus sylvatica H g. a 1.4 — — 3,000 — —Populus sp. H g. s 0.03 — — 10,000 — —L, longitudinal; R, radial; T, tangential; S, sapwood; H, heartwood; l, permeability to liquid; g, permeability to gas; a, air-dried sample; o,oven-dried sample.easy to calculate the permeability ratios from permeabilityvalue, or vice versa. Choong et al. (1974), forexample, have reported the permeability values forsapwood and heartwood for the longitudinal direction,but not for the transverse directions. Only the meananisotropy ratio is available in this paper. Perré (1992)and Perré et al. (2002) have used an experimentalprocedure to determine the longitudinal permeabilityand the anisotropy ratio on the same sample. In theseinstances, they just obtained the ratios and decided notto calculate the transverse permeability accordingly.The variability is impressive for both the permeabilityvalues and the anisotropy ratios. In spite <strong>of</strong> thescatter in the data, general trends are exhibited for thelongitudinal permeability:. The most permeable species in the longitudinaldirection are among the ring-porous species(Quercus spp.) that have very large vessels (upto 500 mm in diameter).. The diffuse-porous hardwoods are fairly permeabletoo; probably, the large number <strong>of</strong>vessels can <strong>of</strong>fset their smaller diameter (around50 mm).. S<strong>of</strong>twood species are generally less permeable;these trees have no specific elements for sapflow, so the fluid has to pass through the smallopenings, the bordered pits, at the end <strong>of</strong> eachtracheid along the path, which is 1 to 2 mmeach. Certain s<strong>of</strong>twood species, such as Pinusradiata, however, are very permeable.ß 2006 by Taylor & Francis Group, LLC.


The transverse permeability and the anisotropyratios are very variable. The heartwood part <strong>of</strong> logsis usually much less permeable than the sapwood part(Comstock, 1967). This is due to tyloses developmentand extractives deposition (tannins, gums, etc.) inhardwoods and due to the aspiration <strong>of</strong> borderedpits in s<strong>of</strong>twoods.<strong>36</strong>.2.2.1.1 Pit AspirationIn s<strong>of</strong>twoods, the tracheids have considerable pittingin their radial walls. These bordered pits are specializedvalves to seal and isolate tracheids if they becomedamaged or embolized, but remain open for sap flow.At the pit location, the double cell wall takes theshape <strong>of</strong> the external part <strong>of</strong> a torus. The externaldiameter <strong>of</strong> this torus is in the range 10 to 20 mm,depending on the position in the annual growth ring(the diameter is smaller in latewood). The torus issuspended by the margo, a net <strong>of</strong> radially orientedmicr<strong>of</strong>ibrils, with openings up to some micrometerswide. In normal operation, the sap flows from onetracheid to the other simply by using these smallopenings (Figure <strong>36</strong>.13a).When gas invades one tracheid, the gas–liquidinterface is blocked in the margo due to capillaryforces. These forces press the torus against the oppositeborder <strong>of</strong> the pit (Figure <strong>36</strong>.13b). Then, hydrogenbonds keep the torus in this position; the pit is nowimpermeable (Figure <strong>36</strong>.13c). Such sealing <strong>of</strong> a pit isknown as aspiration.This subtle mechanism is vital for trees, but causessome difficulties in wood drying. Indeed, pit aspirationoccurs as soon as water is removed from the wood,sometimes even during the heartwood formation. Inparticular, it is impossible to avoid pit aspiration whendrying s<strong>of</strong>twoods under normal conditions.Only freeze-drying, or changing the liquid phase fora solvent with low surface tension before drying, allowsair-dried samples without pit aspiration to be obtained(Comstock and Côté, 1968; Meyer, 1971; Bolton andPetty, 1978; Fumoto et al., 1984). The permeabilityvalues depend on the species and on the author,but all these data show that air-dried samples, withaspirated pits, have permeability values considerablysmaller than unaspirated samples (typically rangingfrom 1 to 10%). Due to thicker cell walls, smaller pitradii, and more rigid structures, the percentage <strong>of</strong> aspiratedpits is much less in the latewood part <strong>of</strong> samples(Siau, 1984).<strong>36</strong>.2.2.2 Generalized Darcy’s Law:Multiphase FlowWhen two phases coexist, the generalized Darcy’s lawmust be used. The volumetric flow rate <strong>of</strong> each phaseis considered to be proportional to the pressure gradient<strong>of</strong> the corresponding phase. The phenomenologicalcoefficient is the product <strong>of</strong> the permeabilityK by a function <strong>of</strong> saturation called ‘‘relative permeability’’to the considered phase.Middle lamellaCapillary forcesMargo≅10 µmTorusGasLiquidPosition <strong>of</strong> theair/liquidmeniscusGasGasLiquidLiquidSecondary wallConductive bordered pitPit aspirationAspirated pitFIGURE <strong>36</strong>.13 The mechanism <strong>of</strong> pit aspiration: a clever strategy to limit the damage caused by any gas invasion due toinjury or cavitation <strong>of</strong> the sap column. (Adapted from Siau, J.F., Transport Processes in <strong>Wood</strong>, Springer, Berlin, 1984.)ß 2006 by Taylor & Francis Group, LLC.


For the gaseous phasev g ¼For the liquid phasev ‘ ¼Kk rg(S)m grP g (<strong>36</strong> :6)Kk r‘(S)m ‘rP ‘ (<strong>36</strong> :7)The liquid pressure is related to the gaseous pressurethrough the capillary pressure function:P ‘ ¼ P g P c (S) (<strong>36</strong> :8)Equation <strong>36</strong>.6 and Equation <strong>36</strong>.7 must be consistentwith Darcy’s law (Equation <strong>36</strong>.5) when one singlefluidphase occupies the porous medium. Consequently,the relative permeability functions fulfill thefollowing conditions:Gas only Liquid onlyk r‘ (0) ¼ 0, k r‘ (1) ¼ 1k rg (0) ¼ 1, k rg (1) ¼ 0(<strong>36</strong> :9)Although depending on both initial and boundaryconditions, the relative permeability values are usuallysupposed to be the function <strong>of</strong> saturation only.Even with this simple assumption, their experimentaldetermination remains very challenging.Very few results are available in the literature forwood. Some data have been published by Tesoro et al.(1972, 1974). A recent paper by Tremblay et al. (2000)describes an indirect method to measure this property.Based on the geometrical model <strong>of</strong> tracheids proposedby Comstock (1970), Spolek and Plumb (1980)derived an expression for the relative permeability ins<strong>of</strong>twoods. This model assumes that all tracheids areexactly similar and that the wetting-phase distributionis ideal. Their final expression involves an irreduciblesaturation, below which no liquid flux is possible.This concept <strong>of</strong> irreducible saturation has beenwidely discussed, especially in the context <strong>of</strong> petroleumproduction (Dullien, 1992). Depending on theboundary conditions, one part <strong>of</strong> the fluid phasethat occupied the medium at the beginning <strong>of</strong> theexperiment remains in the medium, even though itsflow rate has vanished. This part seems to be trappedin the medium in what is called a ‘‘pendular’’ state.The amount <strong>of</strong> the residual part increases with theimbibition or drainage velocity. In addition, theamount <strong>of</strong> trapped phase after the experiment reduceswith time; surface spreading in the solid phase and theedge-capillary pressure are the two mechanisms thatcan explain this observation (Dullien, 1992). Due tothe microporosity that exists in the cell walls, suchmechanisms are likely to exist in wood also. However,the concept <strong>of</strong> irreducible saturation has led to unrealisticcomputed moisture content pr<strong>of</strong>iles duringdrying (Perré , 1987), though later modeling <strong>of</strong> s<strong>of</strong>twooddrying incorporates a similar concept successfullyas a mean to account for pit aspiration (Nijdamet al., 2000). In this case, ‘‘irreducible saturation’’corresponds to the condition when pit aspiration issufficiently advanced such that the remaining freemoisture is immobilized in isolated pockets in thewood.Based on the measurements in the longitudinaldirection reported by Tesoro et al. (1972), on thecurve computed from the tracheid model (Spolekand Plumb, 1980), and on those considerationsregarding the concept <strong>of</strong> irreducible saturation, thefollowing functions have been proposed for s<strong>of</strong>twoodsby Perré et al. (1993) (Figure <strong>36</strong>.14):In the transverse direction (radial or tangential)k T rg ¼ 1 þ (2S 3)S2 and k T r‘ ¼ S3 (<strong>36</strong>:10)In the longitudinal directionk L rg ¼ 1 þ (4S 5)S4 and k L r‘ ¼ S8 (<strong>36</strong>:11)<strong>36</strong>.2.2.3 Capillary PressureSome works can be found on the determination <strong>of</strong>capillary pressure functions in wood. Mercury porosimetryexhibits a dramatic effect <strong>of</strong> the sample thicknessin the longitudinal direction (Trénard, 1980). Forshort samples, the cell lumens are directly accessible tothe mercury; whereas for longer samples, the liquid hasto pass through the small openings, the pits that existbetween cells, clearly illustrating their bottleneckingeffect. The centrifuge method has been used successfullyby Spolek and Plumb (1981) on s<strong>of</strong>twoods and byChoong and Tesoro (1989) on various species. Todetermine the moisture content–water potential relationship<strong>of</strong> wood, Cloutier and Fortin (1991) used atension plate and a pressure plate. This is a drainagemethod and their results could easily be converted intoa classic capillary pressure function.Because certain anatomical features govern themorphology <strong>of</strong> wood pores, particular methods canbe applied to wood:. To compute a capillary pressure curve, Spolekand Plumb (1981) have developed the geometricalß 2006 by Taylor & Francis Group, LLC.


FIGURE <strong>36</strong>.14 Relative permeability curves calculated using equations (<strong>36</strong>.10) and (<strong>36</strong>.11).10.8TransverseLongitudinalRelative permeability0.60.40.200 0.2 0.4 0.6 0.8Saturation1model <strong>of</strong> the tracheid shape proposed by Comstock(1970). Although it may be simplistic toassume that all tracheids have exactly the sameshape,theyobtainedagoodtrendforthecapillarypressure function by this means.. Because the longitudinal direction <strong>of</strong> wood isvery marked, it is quite simple to obtain thethree-dimensional structure <strong>of</strong> the materialfrom a cross section. Figure <strong>36</strong>.15 depicts theexamples <strong>of</strong> capillary pressure curves calculatedfrom microscopic images <strong>of</strong> cross sections <strong>of</strong>wood. In this case, the pore-size distributionhas been calculated using image processing(Perré , 1997; Perré and Turner, 2002).<strong>36</strong>.2.2 .4 Bound -Water Diffusio nMacroscopic bound-water diffusion results fromtransport mechanisms that take place at the microscopicscale, i.e., diffusion <strong>of</strong> the bound waterthrough the cell walls and vapor diffusion due toFick’s law. At the microscopic scale, these two fluxescan be expressed as follows:f b ¼ r s D b rX b (Bound-water flux) (<strong>36</strong> :12)f v ¼ r g D v rv v (Vapor flux) (<strong>36</strong> :13)In Equation <strong>36</strong>.12 and Equation <strong>36</strong>.13, D b and D vrepresent the microscopic bound-liquid and vapordiffusivities, respectively having units m 2 s 1 and v vis the mass fraction <strong>of</strong> vapor in the gaseous phase.By using the bound-liquid diffusivity data <strong>of</strong>Stamm (1963), it is possible to obtain the followingleast-squares, best-fit correlation for D b :4300D b ¼ exp 12 :82 þ 10 :90 X b (<strong>36</strong>:14)Twhere T is the temperature in Kelvin.On assuming isothermal conditions and constanttotal pressure, the microscopic vapor flux can be expressedwith the gradient <strong>of</strong> the bound-water contentas the driving force by Equation <strong>36</strong>.13. M vf v ¼RT D @P vv rX b (<strong>36</strong>:15)@X bWithin the anatomical structure <strong>of</strong> wood, anycombination in series or parallel <strong>of</strong> vapor diffusion(in lumen and pits) and bound-water diffusion (in thecell walls) is a possible pathway to drive water fromhigh to low moisture content regions (Figure <strong>36</strong>.16).Because Equation <strong>36</strong>.12 and Equation <strong>36</strong>.15 use thesame driving force, the expressions for the macroscopicbound-water diffusivity in the radial and tangentialdirectionscanbecalculatedevenusinghomogenizationtechniques according to these microscopic properties,together with the pore morphology (Perré and Turner,2002). Equation <strong>36</strong>.14, derived from specific experimentalmeasurements, exhibits a dramatic increase <strong>of</strong>ß 2006 by Taylor & Francis Group, LLC.


0.4Capillary pressure (P c /P atm )0.30.20.1458 kg m 3 641 kg m 3Late woodMiddle woodEarly wood349 kg m 30(a) 0 0.2 0.4 0.6 0.8 1.0Saturation0.8Capillary pressure (P c /P atm )0.60.40.2656 kg m 3543 kg m 3SaturationLate woodEarly wood(b)00 0.2 0.4 0.6 0.8 1.0FIGURE <strong>36</strong>.15 Capillary-function curves determined using image analysis. (a) spruce (Picea abies): the cells have thickerwalls and smaller radial extension in latewood part, hence the highest value <strong>of</strong> the capillary-pressure curve. One also has to beaware that full saturation is obtained with a lower amount <strong>of</strong> water in latewood (the porosity <strong>of</strong> this part is very small); (b)beech (Fagus sylvatica): because beech is a pore diffuse-porous hardwood species, no significant difference is observedbetween these parts. The low capillary pressure obtained for saturation values above 0.2 corresponds to the meniscus radiilocated in the vessel elements. The dramatic increase for low saturation values is due to the small lumen diameters <strong>of</strong> theparenchyma and fiber cells.the bound-water diffusivity as the bound moisture contentincreases. For this reason, the same trend is predictedfrom the calculations or measurements at themacroscopic scale; bound-water diffusion is alwayseasier for higher bound-water contents.When using the gradient <strong>of</strong> bound water as adriving force, the macroscopic flux readsf b ¼ r 0 D b rX b (<strong>36</strong>:16)This expression is consistent with the derivation <strong>of</strong> thesecond Fick’s law by using the mass balance equation:@X b@t¼r(D b rX b ) (<strong>36</strong>:17)For the sake <strong>of</strong> simplicity, the foregoing expressionshave always assumed isothermal conditions. However,this assumption fails for certain drying processesß 2006 by Taylor & Francis Group, LLC.


Low moisture contentHigh moisture contentFIGURE <strong>36</strong>.16 Moisture diffusion in the hygroscopicrange: any combination in series or parallel <strong>of</strong> vapor diffusion(in lumen and pits) and bound-water diffusion (in thecell walls) is able to drive water from high to low moisturecontent regions.(e.g., impingement drying, radio-frequency, or microwaveheating). The problem <strong>of</strong> mass migration due to athermal gradient is a matter <strong>of</strong> scientific debate. Siau(1984, 1995) gives a good review <strong>of</strong> possible formulationsthat can be used in nonisothermal conditions.<strong>36</strong>.2.2.5 Physical FormulationSeveral sets <strong>of</strong> macroscopic equations are proposed inthe literature for the simulation <strong>of</strong> the drying process.The first fundamental difference between them lies inthe number <strong>of</strong> state variables used to describe theprocess:. One: moisture content (or an equivalent variable:saturation, water potential, etc.). Two: moisture content (or equivalent) and temperatureT (or an equivalent variable: enthalpyetc.). Three: moisture content (or equivalent), T (orequivalent), and gaseous pressure P g (or anequivalent variable: air density, intrinsic airdensity, etc.)Perré (1999) gives a critical review <strong>of</strong> the possibilitiesand limitations given by these different sets <strong>of</strong>equations. The use <strong>of</strong> moisture content alone is thebasis <strong>of</strong> many correlations <strong>of</strong> lumber-drying rates(see Keey et al., 2000). Doe et al. (1996a) have foundthat, at relatively low temperatures (


Water Conservation@@t « wr w þ « g r v þ r bþr rw v w þ r v v g þ r b v b¼rðrg¼Deff rv v ÞðT1ÞEnergy Conservation@@t « wr w h w þ « g ðr v h v þ r a h a Þþr bhb þ r o h s « g P g þr rw h w v w þ ðr v h v þ r a h a Þv g þ h b r b v b¼¼r r gDeff ðh v rv v þ h a rv a Þþl eff rT þ FðT2ÞAir Conservation@@t « ¼gr a þr ra v g ¼r rgDeff rv a(T3)where the gas- and liquid-phase velocities are given by the generalized Darcy’s law:v ‘ ¼¼K ‘¼kr‘m ‘rw ‘ , rw ‘ ¼rP ‘ r ‘ grx (T4)where ‘ is w, g, the quantity w is known as the phase potential, and x is the depth scalar. All other symbolshave their usual meaning.Boundary ConditionsFor the external drying surfaces <strong>of</strong> the sample, the boundary conditions are assumed to be <strong>of</strong> the following form: 1 x 1J w j x¼0 þ ^n ¼ h m cM v ln1 x v j x¼0(T5)J e j x¼0 þ ^n ¼ h(Tj x¼0 T 1 )P g j x¼0 þ ¼ P atmwhere J w and J e represent the fluxes <strong>of</strong> total moisture and total enthalpy at the boundary, respectively, and xdenotes the normal position from the boundary in the external medium.<strong>36</strong>.2.3 PROCESS OF DRYING<strong>36</strong>.2.3 .1 Lo w-Temper ature Convecti ve Dry ingLow-temperature convective drying is the most widespreadindustrial process for seasoning wood in kilns.In this case, the role <strong>of</strong> internal gaseous pressure isalmost negligible and transfer occurs mainly in thedirection <strong>of</strong> the board thickness. Two periods <strong>of</strong> dryingmay be distinguished: (1) a constant drying-rateperiod and (2) a decreasing drying-rate period.<strong>36</strong>.2.3 .1.1 The Cons tant <strong>Drying</strong>- Rate Peri odThis stage is very common for certain porous media,but is rarely seen with wood. However, it exists almostalways for fresh boards consisting <strong>of</strong> sapwoodthat are dried under moderate conditions (Perré et al.,1993; Perré and Martin, 1994). During this period,the exposed surface <strong>of</strong> the board is still above theFSP. As a result, the vapor pressure at the surface isequal to the saturated vapor pressure, and is a function<strong>of</strong> the surface temperature only.Coupled heat and vapor transfer occur across inthe boundary layer (Figure <strong>36</strong>.17). The heat flux suppliedby the airflow is used solely for transforming theliquid water into vapor. During this stage, the dryingrate is constant and depends only on the externalconditions (temperature, relative humidity, velocity,and flow configuration). The temperature at the surfaceis equal to the wet-bulb temperature. Moreover,because no energy transfer occurs within the mediumduring this period, the whole temperature <strong>of</strong> theboard remains at the wet-bulb temperature.ß 2006 by Taylor & Francis Group, LLC.


Boundary layersExternal flowTP vHeatVaporCapillary migrationLow moisturecontent=small radiusLiquid flow<strong>Wood</strong>High moisturecontent=large radiusFIGURE <strong>36</strong>.17 Constant drying-rate period: the moisture migrates inside the medium mostly by capillary forces; evaporationoccurs at the exchange surface with a dynamical equilibrium within the boundary layers between the heat and the vaporflows. (Adapted from Perré, P., The numerical modeling <strong>of</strong> physical and mechanical phenomena involved in wood drying: anexcellent tool for assisting with the study <strong>of</strong> new processes, Tutorial, Proceedings <strong>of</strong> the Fifth International IUFRO <strong>Wood</strong><strong>Drying</strong> Conference, Québec, Canada, 1996, 9–38.)The exposed surface is supplied with liquid watercoming from the inside <strong>of</strong> the board by capillaryaction; the liquid migrates from regions with highmoisture content (liquid–gas interfaces within largepores) toward regions with low moisture content(liquid–gas interfaces within small pores).The constant drying-rate period lasts as long as thesurface is supplied with liquid. Its duration dependsstrongly on the drying conditions (magnitude <strong>of</strong> theexternal flux) and on the medium properties. The liquidflow inside the medium is expressed by Darcy’slaw (permeability gradient <strong>of</strong> liquid pressure).<strong>36</strong>.2.3 .1.2 The Decr easing <strong>Drying</strong>- Rate Pe riodOnce the surface attained the hygroscopic range,the vapor pressure becomes smaller than the saturatedvapor pressure (Figure <strong>36</strong>.18). Consequently, theexternal vapor flux is reduced and the heat flux suppliedto the medium is temporarily greater than whatis necessary for liquid evaporation. The energy inexcess is used to heat the board, the surface at firstand then the inner part by conduction. A new, moresubtle, dynamic equilibrium takes place. The surfacevaporpressure,hencetheexternalvaporflow,dependson both temperature and moisture content. To maintainthe energy balance, the surface temperature increasesas the surface moisture content decreases. Thisleads to a decreasing drying rate (the heat supplied bythe airflow becomes smaller and smaller).A two-zone process develops inside the wood: (1)an inner zone, where liquid migration prevails, and(2) a surface zone, where both bound-water andwater-vapor diffusion take place. During this period,a conductive heat flux must exist inside the board toincrease the temperature and to evaporate the liquiddriven by gaseous diffusion. The region <strong>of</strong> liquidmigration naturally reduces as the drying progressesand finally disappears. The process is finished whenthe temperature and the moisture content attain theoutside air temperature and the EMC, respectively.<strong>36</strong>.2.3 .2 Dry ing at High Temper ature: The Effect<strong>of</strong> Inter nal Pressur e on Mass Tr ansferTo reduce the drying time without decreasing thequality <strong>of</strong> the dried product, the drying conditionsmust be such that the temperature <strong>of</strong> the product isabove the boiling point <strong>of</strong> water. Such conditionsensure that an overpressure exists within the material,which implies that a pressure gradient drives themoisture (liquid or vapor) toward the exchange surfaces(Lowery, 1979; Kamke and Casey, 1988).At normal atmospheric pressure, the boiling point<strong>of</strong> water equals 1008C. Consequently, in order toß 2006 by Taylor & Francis Group, LLC.


External flowDiffusion <strong>of</strong> vaporand bound waterFew vapormoleculesHeatVaporBound waterand vaporCapillary migrationMany vapormoleculesLow moisturecontent=small radiusLiquid flowHigh moisturecontent=large radiusFIGURE <strong>36</strong>.18 Second drying period: a region in the hygroscopic range develops from the exposed surface. In that region,both vapor diffusion and bound-water diffusion act. Evaporation takes place partly inside the medium. Consequently, a heatflux has to be driven toward the inner part <strong>of</strong> the board by conduction. (Adapted from Perré , P., The numerical modeling <strong>of</strong>physical and mechanical phenomena involved in wood drying: an excellent tool for assisting with the study <strong>of</strong> new processes,Tutorial, Proceedings <strong>of</strong> the Fifth International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec, Canada, 1996, 9–38.)obtain an internal overpressure, the temperature <strong>of</strong>the porous medium must be above that level during atleast one part <strong>of</strong> the process. This is the aim <strong>of</strong> convectivedrying at high temperature (moist air orsuperheated steam) and a possible aim <strong>of</strong> contactdrying or drying with an electromagnetic field (microwaveor radio frequency).However, as shown in Figure <strong>36</strong>.19, it is possibleto reduce the boiling point <strong>of</strong> water by decreasing theexternal pressure and, consequently, to obtain a hightemperatureeffect with relatively moderate dryingconditions. This is the principle <strong>of</strong> vacuum drying,particularly useful for lumber that would be damagedby high temperature levels.Whenever an overpressure exists inside a board,the large anisotropy ratios imply intricate transfermechanisms. Heat is <strong>of</strong>ten supplied in the thicknessdirection while, in spite <strong>of</strong> the length, the effect <strong>of</strong>the pressure gradient on gaseous (important for lowmoisture content) or liquid migration (important forhigh moisture content) takes place in the longitudinaldirection (Figure <strong>36</strong>.20). This is a result <strong>of</strong> the anatomicalfeatures <strong>of</strong> wood. In the case <strong>of</strong> very intensiveinternal transfer, the end piece can be fully saturatedand, sometimes, moisture can leave the sample in theliquid state. (This is quite easy to observe duringmicrowave heating.)Pressure (kPa)10050Atmospheric pressureExternal pressureBoiling temperature00 20 40 60 80 100 120Temperature (C)Saturated vaporFIGURE <strong>36</strong>.19 Vacuum drying seeks to reduce the boilingpoint <strong>of</strong> water in order to obtain a high-temperature effectwith moderate drying conditions. (saturated vapor pressurevalues from Lide 1995.)<strong>36</strong>.2.3 .3 Ty pical <strong>Drying</strong> Behavior : Differenc ebe tween Sapwood and Hear twoodIn a tree, freshly cut down and sawn, it is easy todistinguish sapwood from heartwood (by touch or bysight). But a few days later, the loss <strong>of</strong> surface moisturecontent makes it impossible to do that. Nevertheless, inß 2006 by Taylor & Francis Group, LLC.


HeatVaporVessel ortracheidLiquidPitsEndpiecefully saturatedOverpressureLiquid evacuationpossible inmicrowave heatingFIGURE <strong>36</strong>.20 <strong>Drying</strong> at high temperature (second drying period): a high-temperature regime means that an overpressuredevelops inside the medium. Depending on the moisture content, this overpressure induces liquid or gaseous flow; inaddition, as wood is strongly anisotropic, the most part <strong>of</strong> the flow occurs in the longitudinal direction (see the magnifiedviews). (Adapted from Perré, P., The numerical modeling <strong>of</strong> physical and mechanical phenomena involved in wood drying:an excellent tool for assisting with the study <strong>of</strong> new processes, Tutorial, Proceedings <strong>of</strong> the Fifth International IUFRO <strong>Wood</strong><strong>Drying</strong> Conference, Québec, Canada, 1996, 9–38.)the case <strong>of</strong> high-temperature drying, the increase <strong>of</strong>internal pressure gives rise to longitudinal migration<strong>of</strong> liquid toward the end pieces, provided that thepermeability and the moisture content are highenough. This is a good way to spot sapwood after afew hours <strong>of</strong> drying (Figure <strong>36</strong>.21b). This phenomenoncan be observed in industrial kilns (Figure <strong>36</strong>.21c).To illustrate the effect <strong>of</strong> these differences on thedrying process, Figure <strong>36</strong>.22 depicts drying experimentscarried out with superheated steam at 150 8Con both sapwood and heartwood <strong>of</strong> Norway spruce(Picea abies). They are representative <strong>of</strong> the trendsobserved by different authors (Salin, 1989; Pang et al.,1994; Perré and Martin, 1994).After the initial transient period, the constant drying-rateperiod takes place for the sapwood board.During this period, which lasts several hours, alltemperatures are equal to the wet-bulb temperatureand the overpressure remains very small. At the beginning<strong>of</strong> the second drying period (around 350min), an important overpressure develops due tothe temperature increase. It disappears only when the(a) A stack at the beginning <strong>of</strong> the dryingHeartwoodSapwoodSapwood Heartwood(b) The same stack after a few hours <strong>of</strong> dryingat high temperature(c)FIGURE <strong>36</strong>.21 A stack <strong>of</strong> boards during high-temperature drying (shaded areas indicate wet zones).ß 2006 by Taylor & Francis Group, LLC.


Air flow041 381.0 m65Different locationsin the sample30 mm1501.0T 11300.8Temperature (˚C)11090T 8T 5T 3P 3P 50.60.4Overpressure (P atm )(a)70500100 200 300 400 500 600 700 0 0.2Time (min)1501.25T 1T 8T 5T 3P 3P 51301.00Temperature (˚C)110900.750.50Overpressure (P atm)(b)700.25500 100 200 300 400 500 0Time (min)FIGURE <strong>36</strong>.22 Experiment on spruce (Picea abies) dried with superheated steam at 1508C. Temperature and internalpressure at different locations. Note the difference between sapwood (a) and heartwood (b).entire board enters the hygroscopic range. At thismoment, all temperatures approach the dry-bulbtemperature.The results obtained for heartwood are quite different.No constant drying-rate period can be observed.Just a short plateau at the boiling pointß 2006 by Taylor & Francis Group, LLC.


Moisture content (%)15010050SapwoodHeartwood00 200 400 600Time (min)FIGURE <strong>36</strong>.23 Moisture content loss obtained for sapwoodand for heartwood (same experiments as in Figure <strong>36</strong>.22).is detectable at the rear end <strong>of</strong> the board (T 8 ).Consequently, the overpressure remains high (especiallyfor the center pressure P 3 ) up to the end <strong>of</strong> thedrying. The maximum pressure is higher for heartwoodthan for sapwood. The differences in dryingkinetics are <strong>of</strong> great interest. In spite <strong>of</strong> the high initialmoisture content <strong>of</strong> sapwood (170% against 60%), thepermeability <strong>of</strong> heartwood is so low that the curvescross each other at 450 min <strong>of</strong> drying (Figure <strong>36</strong>.23).This is consistent with the observations on entirestacks (Salin, 1989).The strategy <strong>of</strong> simulating the differences betweenheartwoodandsapwoodliesinonlytwosets<strong>of</strong>parameters:the permeability and the initial moisture content(fortheseexperiments,180%forsapwoodand70%forheartwood). The values <strong>of</strong> permeability used to differentiatesapwood from heartwood (Table <strong>36</strong>.8) arebased on the considerations concerning pit aspiration.By using only these differences, Perré and Turner(1996) found that all the trends observed for sapwoodand heartwood were found in the simulated results.The most spectacular effect is the longitudinal flowdue to the overpressure (Figure <strong>36</strong>.24). In the case <strong>of</strong>high-temperature convective drying, the sapwoodboard delivers a large supply <strong>of</strong> water to the endpiece after 5 h, while the heartwood end piece isalready within the hygroscopic range. These carpetplots should be compared with Figure <strong>36</strong>.21.<strong>36</strong>.2.4 MECHANICAL A SPECTS OF WOOD DRYING<strong>36</strong>.2.4 .1 Mech anical Behavior <strong>of</strong> W oodIndustrial wood drying consists <strong>of</strong> not only removingmoisture from greenwood but also ensuringthat its quality (fitness for purpose) is adequatein end use. Because wood shrinks during drying,deformations and stresses develop that can lead toTABLE <strong>36</strong>.8Intrinsic Permeabilities Used in the <strong>Drying</strong> ModelingDirection Heartwood SapwoodLongitudinalGas 10 13 m 2 2.10 13 m 2Liquid 10 13 m 2 10 12 m 2TransverseGas 10 16 m 2 2.10 16 m 2Liquid 10 16 m 2 10 15 m 2unusable products (Figure <strong>36</strong>.25). The understanding<strong>of</strong> these aspects must account for the complex mechanicalbehavior <strong>of</strong> wood, including its memory effect.Shrinkage is the ‘‘driving’’ force for drying stress,i.e., without shrinkage, no drying stresses would develop.Figure <strong>36</strong>.26 exhibits the dimensional variation<strong>of</strong> an unladen sample with the moisture content(the latter is assumed to be uniform). Under normalconditions, the dimensions do not change until themoisture content attains a moisture content close tothe acknowledged FSP. This condition is sometimescalled SIP. Then, the dimension variations are almostproportional to the change in moisture content. Thisstrain field is called free shrinkage.A sample subjected to tensile or compressive stress(Figure <strong>36</strong>.27) exhibits the instantaneous deformation(elastic part) at first, which then increases with time(viscoelastic creep). After cycling the moisture contentto and from a higher moisture content, the creephas been significantly greater due to mechanosorptiveaction.Thus, a sample subjected to a compressive stress(as shown in case 1, Figure <strong>36</strong>.28) exhibits a smallerlength at the end <strong>of</strong> drying than an unloaded specimen(case 2), which itself has a smaller length thanthe sample subjected to a tensile stress (case 3). Inthis experiment, the viscoelastic behavior acts because<strong>of</strong> time and the mechanosorptive behavioracts because <strong>of</strong> the removal <strong>of</strong> water molecules dueto drying.These ideas can now be applied to the drying <strong>of</strong>lumber boards, which is assumed to be stress-free atthe beginning. At the beginning <strong>of</strong> drying (constantdrying-rate period), sap throughout the entire boardremains free. No shrinkage occurs; hence, stressbuildup is absent.At the beginning <strong>of</strong> the second drying period,shrinkage exists close to the exposed surfaces(Figure <strong>36</strong>.29a). At this moment, if the section wascut into slices, the outer slices would have a shorterlength than the inner ones (Figure <strong>36</strong>.29b). This displacementfield is not compatible and induces, in theß 2006 by Taylor & Francis Group, LLC.


504030Length (cm)2010Moisture content1.00.50.00 1 20Width (cm)50403020Length (cm)100Moisture content1.00.50.00 1 2Width (cm)504030Length (cm)2010150125100755000 12TemperatureWidth (cm)504030Length (cm)2010015012510075500 1 2TemperatureWidth (cm)50(a)4030Length (cm)20100Pressure1.81.61.41.21.00 1 2Width (cm)50(b)4030Length (cm)20100Pressure1.81.61.41.21.00 1 2Width (cm)FIGURE <strong>36</strong>.24 High-temperature drying (140/858C). Carpet plot after 5 h <strong>of</strong> drying. Internal overpressure, resaturation <strong>of</strong>the end piece, thermal conduction along the thickness, and end piece close to the wet-bulb temperature are evident on theseplots. Note the high value <strong>of</strong> internal pressure and the absence <strong>of</strong> end-piece resaturation obtained for heartwood (b).actual section, a tensile stress in the surface layers and(because <strong>of</strong> equilibrium conditions) a counteractingcompressive stress in the core layers (Figure <strong>36</strong>.29c).During this period, surface checking is possible.From this point onward, the wood layers dry underload.As the drying proceeds, viscoelastic creep develops,together with mechanosorptive creep. The outer slicesappear similar in configuration to that exhibited forslice n83inFigure<strong>36</strong>.28,whiletheinternalslicesresembleslice n 81. Consequently, in spite <strong>of</strong> the flat moisturecontent pr<strong>of</strong>ile, slicing the section at the end <strong>of</strong> thedrying wouldgivepictureFigure<strong>36</strong>.30b; the coreslices,dried under compression, are smaller than the outerones, dried under tension. In the actual section, compressivestress exists in the inner part (Figure <strong>36</strong>.30c).This phenomenon is known as stress reversal or casehardening.The residual stress level depends on manyparameters (growth history, sawing pattern, dryingconditions, species, thickness, etc.), which provide most<strong>of</strong> the problems <strong>of</strong> drying optimization. In addition,one must keep in mind that gradients <strong>of</strong> moisturecontent, strain, and stress exist along the thickness.This explains the curvature <strong>of</strong> the slices observed inprong test or cup method commonly used in industryto assess stress levels (Figure <strong>36</strong>.31). When the innertensile stress is too high, internal checking occurs(Figure <strong>36</strong>.25). An interesting simulation <strong>of</strong> this testcan be found in Dahlblom et al. (1994).<strong>36</strong>.2.4 .2 Dry ing Stres s Fo rmulationDuring drying, shrinkage appears in all parts <strong>of</strong> theboard for which the moisture content X is within thehygroscopic range. The shrinkage strain is proportionalto the difference between the local moisture content andthelocalvalue<strong>of</strong>themoisturecontentatfibersaturationat the same temperature. A deformation field noted, « sh ,is defined in the material’s axes by Equation S1.If this deformation field does not fulfill the geometricalcompatibility, a strain tensor « mec related tostress is generated. The constitutive equation, whichrepresents the mechanical behavior <strong>of</strong> the material,relates this strain tensor « mec and the stress tensor.Due to the memory effect <strong>of</strong> wood, this tensor « mechas to be divided into two parts: (1) an elastic strain,« elas , connected to the actual stress tensor and (2) amemory strain, « mem , which includes all the strain dueß 2006 by Taylor & Francis Group, LLC.


to the history <strong>of</strong> that point (« mem can deal with plasticity,creep, mechanosorption, etc.).The geometrical compatibility applies to thetotal strain field « tot . When solving the mechanicalproblems in terms <strong>of</strong> displacement, the total straintensor is deduced from the displacement field and thisgeometrical condition is automatically fulfilled withinthe domain. The stress field must satisfy the local mechanicalequilibrium and the boundary conditions.Finally, the complete formulation <strong>of</strong> the stress problemis given by Equation S1 through Equation S4.<strong>36</strong>.2.4.3 Memory EffectWhile describing the strain field, « mem , lies the entireproblem <strong>of</strong> developing a constitutive model forwood, which requires both theoretical and numericalwork. Comprehensive formulations are also verydifficult to characterize (Ranta-Maunus, 1975). Theproblem lies in the fact that the memory effect <strong>of</strong>wood depends not only on the temperature and moisturecontent values but also on their variations in timeand on the history <strong>of</strong> their variations in time. This2 3A 0 0« sh ¼ H(~x) 4 0 B 0 5 (S1)0 0 CwithH(~x) ¼ 0 X(~x) X fsp if X (~x) $ X fspif X (~x) # X fsp8« tot« mec ¼ « elas þ « mem (S2)ij ¼ 1 2 (u i, j þ u j,i) over V>< s ij, j þ rf i ¼ 0 over Vs ij ¼ a ijkl (« totkl« 0 kl ) over V with «0 = « sh + « mem and 8i , G Di G Ti = Gs ij n j ¼ T ion G >:Tiu i ¼ D i ¼ 0 on G Di(S3)Remarks:. This static formulation requires that boundary and volumetric forces satisfy the global equilibrium.. G is the surface surrounding the domain V. G Di refers to the subdomain <strong>of</strong> G where the i component<strong>of</strong> the displacement is known and G Ti to the subdomain <strong>of</strong> G where the i component <strong>of</strong> the tractionforce is known. In order to ensure the uniqueness <strong>of</strong> the solution, additional conditions are requiredon the boundary conditions: 8i, mes(G Di ) > 0. Otherwise, the solution is defined within a rigid bodymotion.. As wood is orthotropic, each behavior law involves nine independent terms. In fact, it is morecommon to define the inverse <strong>of</strong> a ijkl that, for the case <strong>of</strong> linear elasticity, leads to the generalizedHooke’s law:223« LL« RR« TT6 2« LR ¼ g LR¼74 2« LT ¼ g LT52« RT ¼ g RT641E L LR RLE R1E L LTE LE R TL RT 1E R0 0 0E T TR0 0 0E T0 0 0E T10 0 00 0G LR10 0 0 0 0G LT0 0 0 0 01G RT326475s LLs RRs TTs LRs LTs RT375(S4)ß 2006 by Taylor & Francis Group, LLC.


FIGURE <strong>36</strong>.25 Two examples <strong>of</strong> mechanical degrade during wood drying: board deformation and internal checking.GreenwoodSIPOven-driedLength0SIPMoisture contentFIGURE <strong>36</strong>.26 <strong>Wood</strong> shrinkage: the shrinkage intersection point (SIP), <strong>of</strong>ten close to 30%, depends on species andtemperature.ß 2006 by Taylor & Francis Group, LLC.


ConstantmoisturecontentTime t = 0 +LoadTime tLoadAfter cycles <strong>of</strong>moisture contentLoadMoisturecontentLoadTimeTimeLengthvariationViscoelasticMechanosorptiveTimeFIGURE <strong>36</strong>.27 Viscoelastic and mechanosorptive behavior <strong>of</strong> wood. (Adapted from Perré, P., The numerical modeling <strong>of</strong>physical and mechanical phenomena involved in wood drying: an excellent tool for assisting with the study <strong>of</strong> new processes,Tutorial, Proceedings <strong>of</strong> the Fifth International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec, Canada, 1996, 9–38.)subjectremainsamatter<strong>of</strong>somescientificdebate(Keeyet al., 2000).Nevertheless, in the case <strong>of</strong> drying, the moisturecontent only decreases and some simplificationsapply. Here, only the most common way to expresscreep and mechanosorptive effect will be presented.The general formulation <strong>of</strong> the time dependency <strong>of</strong>the creep property involves the whole stress history:« ij ¼ð t 1J ijkl (t t 0 ) d s kldt 0 dt0 (<strong>36</strong> :18)where J ijkl (t) is the creep compliance tensor and t isthe actual time. The experimental creep function is<strong>of</strong>ten analyzed as a number <strong>of</strong> Kelvin elements inseries (Figure <strong>36</strong>.32), each having the property <strong>of</strong>a spring and dashpot in parallel (Genevaux, 1989;Martensson, 1992; Mohager and Toratti, 1993; Hanhijärvi, 1999 Passard and Perre, 2005). In the case <strong>of</strong>uniaxial load, this leads toJ(t) ¼ J 0 1 þ XNa n (1 et t!n ) (<strong>36</strong>:19)n ¼1The temperature and moisture dependency <strong>of</strong> thatfunction can be expressed using a material time orchanging the characteristic time t n . The thermal activation,for example, is <strong>of</strong>ten expressed with the aid <strong>of</strong>an Arrhenius law: t n ¼t 1 n exp DW n(<strong>36</strong>:20)RTTime t = 0 + highmoisture contentTime t lowmoisture contentLoad1 2 3 <strong>Drying</strong>1 2 3LoadFIGURE <strong>36</strong>.28 Dimension changes <strong>of</strong> a specimen loaded during drying. (Adapted from Perré, P., The numerical modeling <strong>of</strong>physical and mechanical phenomena involved in wood drying: an excellent tool for assisting with the study <strong>of</strong> new processes,Tutorial, Proceedings <strong>of</strong> the Fifth International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec, Canada, 1996, 9–38.)ß 2006 by Taylor & Francis Group, LLC.


Second drying periodSecond drying periodEnd <strong>of</strong> dryingFSPTensilestressCompressivestressProngtest(a)(b)FIGURE <strong>36</strong>.29 Appearance <strong>of</strong> drying stresses followingshrinkage. (Adapted from Perré , P., The numerical modeling<strong>of</strong> physical and mechanical phenomena involved inwood drying: an excellent tool for assisting with the study<strong>of</strong> new processes, Tutorial, Proceedings <strong>of</strong> the Fifth InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Qué bec, Canada,1996, 9–38.)where DW n is the activation energy associated toelement n.The mechanosorptive effect occurs as soon as themoisture content changes during load. The simplerway to express this effect consists in assuming thatthe strain rate depends linearly on both the stress fieldand the time derivative <strong>of</strong> the moisture content ḣ :_« msij ¼ m ijkl s kl (sign ḣ)ḣ (<strong>36</strong>:21)The mechanosorptive strain rate is always in the direction<strong>of</strong> the stress field, hence the factor sign _u inEquation <strong>36</strong>.21.A three-dimensional resolution is very costly interms <strong>of</strong> calculation time and computer memoryspace. The need for such a cost is justified wheneverthe objective <strong>of</strong> the calculation lies in evaluating theoverall deformation <strong>of</strong> the board. In this way, theeffect <strong>of</strong> reaction wood, fiber angle, and propertyFSPX eq(a)End <strong>of</strong> drying(b)(c)(c)TensilestressCompressivestressFIGURE <strong>36</strong>.30 Stress reversal due to the memory effect <strong>of</strong>wood. (Adapted from Perré, P., The numerical modeling <strong>of</strong>physical and mechanical phenomena involved in wooddrying: an excellent tool for assisting with the study<strong>of</strong> new processes, Tutorial, Proceedings <strong>of</strong> the Fifth InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec,Canada, 1996, 9–38.)CupmethodFIGURE <strong>36</strong>.31 Two experimental methods used to assessdrying stresses.variations can be analyzed. Good examples <strong>of</strong> thesepossibilities can be found in the literature (Dahlblomet al., 1994, 1996; Ormarsson, 1999). Nevertheless, tostudy the stress development within a section farfrom the ends <strong>of</strong> the board, a two-dimensional simulationis sufficient. A ‘‘planar displacement’’ formulationhas to be used in this case, assuming smalldisplacements (Perré and Passard, 1995; Chen et al.,1997a)orlargedisplacements(MaugetandPerré ,1999).<strong>36</strong>.2.4 .4 Stres s Deve lopm ent dur ing Dry ing:So me Exampl esAll examples presented in this section, except the nonsymmetriccase, refer to a flatsawn board <strong>of</strong> heartwood,20-mm thick and 40-mm wide. They have beencomputed using the computer code Transpore (Perréand Turner, 1996b and c, 1999; Mauget and Perré ,1999; Perré and Passard, 2002;). Figure <strong>36</strong>.33 depictsthe moisture and stress fields calculated at differentdrying stages for quite severe conditions at a mediumtemperature level (T d ¼ 80 8C, T w ¼ 60 8C). After 6 h<strong>of</strong> drying, the external part <strong>of</strong> the section is within thehygroscopicrange,whichgivesrisetotensilestressduetoshrinkageinthezonesclosetotheexchangesurface.As a consequence <strong>of</strong> the mechanical equilibrium, internalzones undergo compressive stress. A negativeshear stress exists close to the edge (the right angleshavedecreased).Attheend<strong>of</strong>thedryingprocess(60h)the moisture content is almost equal to the EMC (7%)throughout the section. The stress reversal due to thememory effect <strong>of</strong> wood is clearly exhibited. It can benoted that the shear stress also changes in sign.When a face <strong>of</strong> a board is insulated, the dryingconditions are not symmetrical anymore andß 2006 by Taylor & Francis Group, LLC.


Element 1 Element 2 Element NElement 0FIGURE <strong>36</strong>.32 Modeling the viscoelastic behavior <strong>of</strong> wood with the number <strong>of</strong> Kelvin elements in series.significant section deformations can be observed experimentally(Brandã o and Perré , 1996). Figure <strong>36</strong>.34is an example <strong>of</strong> nonsymmetrical drying. In order toincrease the section deformation, a thin quartersawnboard has been simulated, here 5 mm by 80 mm. Thelarge displacement formulation is essential in thiscase. In this configuration, one part <strong>of</strong> the dryingstress is transformed into section deformation; hencethe stress-reversal phenomenon induces a negativefinal curvature.Thickness (cm)1.00.80.60.40.2Moisture content0.650.580.510.440.370.290.220.15Sigma xx0.00.00.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0Width (cm)Width (cm)Thickness (cm)1.00.80.60.40.21.541.240.930.630.320.01−0.29−0.60Thickness (cm)1.00.80.60.40.2Sigma xy0.00.0 0.5 1.0 1.5 2.0(a)Width (cm)0.150.090.03−0.03−0.09−0.15−0.21−0.28Thickness (cm)1.00.80.60.40.2Sigma yy0.00.0 0.5 1.0 1.5 2.0Width (cm)1.291.040.790.530.280.03−0.22−0.47Thickness (cm)Moisture content1.00.80.60.40.20.00.0 0.5 1.0 1.5 2.0Width (cm)0.080.080.070.070.070.070.070.07Thickness (cm)1.00.80.60.40.2Sigma xx0.00.0 0.5 1.0 1.5 2.0Width (cm)0.12−0.29−0.70−1.11−1.52−1.93−2.34−2.75aThickness (cm)1.00.80.60.40.2Sigma xy0.240.210.180.150.120.090.060.03Thickness (cm)1.00.80.60.40.2Sigma yy−0.10−0.35−0.60−0.86−1.11−1.<strong>36</strong>−1.62−1.870.00.0 0.5 1.0 1.5 2.0(b)Width (cm)0.00.0 0.5 1.0 1.5 2.0Width (cm)FIGURE <strong>36</strong>.33 Moisture content and stress fields after (a) 6 h and (b) 60 h (T d ¼ 808C, T w ¼ 608C).ß 2006 by Taylor & Francis Group, LLC.


7 hMoisture: 0.05 0.09 0.14 0.18 0.22 0.26 0.31 0.3530 hFIGURE <strong>36</strong>.34 Example <strong>of</strong> nonsymmetrical drying: moisture content and section deformation (80/608C).The next example illustrates three different constantdrying conditions chosen to analyze the possibilities<strong>of</strong> prediction (T d stands for dry-bulb temperatureand T w for wet-bulb temperature):1. T d ¼ 40 8C, T w ¼ 358C; mild conditions at lowtemperature (EMC ¼ 15%)2. T d ¼ 80 8C, T w ¼ 60 8C; rather severe conditionsat medium temperature (EMC ¼ 7%)3. T d ¼ 80 8C, T w ¼ 76 8C; mild conditions atmedium temperature (EMC ¼ 14%)Figure <strong>36</strong>.35 depicts the variations <strong>of</strong> the averagedmoisture content and the stress level (direction parallelto the exchange surface) at different positions vs.time. All tests show a first drying period, withoutdrying stress, then a stage with tensile stress in theperipheral zones, and finally the last drying stage thatexhibits the stress reversal. However, the duration <strong>of</strong>each stage and the stress level depend strongly on thedrying conditions.For the mild drying conditions at low temperature(T d ¼ 408C, T w ¼ 35 8C), the drying time is ratherimportant.Thefirstdryingperiodlastsforaround10hand the stress level is high for both the second dryingstage and the final drying stage. The drying conditionsare mild concerning heat and mass transfer,while the temperature level is not high enough forthe creep field to relax the stress field; the final stresslevel reveals the importance <strong>of</strong> the memory effect.As a consequence <strong>of</strong> the low relative humidity <strong>of</strong>air, the second test (T d ¼ 80 8C, T w ¼ 60 8C) is veryfast. However, both the maximum tensile stress leveland the final stress reversal are important; the rapidexternal transfer imposes a high moisture contentgradient within the board. In addition, the viscoelasticcreep is not sufficient to cancel the memory effect.At the beginning <strong>of</strong> drying, the board temperature isclose to the wet-bulb temperature, which is below theglass-transition zone (Ge<strong>of</strong>fray 1984, Goreng 1963,Salmen 1984, Ostberg et al. 1990, Passard and Perre2001). At the end <strong>of</strong> drying, the temperature level issufficient for greenwood, but not for the dry part <strong>of</strong>the board to activate the viscoelastic behavior. Consequently,the outer parts, which are close to EMC,are below the glass-transition zone.In the third test (T d ¼ 80 8C, T w ¼ 76 8C), thedifference in moisture content between surface andcore remains low. The first drying period lasts animportant part <strong>of</strong> the total drying time. Due to thehigh value <strong>of</strong> EMC, the board temperature is alwaysabove the s<strong>of</strong>tening zone; consequently, all stresslevels remain very low. These conditions allow wood<strong>of</strong> good quality to be obtained relatively free <strong>of</strong> stressreversal with a moderate drying time (less than 150 hagainst 400 h for the low-temperature test).These simulations are in good agreement withnonsymmetrical drying experiments performed onoak (Quercus rubra) boards using the same dryingconditions (Figure <strong>36</strong>.<strong>36</strong>; Perré , 2001). However, aß 2006 by Taylor & Francis Group, LLC.


Averaged moisture content (%)907560453015MoisturecontentSurface5 mm10 mmCenter1.51.00.50−0.5−1.0s xx (MPa)00 100 200 300 400 −1.5(a)Time (h)Averaged moisture content (%)907560453015MoisturecontentSurface5 mm10 mmCenter1.51.00.50−0.5−1.0s xx (MPa)0(b) 0 50 100 150 −1.5Time (h)Averaged moisture content (%)907560453015MoisturecontentSurface5 mm10 mmCenter1.51.00.50−0.5−1.0s xx (MPa)(c)00 50 100 150 200 −1.5Time (h)FIGURE <strong>36</strong>.35 Averaged moisture content and s xx vs. time: (a) T d ¼ 408C, T w ¼ 358C; (b) T d ¼ 808C, T w ¼ 608C; and(c) T d ¼ 808C, T w ¼ 768C.closely similar schedule for another hardwood(Noth<strong>of</strong>agus truncata) resulted in gross deformationsand thermal degradation due to the high extractivescontent <strong>of</strong> the wood (Grace, 1996).Based on this reasoning, new drying procedureshave been devised and tested on different tropicalspecies, including numerous tests in industrial kiln <strong>of</strong>100-m 3 capacity (Aguiar and Perré , 2000b). The productquality was always very good, <strong>of</strong>ten with very littlechecking and rather less deformation than withconventional drying. Most importantly, this methodneeds only one half to one third <strong>of</strong> the time requiredfor drying according to conventional schedules.The code can also be used to test different dryingschedules (Figure <strong>36</strong>.37). The first one (Schedule A) isrecommended for s<strong>of</strong>twoods while the second oneß 2006 by Taylor & Francis Group, LLC.


FIGURE <strong>36</strong>.<strong>36</strong> Final section curvature <strong>of</strong> oak samples dried on their upper face with conditions n 2 (T d ¼ 808C, T w ¼608C) and n3 (T d ¼ 808C, T w ¼ 768C), respectively.(Schedule B) is recommended for hardwoods. ScheduleB proposes lower temperature levels and higherrelative humidity values. In this drying schedule,EMC decreases significantly only at the end <strong>of</strong> theprocess, when the board is supposed to be dry with alow moisture content gradient. As the first consequence<strong>of</strong> these drying conditions, one can notice amuch longer drying time for Schedule B (130 hagainst 50 h). The first drying period lasts also for alonger time for the second procedure (30 h instead <strong>of</strong>10 h). It may be noted that the first drying–periodduration represents about the same percentage <strong>of</strong> thetotal drying time for both schedules. This remark stillstands for the stress level. One can easily consider arelative time (current time over total drying time) atwhich all curves have the same shape and the samestress magnitude.In the first example, the advantages to be gainedfrom using a high relative humidity level (low moisturecontent gradient, high hygroactivation <strong>of</strong> theviscoelastic behavior) hardly <strong>of</strong>fset the negative effect<strong>of</strong> the low temperature levels (slow moisturemigration and low thermoactivation <strong>of</strong> the viscoelasticbehavior). A careful analysis <strong>of</strong> these approachesis very promising. New rules can bederived to improve existing drying schedules or todevise innovative drying procedures.Schedule AAverageMoistureContent (%)Dry-BulbTemperature(8C)Wet-BulbTemperature(8C)RelativeHumidity(%)Green 71 66 8050 76.5 68.5 7030 82 70.5 6020 88 67.5 40Schedule BAverageMoistureContent (%)Dry-BulbTemperature(8C)Wet-BulbTemperature(8C)RelativeHumidity(%)Green 40.5 38 8560 40.5 37 8040 43.5 39 7535 43.5 38 7030 46 39.5 6525 51.5 43 6020 60 47.5 5015 65.5 49 40ß 2006 by Taylor & Francis Group, LLC.


Moisture content (%)12010080604020Average moisturecontent(%)Green503020Dry-bulbtemperature(8C)7176.58288Wet-bulbtemperature(8C)6668.570.567.5Temperature edgeTemperature centerMC edgeMC centerMC surfaceMC averageRelativehumidity(%)80706040908070605040Temperature (8C)Moisture content (%)10080604020Average moisturecontent(%)Green60403530252015Dry-bulbtemperature(8C)40.540.543.543.54651.56065.5Wet-bulbtemperature(8C)3837393839.54347.549Relativehumidity(%)858075706560504070Temperature edgeTemperature centerMC edgeMC centerMC surfaceMC average60504030Temperature (8C)00 20Time (h)40 60 302Surface5 mm110 mmCenter00 50Time (h)100 150 202Surface5 mm110 mmCenters xx (MPa)0s xx (MPa)0−1−1−20 20 40 60Time (h)Schedule A−20 25 50 75 100 125 150Time (h)Schedule BFIGURE <strong>36</strong>.37 Simulation <strong>of</strong> two different drying schedules: moisture content, temperature, and stress level at differentpositions vs. time.An alternative procedure in improving kiln scheduleshas beenthe estimation<strong>of</strong> strainlevels toprovideasafe envelope <strong>of</strong> dry- and wet-bulb temperatures in kilnoperation. One industrial method uses acoustic emissionson sample boards to determine a stress thresholdto keep the surface strain under 50 to 75% <strong>of</strong> the estimatedultimate value (Doe et al., 1996b). Later optimizedschedules have been developed by Langrish et al.(1997) using a model predictive control technique tokeep within the strain criterion. The technique reducedthe number <strong>of</strong> small- and medium-sized cracks, bothinternallyandatthesurface,toless than one quarter <strong>of</strong>those observed in the original conventional schedule.<strong>36</strong>.2.5 DRYING QUALITY<strong>36</strong>.2.5 .1 F actors Affecti ng the Dry ing Dura tionLet us assume that the duration <strong>of</strong> drying is the singleimportant factor. In this case, heat and mass transferonly are to be considered (Figure <strong>36</strong>.38). Some parametersdepending on the load are important, but they arenot under control:. The thickness is a very important parameter(roughly speaking, the drying time increases asthe thickness doubled). The transfer properties <strong>of</strong> the wood (diffusivity,permeability, capillary pressure, thermal conductivity,etc.)In conventional drying, the controlled parametersare the dry- and wet-bulb temperatures as well as thevelocity <strong>of</strong> the airflow. These three parameters determinethe external heat- and mass-transfer rates:. The ‘‘drying potential’’ <strong>of</strong> the air flow is theheat-transfer coefficient (which increases withthe air velocity) times the difference betweendry- and wet-bulb temperatures.ß 2006 by Taylor & Francis Group, LLC.


ParametersLow valueHigh valueExternalVelocity<strong>Drying</strong> potential(dry bulb−wet bulb)ThicknessInternalMass diffusivityThermal diffusivityTemperatureFIGURE <strong>36</strong>.38 Guidelines on how to obtain a fast drying operation (from good to excellent and from poor todisastrous). (Adapted from Perré, P., The drying <strong>of</strong> wood: the benefit <strong>of</strong> fundamental research to shift fromimprovement to innovation, Proceedings <strong>of</strong> the Seventh International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Tokyo, Japan, 2001,2–13.). The air velocity also plays an important role inuniform drying within the stack. However, itseffect becomes less important as the drying progresseswhen internal transfer mainly controlsthe moisture migration.In addition, we have to keep in mind some moresubtle effects:. The internal transfer (diffusion, liquid migration)becomes easier when the temperaturelevel increases.. Above the boiling point <strong>of</strong> water, an additionaldriving force, the gradient <strong>of</strong> total pressure, actswith a dramatic effect (Perré , 1995). <strong>Drying</strong>by internal vaporization takes place in suchconditions.. The internal transfer rates depend on the localmoisturecontent(liquidmigrationisusuallymuchmore effective than bound or vapor diffusion).In addition, diffusion becomes very slow whenthe bound-water content decreases toward zero.In general, the drying time is reduced when thevelocity and the temperature <strong>of</strong> air are high and itsrelative humidity is low. However, an excessively lowrelative humidity may produce a surface zone with lowmoisture content, thus reducing moisture migrationclose to the surface. All high-temperature arrangements(convective drying at high temperature, vacuumdrying, contact drying, etc.) are processes that accelerateinternal moisture migration due to the overpressuresgenerated within the product.Finally, drying with electromagnetic heating(microwave or radio frequency) <strong>of</strong>fers an entirelynew possibility: any internal temperature can beattained without resorting to a heating medium suchas a hot gas.<strong>36</strong>.2.5 .2 Fac tors Affecti ng the Dry ing Qual ity<strong>Drying</strong> stresses originate from shrinkage; as soon asthe shrinkage field within the board is not geometricallycompatible, a stress field develops in the material,which is responsible for mechanical degradation. Inorder to reduce the stress level throughout the process,and thereby the surface checking, the internalchecking, and the residual stress, several conditionsshould be fulfilled (Figure <strong>36</strong>.39):. Low shrinkage coefficients, not under control. Small thickness, not under control. Low moisture content values between surfaceand core. Retaining important possibilities <strong>of</strong> viscoelasticcreep (mechanosorptive creep is always a source<strong>of</strong> stress reversal); such an effect is obtained athigh temperatures, provided the moisture contentis sufficiently high (Irvine, 1984)It may be noted that a low-temperature level is sometimesdesired (for example to avoid collapse), becausea high-temperature level may produce thermal degradationor discoloration.<strong>36</strong>.2.5 .3 Cr iteria for Obtaini ng a Fast and GoodDry ing Pro cessA fast and good drying process should incorporate thecriteria listed in Section <strong>36</strong>.2.5.1 and Section <strong>36</strong>.2.5.2,ß 2006 by Taylor & Francis Group, LLC.


Second drying periodTensionParametersLow valueHigh valueCompressionThicknessTensionEnd <strong>of</strong> dryingCompressionMoisture content(bound) gradient(center−surface)ShrinkageTensionCompressionSurface moisturecontentTemperatureFIGURE <strong>36</strong>.39 Guidelines on how to obtain a good-quality product (from good to excellent and from poor todisastrous). (Adapted from Perré, P., The drying <strong>of</strong> wood: the benefit <strong>of</strong> fundamental research to shift fromimprovement to innovation, Proceedings <strong>of</strong> the Seventh International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Tokyo, Japan, 2001,2–13.)which are, to a large extent, contradictory. Numerousmechanisms involved during drying have to be considered(Figure <strong>36</strong>.40).Because <strong>of</strong> this complexity, compromises have tobe found. Nevertheless, some general rules can belisted (Figure <strong>36</strong>.41):. Rule 1: high relative humidit To y. ensure a lowmoisture content gradient, one way is to reducethe drying potential (wet-bulb depression) asmuch as possible. In addition, this conditionimposes a relatively high value <strong>of</strong> EMC (onlyone part <strong>of</strong> shrinkage is effected and the influence<strong>of</strong> temperature on the viscoelastic creep isnot inhibited by a relatively low moisture contentlevel). However, a high relative humidityvalue can activate the development <strong>of</strong> fungi.. Rule 2: high tem perature. A high value <strong>of</strong> temperatureis most <strong>of</strong>ten a positive factor. Thisaccelerates the internal moisture transfer andactivates the viscoelastic creep. However, careshould be taken with sensitive species; high temperaturelevels can increase the risk <strong>of</strong> collapse,problems <strong>of</strong> color, or even thermal degradation<strong>of</strong> the wood constituents.. Rule 3: high air velocity. A high air velocitypromotes good uniformity <strong>of</strong> drying throughoutthe stack. However, a higher velocity increasesthe electricity consumption and may produce,by the heat-transfer coefficient, an excessivelyhigh external transfer flux, which is opposite tothe effect intended in Rule 1.Concerning moisture transfer, Rule 1 and Rule 2meanthatinternaltransferhastobeincreasedwhereasexternal transfer should be reduced. Exceeding theboiling point <strong>of</strong> water is decisive for internal transfer.However, these rules also require the temperature tobe high with a high value <strong>of</strong> relative humidity. Suchconditions may be difficult to ensure for certaindryers. Innovative drying procedures may need newdryer designs. Finally, too <strong>of</strong>ten, the effect <strong>of</strong> temperatureand moisture content on the viscoelasticbehavior is disregarded in the optimization <strong>of</strong> dryingschedules. The situation strongly differs from onespecies to the other. Usually, s<strong>of</strong>twood species arequite easily dried. On the other hand, hardwoods are<strong>of</strong>ten intractable because <strong>of</strong> their low permeability.<strong>36</strong>.3 KILN SCALE<strong>36</strong>.3.1 LUMBER QUALITYThe ultimate fitness for the purpose <strong>of</strong> dried lumberdepends not only on the chosen drying conditions butalso on the lumber quality itself. This quality may bethought<strong>of</strong>in terms <strong>of</strong> gross defects suchas knotsaswellasintrinsicwoodpropertiessuchasthedegree<strong>of</strong>anisotropy.<strong>Drying</strong>,whichcausesanisotropicshrinkage,interactswith various wood features in various ways. Theobjective <strong>of</strong> kiln seasoning, then, is to acknowledge thisinteraction by setting process conditions that yield driedlumber to the specifications in terms <strong>of</strong> a grade for anend use. There is a world <strong>of</strong> difference between dryingdecorative hardwoods and drying structural s<strong>of</strong>twoods.Increasingly, kiln operators are drying wood fromever younger, fast-growing stands rather than from mature,old-growth forest. The drying behavior <strong>of</strong> this newkind <strong>of</strong> wood is requiring operators to adapt traditionalprocesses on the basis <strong>of</strong> better understanding <strong>of</strong> thedrying mechanism, as outlined in the previous sections.ß 2006 by Taylor & Francis Group, LLC.


Transition zone0.55Relative density at breast height0.500.450.400.350.300.25Douglas-firWestern larchWestern hemlockYellow - cypressLodgepole pineSitka spruceInterior spruceSubalpine firWestern red cedar0.200Juvenile wood10 20 30Age (y)Mature wood40 50 60FIGURE <strong>36</strong>.43 Trends in basic density at breast height for commercial second-growth s<strong>of</strong>twoods in British Columbia.(Adapted from Josza, L.A. and Middleton, G.R., A discussion <strong>of</strong> wood quality attributes and their practical implications,Forintek Canada Special Publ. SP-34, 1994.)In practice, however, it is not easy to distinguish theeffect <strong>of</strong> density with that <strong>of</strong> moisture content.Within-ring density variations can cause problemsbecause <strong>of</strong> the differential shrinkage at the ringboundary. Subsequently, severe drying stresses maycause deformation and internal checking (Booker,1994). Further, a low absolute density in earlywoodcan result in collapse on drying, particularly underhigh-temperature conditions (Booker, 1996).Within-tree variations in density can be highlysignificant. Cown and McConchie (1983) show thatthe density in a 24-year-old radiata pine tree can varyfrom 300 kg m 3 in the top log to greater than 450 kgm 3 in the outer wood <strong>of</strong> the butt log. Consequently,the drying kinetics <strong>of</strong> boards taken from the same logmay be markedly different (Davis, 2001). Trends inbasic density for a number <strong>of</strong> second-growth s<strong>of</strong>twoodsare illustrated in Figure <strong>36</strong>.43.The situation is more complex with hardwoods.The growth rate has little effect on the wood properties<strong>of</strong> diffuse-porous hardwoods, but has a markedimpact on the density <strong>of</strong> ring-porous hardwoods. Unlikes<strong>of</strong>twoods, these produce denser wood when fastgrown.Regardless <strong>of</strong> the species or where the forests areestablished, the variation in wood properties betweentrees is very great and can be great even in boardssawn from the same tree. In particular, the socialstatus <strong>of</strong> the tree in the stand (whether dominatedor dominant trees) has a great effect on the growthrate in diameter and the occurrence <strong>of</strong> reactionwood. Table <strong>36</strong>.10 lists some <strong>of</strong> the characteristics <strong>of</strong>TABLE <strong>36</strong>.10Characteristics and Properties <strong>of</strong> Reaction <strong>Wood</strong> Compared with Corresponding Normal <strong>Wood</strong>Features Compression <strong>Wood</strong> Tension <strong>Wood</strong>Physical characteristics Darker in color and very hard Darker in color and silvery sheen in mosttemperate hardwoodsDensity 10–100% greater 10–30% greaterLongitudinal shrinkage Order <strong>of</strong> magnitude greater (up to several fold) About fivefold greaterWarp on drying Liable to warp badly Can warp and is liable to collapseStrength Comparable strength, does not reflect higher density Superior strengthSource: Keey, R.B., Langrish, T.A.L., and Walker, J.C.F., The Kiln-<strong>Drying</strong> <strong>of</strong> Lumber, Springer, Berlin, 2000.ß 2006 by Taylor & Francis Group, LLC.


eaction wood produced by environmental factors inthe forest. Compression wood has been found to havea significant effect on drying, with lower drying ratesover the moisture content range <strong>of</strong> 40 to 100% forboards <strong>of</strong> P. radiata (Davis et al., 2002). This reductionis attributed to the lower permeability <strong>of</strong> thecompression-wood zone, with its denser wood andmore thick-walled cells.Large micr<strong>of</strong>ibril angles are found in both compressionwood and normal wood near the pith, inducinglarger than normal longitudinal shrinkage and agreater tendency to warp. The longitudinal shrinkage<strong>of</strong> tension wood, although larger than normal wood,is less than in compression wood.<strong>36</strong>.3.1 .3 Sa wmilling Strategi esTimell (1986) describes sawmilling strategies to reducewarp in juvenile lumber and compressionwood. The cutting pattern influences the quality <strong>of</strong>drying either by releasing stresses on sawing or thecutting induces a stress pattern in the board that canbe balanced on drying. Warp has become more <strong>of</strong> aproblem with harvesting from second-growth forests<strong>of</strong> short rotation, which contain proportionatelymore juvenile wood than lumber from old-growthforests. Crook and bow are most severe in the corewood <strong>of</strong> the butt log and where compression wood isencountered; whereas twist is most severe in the corewood <strong>of</strong> the upper logs where spiral-grain angles arelarge and changing rapidly. The variation <strong>of</strong> propertiesabout the mean is the critical factor with core wood.Vá zquez (2001) examines silvicultural practiceand sawmilling strategies to counteract the effect <strong>of</strong>growth stresses in fast-grown Iberian eucalypts (Eucalyptusglobulus). A model <strong>of</strong> the stress distributionthat enables to determine the deformations on sawingis given. The appearance <strong>of</strong> checks and warps can belimited by the choice <strong>of</strong> a suitable sawing pattern, asshown in Figure <strong>36</strong>.44. Pang and Haslett (2002) notethat the residual drying stresses in quartersawnP. radiata boards are less than flatsawn boards, particularlyunder low-temperature drying conditionswhen the effect <strong>of</strong> mechanosorptive stress relief isrelatively minor. The difference in behavior is attributedto the lesser shrinkage in the width direction withquatersawn boards.<strong>36</strong>.3.2 KILN DESIGNAlthough there are differences in detail betweenmanufacturers, a lumber kiln is essentially a specialpurposeroom fitted with overhead fans for circulatingthe drying air and heating coils for maintainingthe air (and thus the wood) temperature at the setlevels. The moisture in the air is controlled by means<strong>of</strong> opening the vents in the kiln’s ro<strong>of</strong>, thus governingthe amount <strong>of</strong> moist air that returns to the fan whichis to be mixed with the fresh air drawn in. Althoughmany kilns are operated batchwise for ease <strong>of</strong> controllingthe drying conditions, which may changethrough out the drying schedule, a kiln can be continuouslyworked by arranging the lumber to beslowly railed through the chamber. In this lattercase, the drying schedule is maintained by varying thetemperature and humidity settings along the length <strong>of</strong>the kiln. There is some increase in interest in continuouskilns, which might provide energy savings andbetter quality control (McLean, 2003).The lumber is stacked externally in a rectangularpile on a low, flat-bed trolley, with rows <strong>of</strong> boardsseparated by wooden stickers <strong>of</strong> uniform thickness toprovide duct-like spaces between the boards for thekiln air to flow through. The boards may be stackedFIGURE <strong>36</strong>.44 Sawing pattern to limit the appearance <strong>of</strong> checks. (Adapted from Vázquez, M.C.T., Tensiones de crecimiento enEucalyptus globulus de Galicia (España). Influencia de la silvicultura y estrategias de aserrado (Growth stresses in E. globulusfrom Galicia (Spain). Influence <strong>of</strong> silviculture and sawing strategies), Maderas: Ciencia Tecnologia, 2(1), 68–89, 2001.)ß 2006 by Taylor & Francis Group, LLC.


in separate packages on bearers and loaded into thekiln by a forklift vehicle. The boards are butted up,with their long faces incident to the airflow. The stackis squared <strong>of</strong>f as far as possible to provide a uniformresistance to the airflow, and thus minimize variationsin drying throughout the kiln. In Scandinavian practice,kiln stacks are normally built from boards <strong>of</strong>random length, so that every second board is placedflush at one end <strong>of</strong> the stack and the other boards flushat the other end (Salin, 2001). Kilns may be singletrackedwith a 2.4-m wide stack, or twin-tracked withtwo stacks side by side, to yield a double-width stack<strong>of</strong> 4.8 m. Figure <strong>36</strong>.45 illustrates a vertical cross sectionthrough a single-tracked, batch kiln.To obtain a uniform air distribution to the airinletface <strong>of</strong> the stack as possible, the plenum spacesat each side <strong>of</strong> the stack must be sufficiently wide(Nijdam and Keey, 2000). An internal ceiling directsthe air through the lumber stack or stacks, with bafflesor curtains to direct the airflow through the lumberpile. Inward-swinging baffles and contoured,right-angled bends to the plenum space from theceiling zone improve the uniformity <strong>of</strong> this airflow(Nijdam and Keey, 2002). Bypass <strong>of</strong> air around thestack is minimized by the fitting <strong>of</strong> side baffles orcurtains. The kiln is designed so that the pressureloss through the heating coils and other ancillaryfittings is small compared with that through thestack <strong>of</strong> lumber.<strong>36</strong>.3.2.1 Airflow ConsiderationsKröll (1978) reports the air distribution in a batchdryer fitted with 15 shelves, a heat exchanger, and afan above a false ceiling, which thus resembles a boxtypelumber kiln in a number <strong>of</strong> respects. In the fifthgap from the top, the air velocity was over twice theaverage, whereas in the top gap there was a smallbackflow with air streaming toward the inlet. Theflow reversal appeared to be the result <strong>of</strong> a vortexgenerated at the top <strong>of</strong> inlet plenum below the halfcirclebend out <strong>of</strong> the ceiling space. Although such airmaldistribution may be extreme in a modern lumberkiln, existing kiln designs may still yield a nonuniformairflow through well-stacked lumber loads, as may beinferred from the kiln audits reported by Nijdam andKeey (1996). Haslett (1998) recommends that the coefficient<strong>of</strong> variation across the outlet face <strong>of</strong> the stackshould not exceed 0.12 at velocities through the stackbetween 4.5 and 8 m s 1 when the dry-bulb temperatureis greater than 908C.Industrial rules <strong>of</strong> thumb generally equate theceiling-space height to the plenum-space width andto the combined sticker-spacing height. Hydraulictests on a model kiln confirm the former rule(Nijdam and Keey, 1999), and the latter is verifiedby a pressure-drop analysis (Nijdam, 1998).The transverse gaps between the boards are notsimple smooth ducts, but there may be irregularitiesRo<strong>of</strong> ventHeater coilsCeilingspaceReversible fanWeightPlenumchamberPlenumchamberFilletFIGURE <strong>36</strong>.45 A vertical cross section through a single-tracked, box kiln. (Adapted from Keey, R.B., Langrish, T.A.L., andWalker J.C.F., The Kiln-<strong>Drying</strong> <strong>of</strong> Lumber, Springer, Berlin, 2000.)ß 2006 by Taylor & Francis Group, LLC.


where the boards butt up due to shrinkage on dryingor unevenness in thickness and there may be gaps dueto the presence <strong>of</strong> boards with uneven length. Fluiddynamicsimulation <strong>of</strong> the flow over inline slabs(Langrish et al., 1993) has suggested that gaps assmall as 1 mm might be sufficient to disrupt theflow, with circulation within the gaps themselves.The magnitude <strong>of</strong> the side gaps in the board influencesboth the drying rates and the development <strong>of</strong>drying stresses (Langrish, 1999), so that experimentson the drying <strong>of</strong> single boards (as <strong>of</strong>ten done) mayyield uncertain information about the drying <strong>of</strong> a load<strong>of</strong> the same wood in a kiln. With regard to variationsin board thickness, Haslett (1998) recommends thatthe coefficient <strong>of</strong> variation for the board thicknessmust be under 0.04 for successful high-temperaturedrying.Whenever kiln stacks are built from randomlengthlumber so that every second board is flush ateach end <strong>of</strong> the stack, variations in openness <strong>of</strong> thestack result. This gives two different zones: (1) acentral zone in which all the available space is filledand (2) two end zones where alternate boards aremissing (Salin, 2001). This arrangement results inhigher within-stack velocities (about 30% higher) inthe center than in the end zones, with correspondingimplications in the variation in drying behavior.<strong>36</strong>.3.2.2 Moisture-Evaporation ConsiderationsThe airflow through the stack influences the magnitude<strong>of</strong> the local airside mass-transfer coefficient, andthus the evaporation into the airstream. Particularly,at the higher air velocities used in high-temperaturedrying, any variations in these transfer coefficientshave a significant effect on the uniformity <strong>of</strong> dryingthroughout the stack.The air-inlet face <strong>of</strong> the lumber stack presents a set<strong>of</strong> blunt edges to the incident airflow, resulting in anenhancement <strong>of</strong> the mass-transfer coefficients nearthe leading edges (Kho et al., 1990). Computationalstudies (Sun, 2001) <strong>of</strong> the flow over a series <strong>of</strong> slabswith inline gaps suggest that for gaps greater thanabout 2 mm there will be similar, but lesser, enhancementsat subsequent boards downstream.With Scandinavian stacking practice, the endzones <strong>of</strong> the stack dry faster than the central, fullyfilled part (Salin and Öhman, 1998). The lower airvelocity in the ends is more than compensated byhigher heat-transfer coefficients associated with theflow disturbance and smaller wood volume. (Thereis a smaller decrease in temperature and increase inhumidity along the stack in the airflow direction.) Ingeneral, it is expected that the local transfer coefficientsdiminish with distance in the airflow directiondue to a thickening <strong>of</strong> the boundary layer. This variationand the downwind accumulation <strong>of</strong> moisture inthe airstream result in the maximum possible evaporationrate dwindling with distance from the air inlet tothe air outlet from the stack.Traditionally, the variation <strong>of</strong> evaporative ratesacross the stack has been counteracted by the installation<strong>of</strong> bidirectional fans and by periodically reversingthe airflow direction through the stack. Thispolicy has minimal effect on the drying rates in thecenter <strong>of</strong> the stack, but reduces the variation in behaviorbetween the two end zones. If only moisturecontent variations are considered, many reversals arenot needed to achieve this equalization (Pang et al.,1995; Nijdam and Keey, 1996; Wagner et al., 1996).However, if stress development in the surface layerwith the likelihood <strong>of</strong> checking is taken into account,then the flow reversals for a timber such as Pinussylvestris should be less than 2 h apart (Salin andÖhman, 1998). A period <strong>of</strong> 4 h is a common industrialpractice for permeable s<strong>of</strong>twoods such as P. radiata.<strong>36</strong>.3.3 KILN OPERATIONTo understand kiln-wide behavior, it is useful to invokethe concept <strong>of</strong> the characteristic drying curve(van Meel, 1958; Keey, 1978). The concept reducesthe drying kinetics for a specific material <strong>of</strong> specificgeometry to a single function <strong>of</strong> the local averagedmoisture content. The concept when applied to thekiln drying <strong>of</strong> lumber boards is rough, not only due tovariations in drying behavior between boards (Daviset al., 2001) but also due to embedded assumptions inthe concept itself. Nevertheless, it is a sufficient representation<strong>of</strong> drying behavior to determine the effect<strong>of</strong> kiln parameters on the course <strong>of</strong> drying. Thesethings, such as the uniformity <strong>of</strong> the airflow, thenumber <strong>of</strong> airflow reversals, the velocity, temperature,and humidity settings, are all those under thecontrol <strong>of</strong> the kiln operator.The concept <strong>of</strong> a characteristic drying curve leads tothe following expression for the moisture-evaporationrate per unit <strong>of</strong> exposed board surface:N v ¼ f bf(Y W Y G ) (<strong>36</strong>:22)where f is the evaporation rate relative to that at agiven moisture content (either the initial or some criticalvalue <strong>of</strong> transition from unhindered drying), and isa unique function <strong>of</strong> the mean free moisture content;b is the external (airside) mass-transfer coefficient; f isthe humidity-potential coefficient, which takes a constantvalue when the wet-bulb temperature remains thesame throughout the kiln; Y W is the saturation humidityat the wet-bulb temperature; and Y G is the bulk-airß 2006 by Taylor & Francis Group, LLC.


humidity between the boards. Although the values <strong>of</strong>the parameter f depend on the extent <strong>of</strong> drying that hastaken place and the nature <strong>of</strong> the wood itself, the otherparameters are under the control <strong>of</strong> the kiln operatorby varying either the stack extent, the kiln-air velocity,or the dry- and wet-bulb temperatures.Consideration (Keey et al., 2000) <strong>of</strong> the moisturetransfer over a small zone in the kiln leads to the equations,which can be conveniently expressed in nondimensionalform as follows:∂Φ∂qInletOutletz@F@u ¼ @P ¼ f P (<strong>36</strong>:23)@zwhere F is the moisture content relative to unit valuewhen f is 1, P is the humidity potential (Y W Y G )relative to unit value at the air inlet, z is a nondimensionalextent <strong>of</strong> the kiln in the airflow direction andis a weak function <strong>of</strong> the kiln-air velocity, and u isthe relative time <strong>of</strong> drying which itself depends uponthe value <strong>of</strong> z for the kiln stack and the capacity <strong>of</strong> theair to pick up moisture. These equations can be solvedif the parameter f is known as a function <strong>of</strong> themoisture content (averaged over the board thickness).They imply that the rate <strong>of</strong> change <strong>of</strong> moisture contentwith time (i.e., the drying rate) directly dependson the rate at which the bulk air humidifies in itspassage through the kiln. The drying rate is alsodirectly dependent upon the humidity potential (thedriving force for the evaporation) and the parameterf, which reflects the ease <strong>of</strong> moisture movementthrough the wood.These equations have been used to examine theinfluence <strong>of</strong> kiln variables on the course <strong>of</strong> drying,including the impact <strong>of</strong> exhaust-air recycle and theswitching <strong>of</strong> airflow direction (e.g., Tetzlaff, 1967;Ashworth, 1977; Ashworth and Keey, 1979; Keeyand Pang, 1994; Nijdam and Keey, 1996). A summary<strong>of</strong> this work is given by Keey et al. (2000).Figure <strong>36</strong>.46 shows the variation in the dimensionlessdrying rate, @F/ @u, as a function <strong>of</strong> the normalizedmoisture content F for one-way flow througha single-tracked kiln, for which the nondimensionalextent z in the airflow direction is 1. In the case <strong>of</strong> atimber, whose initial green moisture content is equalto the critical point <strong>of</strong> transition between unhinderedand hindered drying by the rate <strong>of</strong> moisture movementthrough the wood, the drying rate falls monotonicallywith both time and distance in the airflowdirection. The effect with time is due to the intrinsicdrying rate as the wood dries out, whereas that withdistance is due to the progressive humidification inthe kiln. Whenever there is free moisture content inthe wood above the critical point, the drying-ratepr<strong>of</strong>iles are more complex. As soon as the critical(a)∂Φ∂q(b)0InletΦOutlet0 Φ1point is reached at the air-inlet face <strong>of</strong> the stack, theintrinsic drying rate falls there, resulting in less progressivehumidification in the kiln; the downstreamdrying rates can now rise until the local critical pointis attained. The greatest effect is seen at the outlet face<strong>of</strong> the stack.Figure <strong>36</strong>.47 shows the effect <strong>of</strong> reversing the airflowdirection through the stack. On switching over theflow, what was once the ‘‘inlet’’ face now becomes the‘‘outlet’’ face, and vice versa, giving a temporary boostto the drying rate at the former outlet and a moderationto that at the former inlet. The rates within the center<strong>of</strong> the kiln are essentially unaffected. With flowswitchovers, the leaflike moisture content pr<strong>of</strong>iles becomemore pinched with a lesser variation in moisturecontent across the kiln.z1 = Φ 0Φ 0FIGURE <strong>36</strong>.46 Normalized drying rates in a kiln with anextent z <strong>of</strong> 1 and one-way airflow as a function <strong>of</strong> boardaveragedmoisture contents: (a) an indicative hardwood withf ¼ F, (b) an indicative sapwood <strong>of</strong> a s<strong>of</strong>twood with f ¼ F 0.5 .(Adapted from Keey, R.B., Langrish, T.A.L., and Walker,J.C.F., The Kiln-<strong>Drying</strong> <strong>of</strong> Lumber, Springer, Berlin, 2000.)ß 2006 by Taylor & Francis Group, LLC.


1.2Normalized moisture content (Φ)1.00.80.60.40.2Air-outlet faceMiddleAir-inlet faceNormalized drying rate (dΦ/dq)1.0Air-inlet face0.8Middle0.6Air-outlet face0.40.2(a)000100 200 300 400 500 0 100 200 300 400 500Normalized time (q)Normalized moisture content (Φ)1.21.00.80.60.40.200(b)OutletMiddleInletNormalized drying rate (dΦ/dq)1.00.80.60.40.2InletMiddleOutletInletOutlet0100 200 300 400 500 0 100 200 300 400 500Normalized time (q)FIGURE <strong>36</strong>.47 Board-average moisture contents and normalized drying rates as a function <strong>of</strong> time and distance in theairflow direction for a twin-stack kiln without reversals in the schedule (a) and for two flow reversals in the schedule (b). Thepr<strong>of</strong>iles represent the drying <strong>of</strong> 100 50-mm sapwood Pinus radiata dried at 77/65.58C and an air velocity <strong>of</strong> 2.5 m s 1 .(Adapted from Tetzlaff, A.R., An investigation <strong>of</strong> drying schedules when kiln-drying radiata pine, B.E. Report, University <strong>of</strong>Canterbury, New Zealand, 1967.)Typically, lumber kilns operate at very high ratios<strong>of</strong> recycled air to that discharged through the vent tothe outside air to maintain the wet-bulb temperatureat the scheduled values. For that reason, commercialkilns appear to operate under very steamy conditionsto the casual observer. The high degree <strong>of</strong> air recyclemeans that small deviations in evaporation are fedback to the air-inlet face <strong>of</strong> the stack, disturbing theconditions there. This disturbance is then propagatedthrough the stack. In theory, with high recycle ratiosand extensive dryers (z > 1), it is possible, once thewood has reached the critical moisture content at theair-inlet face, for internal drying rates to rise abovethose for very wet greenwood at the air-inlet face <strong>of</strong>the stack initially and even to exceed them (Keey,1968). Although the necessary combination <strong>of</strong> factorsto get substantial rate enhancements is unlikely in thekiln drying <strong>of</strong> most lumber species, the potential forincreases in drying rate (and strain development)should be borne in mind.<strong>36</strong>.3.4 PRACTICAL CONSIDERATIONSVarious works give a detailed overview <strong>of</strong> kilnpractice. Such overviews include those by Pratt andTurner (1986), Boone et al. (1988), Hilderbrand (1989),Mackay and Oliveira (1989), Simpson (1992), andHaslett (1998).ß 2006 by Taylor & Francis Group, LLC.


<strong>36</strong>.3.4.1 Schedule DevelopmentKiln schedules are <strong>of</strong> two kinds. A tropical forest containsmany species, <strong>of</strong>ten diffusely spread, so that individualidentification <strong>of</strong> species and separate drying israrely economic or feasible. Economic drying <strong>of</strong> mixedspecies involves the grouping <strong>of</strong> these according to somecriteria. Simpson and Baah (1989), for example, choosebasic density and initial moisture content, on thegrounds that high-density, moist lumber is the mostsusceptible to drying defects and requires the longestdrying times. Using this system, Hidayat and Simpson(1994) were about to define drying-time factors for 12species groups. On the other hand, temperate productionforests may be confined to a single kind <strong>of</strong> tree or alimited range <strong>of</strong> species, and comprehensive speciesspecificschedules can be developed for these. For example,Haslett (1998) specifies various schedules for asingle species, P. radiata, according to the grade andend use <strong>of</strong> the wood.Traditionally, schedules have been developed on acautious, trial-and-error basis from small-scale testson the woods <strong>of</strong> interest. Pandey (2001) notes that theseven standard, empirically derived schedules fornearly 150 Indian woods could be fitted to a diffusionalmodel. Brandaõ and Perré (1996) propose adrying test at 908C on small boards to provide informationon drying rates and deformation criteria. Thisleads to an alternative species-grouping approach tothat put forward by Hidayat and Simpson (1994).Increasingly, however, detailed simulation <strong>of</strong> themass-transfer and strain-development behavior isused as a basis for determining appropriate kilntemperaturesettings (e.g., Langrish et al., 1997; Aguiarand Perré, 2000a; Nijdam et al., 2000).Kiln drying involves both the transfer <strong>of</strong> moisturethrough the boards and the evaporation from theexposed surfaces into the air circulating through thekiln. These processes must be kept in balance tomaintain the fastest possible drying rate. With impermeabletimbers, the transfer processes within thewood are rate-limiting. The wood temperature is theimportant variable. With permeable timbers, particularlyunder high-temperature conditions, externaltransfer mechanisms become important and highkiln-air velocities (7 m s 1 and higher) are used toenhance the heat- and mass-transfer coefficients.To sustain the rate <strong>of</strong> drying as the wood driesout, either the dry-bulb temperature or the wet-bulbdepression may be raised, thus providing a greateroverall driving force for the drying process. However,faster drying rates can lead to steeper moisture contentgradients with the risk <strong>of</strong> excessive strain developmentand checking. Thus most schedules specifyrelatively small wet-bulb depressions, correspondingto relatively high EMC at kiln temperatures, producingmodest surface shrinkage. In a conventionalschedule the wet-bulb depression is typically about5 to 108C, whereas for high-temperature dryingwet-bulb depressions can be 508C or more.Whenever it is important to avoid significant colordevelopment in the wood on drying, low kiln temperatureshave to be used. Even though in New Zealandwhere accelerated conventional schedules (with dry/wet-bulb temperatures <strong>of</strong> 90/608C) are used to dryappearance-grade timbers, commercially, kiln temperaturesnot greater than 508C are employed to getvery pale wood (Keey, 2003).Although higher airflows seem justifiable at thestart <strong>of</strong> a schedule when the evaporation from theexposed surface is more controlling, there is somedoubt whether there is an economical advantage indesigning kilns with variable-speed fans (Riley andHaslett, 1996). However, one manufacturer claims significantfan-energy savings by reducing the rate <strong>of</strong>circulating air toward the end <strong>of</strong> the schedule in thedrying <strong>of</strong> a s<strong>of</strong>twood without causing any significantextension <strong>of</strong> drying time (Fogarty, Priv. Comm., 2002).<strong>36</strong>.3.4.2 Kiln ControlAs the basic control <strong>of</strong> the drying process dependsupon the wet- and dry-bulb readings, control <strong>of</strong> thesetemperatures has been the normal method <strong>of</strong> maintainingthe drying schedule. Most single-zone, steamheatedkilns mount split bulbs under the overheadheating coils, one in each plenum some distance <strong>of</strong>fthe floor (Culpepper, 1990). In kilns divided into twozones along their length, the bulbs are placed midwayin each zone. The thermometer bulbs must be protectedfrom radiation from hot surfaces and have anadequate air circulation over them. Careful location<strong>of</strong> the bulbs is important, as significant air-temperaturegradients can exist in the plenum spaces.Simple lumber-drying installations having one ortwo kilns might have simple electronic-control systemswith discrete programmable controls, timers,and chart recorders for temperature and steamingcontrol. A more sophisticated system for a mediumsizedfaculty might have a computer-based systemrunning on windows-type s<strong>of</strong>tware to give a visualreadout <strong>of</strong> kiln conditions during the schedule (Keeyet al., 2000). Endpoint determination can be based onthe temperature drop across the load (TDAL), whichis a function <strong>of</strong> the extent <strong>of</strong> drying (Taylor andLandoch, 1990; Martin et al., 1995).As pointed out by Morén (2001), TDAL reflectsthe average drying rate in the kiln directly. He suggeststhat monitoring <strong>of</strong> the TDAL provides amethod <strong>of</strong> ‘‘adaptive kiln control,’’ with the TDALß 2006 by Taylor & Francis Group, LLC.


set at a constant value at constant wet-bulb temperatureafter an initialization period, followed by aperiod at constant dry-bulb temperature.The measurement <strong>of</strong> the boards’ moisture content,at the end <strong>of</strong> drying, has frequently been done withhand-held resistivity meters, <strong>of</strong>ten complementedwith more time-consuming gravimetric measurementswith sample boards (EN 13183-1). The Standard EN13183-2, Round and Sawn Lumber—Procedure toMeasure the Moisture Content, employs a hammerprobe with insulated measuring pins, which areinserted to a depth <strong>of</strong> approximately one third <strong>of</strong>the board’s thickness, and is valid over the moisturecontent range from about 7 to 30%. This techniquehas been adapted by inserting probes in the kilnstacks to give a continuous reading <strong>of</strong> the moisturecontent in the later stages <strong>of</strong> drying. Nonpenetrativesystems are available in which the resistivity probesare aluminum fillets placed along the length <strong>of</strong> thelumber charge to measure the resistance <strong>of</strong> all theboards that the probes are in contact with. All resistivemethods, however, are limited to measuring moisturecontents from just above fiber saturation andbelow, with one system having a claimed repeatability<strong>of</strong> +2% when the average moisture content is lessthan 16% (Furniss, Priv. Comm., 2002).The alternative use <strong>of</strong> microwave-sensing technologiesis attractive in enabling a wider range <strong>of</strong>moisture contents to be determined from green todry (Holmes and Riley, 1996; Riley and Holmes,2001). The first prototype tested used a waveguide,40 20 1.2 mm, with angled slots in the long face,connected by a high-temperature coaxial cable to theoscillator and placed in a stack so that the slots werein contact with the lower face <strong>of</strong> a board. A laterdevelopment enabled the possibility <strong>of</strong> the stack’smoisture distribution to be determined. The experimentshave demonstrated the potential for the manufacture<strong>of</strong> a rugged industrial system. However, unlessthe timber is very uniform in properties, this techniqueis uncertain as the relationship between themicrowave signal and the moisture content becomesfuzzy because <strong>of</strong> factors such as variations in wooddensity among the boards.An online contact-free method <strong>of</strong> measuring themoisture content <strong>of</strong> wood, which involves traversingthe boards on a band between sensor heads, could bedeveloped based on techniques used in papermaking.The method would be suitable for sorting boardsafter drying. At least one such method is commerciallyavailable, but the measurement principle hasnot been divulged (Smith, Priv. Comm., 2002).The benefits <strong>of</strong> even a relatively simple kilncontrolsystem are illustrated in an example quotedby Culpepper (1990). The incorporation <strong>of</strong> aprogrammable logic controller at one site increasedthe overall grade recovery from 70.7 to 81.9%, with areduction in energy costs <strong>of</strong> 43% for steam and 10%for electrical power. The total saving represented a1.2-y payback on the investment.<strong>36</strong>.3.4.3 Volatile EmissionsAcidic corrosion is a well-recognized feature <strong>of</strong> kilndrying, particularly with certain species such as oaks.Packman (1960) notes that, <strong>of</strong> some 150 species studies<strong>of</strong> both hardwoods and s<strong>of</strong>twoods, the majority(80%) has a pH between 4 and 6 and only one had apH consistently above 7. In an extreme case, with ahardwood <strong>of</strong> high extractives content, sufficiently severerusting <strong>of</strong> steel trolley wheels and other ferrousfittings in a pilot-plant kiln required their replacementafter a 2-week period. Extensive use <strong>of</strong> aluminumlinings with plastic piping in modern kilns has largelyavoided the problems <strong>of</strong> corrosion.The corrosive nature <strong>of</strong> the volatile substances releasedduring kiln drying is derived from the presence <strong>of</strong>free acetic acid in the wood and from the hydrolysis <strong>of</strong>the various acetyl groups attached to the hemicelluloses.S<strong>of</strong>twoods generally yield greater emissions thanhardwoods, because the latter are normally dried atlower temperatures and have lower resin content.Some volatile substances, such as formaldehyde, comefrom the thermal degradation <strong>of</strong> the hemicelluloses andlignin, whereas green pines release terpenes and theirderivatives. Table <strong>36</strong>.11 lists the concentrations <strong>of</strong> volatileemissions from two high-temperature schedules.Milota (2001) compared the emissions from variousNorth American lumber species in a small-scale kilnTABLE <strong>36</strong>.11Concentrations <strong>of</strong> Volatile Emissions Arisingfrom Two High-Temperature Kiln Schedulesfor Pinus radiataCompoundConcentration, g m 3 at dry/wet-bulb temperatures (8C)120/70 140/90Formaldehyde 19.5 31.0Acetic acid 21.7 38.2Monoterpenes 34.8 66.4Hydroxylated monoterpenes 12.6 10.7Condenser residues(resins and fatty acids)16.4 16.4Source: McDonald, K.A. and Wastney, S., Analysis <strong>of</strong> volatileemission from kiln drying <strong>of</strong> radiata pine, Proceedings <strong>of</strong> theEighth International Symposium on <strong>Wood</strong> Pulp Chemistry, Vol. 3,431–4<strong>36</strong>, 1995.ß 2006 by Taylor & Francis Group, LLC.


with those from the same wood in a commercial kiln.The laboratory kiln functioned like the mill kiln withrespect to venting characteristics, temperature, andhumidity, but it dried the wood faster, possibly becausethe load is narrower. In general, the small-scale methodwas fairly repeatable for the determination <strong>of</strong> volatileorganic chemicals, with a standard deviation <strong>of</strong> 8 to16% <strong>of</strong> the mean value, but the repeatability was poorerfor methanol and formaldehyde.<strong>36</strong>.3.4.4 Equalization and Stress ReliefAt the end <strong>of</strong> kiln drying there is always some board-toboardvariations in moisture content, arising from theinherent differences in drying characteristics <strong>of</strong> individualboards and the progressive humidification <strong>of</strong> thecirculating air. Because kiln drying is basically a processin which the boards are forced to reach a new equilibrium,moisture content pr<strong>of</strong>iles also exist within eachpiece <strong>of</strong> wood. With moisture-based schedules, the kilnconditions are reset at the end <strong>of</strong> the schedule to give anEMC that is 2 to 3% below the desired value. Thisstrategy prevents overdrying <strong>of</strong> the drier boards whileallowing the wetter boards to dry further toward thetarget end moisture content. Some authorities (e.g.,Haslett, 1998) recommend that the s<strong>of</strong>twood lumbershould be slightly overdried to ensure that the moisturecontent variation both within and between boards becomessmall. Any subsequent steaming, which is undertakenfor stress relief, will raise the moisture content <strong>of</strong>the overdried load toward the specified value.Kiln schedules are designed to dry the lumber asfast as possible without causing unacceptable defectsto appear due to excessive strain development. Residualstresses in the wood must be relieved if thelumber is to be further processed. Steaming is a commonmethod <strong>of</strong> doing this.In smaller installations, steaming may be done inthe kiln; but, in larger installations, a separate chambersupplied with low-pressure, saturated steam isused for this purpose. Such chambers are not normallysupplied with fans, and stratification <strong>of</strong> thesteam and air can be a problem. Lumber that ishigh-temperature dried should be cooled so that thewood temperature falls to about 70 to 958C to enablethe wood to pick up moisture. The steaming induces areversal <strong>of</strong> the moisture content pr<strong>of</strong>ile through thewood, with concomitant reduction in the residualstresses.Chen et al. (1997b) examined various stress-reliefstrategies for sapwood boards <strong>of</strong> s<strong>of</strong>twoods and havefound that other procedures are suitable in additionto final cooling and steaming. These include simplecooling under cover or the use <strong>of</strong> a schedule consisting<strong>of</strong> intermittent drying and conditioning cycles.<strong>36</strong>.3.5 LESS-COMMON DRYING METHODSThe majority <strong>of</strong> lumber-drying installations are convectivedrying chambers worked at atmospheric conditionsand temperatures not greatly above ambient. Forstructural-grade, permeable s<strong>of</strong>twoods, however, whencolor development is not a prime concern, elevatedtemperatures can be used to get very fast drying processes.Designs under consideration include kilns beingworked to temperatures <strong>of</strong> 2008C, with air velocitiesbetween the boards up to 15 m s 1 . Continuouslyworked kilns then become attractive with very fastdrying. Such designs will require particular attentionto thermal expansion, reliable heat-exchange equipment,and venting <strong>of</strong> moisture vapor.Many <strong>of</strong> the advantages <strong>of</strong> high-temperature dryingcan be obtained by working under vacuum. Inparticular, such a process gives rise to an internaloverpressure and an additional and efficient drivingforce for internal moisture migration. The lower operatingtemperatures are an advantage in drying heatsensitivewoods and in minimizing color developmentwhen pale products are required.Similar advantages <strong>of</strong> lower working temperaturesare obtained in the use <strong>of</strong> dehumidifying heatpumpsystems, with the added bonus <strong>of</strong> lower thermalenergyuse to compensate for extended drying times.The drying principle <strong>of</strong> these kilns, however, is thesame as vented conventional kilns.Most kilns, either direct-fired or steam-heated, usewastewood as the primary fuel. Other heating arrangementshave been advocated such as the use <strong>of</strong> microwave,radio frequency, and solar energy, with the latterbeing attractive in remote locations for small kilns.<strong>36</strong>.3.5.1 Vacuum <strong>Drying</strong>Vacuum dryers have been commercially available formany years and their use is regarded as a standardpractice in Europe for the drying <strong>of</strong> high-qualityhardwoods economically, which would otherwise bedifficult to dry (Hilderbrand, 1989). Descriptions <strong>of</strong>vacuum drying are given by Ressel (1994), Audebertand Temmar (1997), and Jomaa and Baixeras (1997).Because <strong>of</strong> the enhanced internal moisture migrationunder vacuum, the rate <strong>of</strong> drying can be as rapid asthat at a much higher temperature at atmospheric pressure.However, the higher specific volume <strong>of</strong> vapor associatedwith the reduced pressure is a severe limitation forheat transport by convection (Perré et al., 1995).Several industrial solutions have been proposed:. The use <strong>of</strong> plates heated by electrical resistanceor circulation <strong>of</strong> heated water are placed betweeneach layer <strong>of</strong> boards; heat is supplied toß 2006 by Taylor & Francis Group, LLC.


oards by conduction whereas the vacuum levelis used to enhance the internal mass transfer.. Discontinuously operating kilns, with twoperiods <strong>of</strong> about 1 h alternate: a heating periodat atmospheric pressure and a drying period atreduced pressure.. Finally, the most recent ‘‘high-vac’’ kilns use aslightly higher pressure level (more than 100mbar), together a very high linear air velocity(10 m s 1 or more), to compensate for the loss <strong>of</strong>thermal capacity <strong>of</strong> the air; this method hasproved to be very effective.In the latter method, uniformity <strong>of</strong> the air distributionthrough the load is important to ensure evenness <strong>of</strong>drying, with regions <strong>of</strong> low velocities resulting inhigher final moisture contents (Ledig and Militzer,1999). The positions <strong>of</strong> fans and heating coils havean important bearing on the temperature and on thefinal moisture content <strong>of</strong> the load (Hedlund, 1996).Vacuum dryers with overhead fans provide a fairlycontrolled airflow path through the load, but otherfan locations can result in ill-defined pressure andsuction sides. An overhead-fan dryer, however, wasfound to yield a systematic variation in temperaturebetween the door and the other end <strong>of</strong> the dryer,which might have been reduced by dividing the unitinto separate temperature-control zones. Techniquesto overcome the inherent poor heat transfer invacuum dryers include the use <strong>of</strong> heated plates betweenthe boards or intermittent heating with superheatedsteam. Another suggested technique employs aheating cycle at atmospheric pressure, when the heattransfer is better, followed by a vacuum-drying cycle(Guilmain et al., 1996). Tests on drying oakwood atpilot and industrial scales showed that the discontinuousprocess was faster, with less susceptibility tomechanical damage <strong>of</strong> the wood, but the thermalconsumption was higher than under continuousvacuum conditions.Behnke and Militzer (1996) have produced avacuum-dryer model for design and process-controlpurposes based on a characteristic drying curve for thewood’sdryingbehavior.Hilderbrand(1989)claimsthatcommercialdryingtimeinvacuumkilnsvariesbetweenone half and one third <strong>of</strong> those found in conventionalconvective kilns under atmospheric pressure.<strong>36</strong>.3.5 .2 Dehum idifi er Kiln s<strong>Drying</strong> at low temperatures, which is a feature <strong>of</strong>seasoning refractory timbers, is energy-inefficient.A dehumidifier kiln reduces the thermal-energy consumptionby incorporating an air-conditioning unitthat recovers heat by cooling the kiln air below itsdew point and, in effect, recycling the latent heat <strong>of</strong>condensation. As moisture is removed as condensedliquid rather than vapor in warm discharged air, theassociated thermal loss is avoided. However, a smallamount <strong>of</strong> venting is needed for humidity-controlpurposes. Volatile organic chemicals normally removedwith the vented moist air now appear in thecondensate stream, which potentially could be sent toa separate unit for chemical recovery.Figure <strong>36</strong>.48 shows the layout <strong>of</strong> a heat-pumpdehumidifying kiln. Moist air is drawn over the evaporatorand condenser consecutively in a Rankine cycleheat pump. Besides these basic elements such as anevaporator, a condenser with its associated compressor,and expansion valve, there is an accumulator thatprevents the refrigerant from entering and damagingthe compressor and a subcooling heat exchanger toenhance the effectiveness <strong>of</strong> the heat pump.The performance <strong>of</strong> dehumidifier kilns is normallyexpressed in terms <strong>of</strong> the specific moisture-extractionrate (SMER), which is the amount <strong>of</strong> moistureextracted per unit energy input. Two such ratiosmay be defined: one representing how effectively thedehumidifier extracts moisture from the air as condensateand the other (the kiln SMER) representinghow efficiently the kiln removes moisture from thelumber including the condensate and venting. Thekiln SMER for convective kilns is limited to about0.8 to 0.9 kg kWh 1 , compared with values in therange <strong>of</strong> 1.5 to 2.5 kg kWh 1 for commercial dehumidifierkilns (Davis, 2001). Some <strong>of</strong> the lower valuesmay reflect the poor insulation <strong>of</strong> the tested kilnsrather than a defect in the process.Dehumidifying kilns are limited in operating temperatureby the working limits <strong>of</strong> the compressor(


AirCirculation fanWeightsCondenserCompressorAccumulatorStickersTimberLiquidwaterEvaporatorExpansionvalveSubcoolerFIGURE <strong>36</strong>.48 A typical configuration for a heat-pump dehumidifying kiln. (Adapted from Davis, C.P., <strong>Drying</strong> Pinusradiata boards in dehumidifier conditions, Ph.D. thesis, Otago University, New Zealand, 2001.)thumb, these high-frequency heating methods becomeeconomically attractive for new kilns if the dryingrate is increased fourfold over that for conventionaldrying. In general, the use <strong>of</strong> dielectric and microwaveheating may become attractive for the small-scaledrying <strong>of</strong> high-value hardwood species that are difficultto dry by conventional means. For example, Smithand Smith (1994) report the use <strong>of</strong> radio-frequencyheating for the drying <strong>of</strong> oakwood in a small vacuumkiln <strong>of</strong> 23-m 3 capacity, which had a lower capital costbut higher energy costs than a conventional dryerfor the same duty. For very small power requirements,microwave heating is more attractive; whenthe power requirement exceeds 50 kW, however,economics favor the higher-power tubes in theradio-frequency range. In one Canadian system, radi<strong>of</strong>requencydrying is used to finish the seasoning <strong>of</strong>conventionally dried lumber that has not met targetmoisture content.Heating is generated in the dipolar rotation <strong>of</strong>water molecules as they try to orient themselves inthe rapidly changing polarity <strong>of</strong> the applied electricalfield. The power developed per unit volume is given byP ¼ kE 2 f « 0 tan d (<strong>36</strong>:24)where k is the dielectric constant, E is the electricfield strength, f is the field’s frequency, «’ is the relativepermeability, and tan d is the loss tangent ordissipation factor. The field’s strength and its frequencyare fixed by the equipment, whereas theother parameters are material-dependent. As the dielectricconstant <strong>of</strong> water is over an order <strong>of</strong> magnitudegreater than the woody materials, moisture ispreferentially heated, a process that leads to a moreuniformly moist product with time. This feature isone <strong>of</strong> the attractions <strong>of</strong> the technique, for example,in moisture leveling in the manufacture <strong>of</strong> plywood toavoid delamination during subsequent hot pressing(Schiffmann, 1995).There is also a contribution due to ionic conductionbecause <strong>of</strong> the presence <strong>of</strong> ions in the sap. Thismode <strong>of</strong> heating is not significantly dependent oneither the temperature or the frequency <strong>of</strong> the appliedfield, but is directly dependent on the charge densityand mobility <strong>of</strong> the ions.Because the heating is internally generated, ratherthan convectively warmed at the exposed surface <strong>of</strong> theboards, high and damaging internal pressures can becreated in the process. For example, internal overpressures<strong>of</strong> 60 kPa have been reported by Antti (1992) forpower inputs <strong>of</strong> 1.25 kW on drying 100 50 1660-mm boards. Under vacuum drying, such overpressuresbecome less damaging. Thus, high-frequency heatinghas been advocated for use with vacuum drying because<strong>of</strong> the difficulty in achieving adequate convective heatingunder vacuum, and a summary <strong>of</strong> its historic developmentis given by Resch and Gautsch (2001). Thisß 2006 by Taylor & Francis Group, LLC.


KilnSolarcollectorCondenserControlvalveAirflowFanRefrigerantdischarge line<strong>Wood</strong>stackCompressorBlowerMotorDamperEvaporatorRefrigerantsuction lineFIGURE <strong>36</strong>.49 A solar-dehumidifier dryer. (Adapted from Chen, P.Y.S., Helmer, W.A., and Rosen, H.N., Experimentalsolar-dehumidifier kiln for drying lumber, Forest Prod. J., 32(9): 35–41, 1982.)technique is attractive for beech and oak timbers in theEuropean market because <strong>of</strong> the retention <strong>of</strong> their naturallight color with low-temperature drying.Perré and Turner (1997, 1999a) have described anumerical model <strong>of</strong> microwave drying <strong>of</strong> s<strong>of</strong>twoodwith an oversized waveguide. In this work, internaloverpressure reaching two to three times the atmosphericpressure has been reported both in experimentaland numerical results.<strong>36</strong>.3.5.4 Solar <strong>Drying</strong>Solar drying <strong>of</strong> lumber has attractions in remotelocations with favorable climates because <strong>of</strong> the‘‘free’’ nature <strong>of</strong> the energy source. Imré (1995) hasclassified solar-heated dryers into three main groups:1. Solar natural dryers that use only the sun2. Semiartificial solar dryers with a fan to supply acontinuous flow <strong>of</strong> air through the load3. Solar-assisted artificial dryers, which may use anauxiliary energy source for boosting the heatingrateMixed types include a solar-dehumidifier dryer withforced-air recirculation, as shown in Figure <strong>36</strong>.49.Many solar dryers described in the literature aresimple greenhouse kilns (e.g., Langrish and Keey,1992). In these units, the solar collector is fitted withinthe structure that holds the load and the airflow ismaintained by fans. The solar energy, in other cases,is collected externally in heat-storage systems orpanels, as illustrated in Figure <strong>36</strong>.49. Greenhousekilns have attractions in simplicity <strong>of</strong> constructionand operation.The daily world-average solar radiation on a horizontalsurface is 3.82 kWh m 2 (McDaniels, 1984),with values in tropical countries being higher (up to7.15 kWh m 2 ) (Imré, 1995). However, Plumptre(1989), reported by Keey et al. (2000) on reviewing35 solar kiln designs, notes that the location <strong>of</strong> thesewas spread almost uniformly over the range in latitudefrom 0 to 508.Langrish and Keey (1992) observed one operationalfeature <strong>of</strong> the use <strong>of</strong> a greenhouse kiln. Withthe kiln’s vents shut overnight and with the drop inambient temperature, the relative humidity in the kilnwould rise sufficiently for moisture to condense on thewood’s surface. This provided a degree <strong>of</strong> conditioning,which prevented the development <strong>of</strong> excessivechecking in a refractory hardwood being dried.REFERENCESAguiar, O. and Perré, P., 2000a. The ‘‘flying wood’’ testused to study the variability <strong>of</strong> drying behaviour <strong>of</strong>oak, in Quality <strong>Drying</strong> <strong>of</strong> Hardwood. 2nd Workshop<strong>of</strong> COST Action E15, Sopron, Hungary, 10 pp.Aguiar, O. and Perré, P., 2000b. Pocesso de secagem aceleradade madeira baseado nas suas propriedadesreológicas (Accelerated drying process for woodbased on its rheological properties)—Industrialß 2006 by Taylor & Francis Group, LLC.


patten N 1265, Instituto nacional de proteçãoindustrial, Brazil.Aléon, D., Chanrion, P., Négrié, G., and Perré, P., 2003.FormaXylos 4—Le séchage (Training in <strong>Wood</strong> Science:<strong>Drying</strong>, Vol. 4), CD-Rom français/English,CTBA, Paris.Antti, A.L., 1992. Microwave drying <strong>of</strong> hardwood: simultaneousmeasurements <strong>of</strong> pressure temperature andweight reduction, Forest Prod. J., 42(6): 49–54.Ashworth, J.C., 1977. The mathematical simulation <strong>of</strong>batch-drying <strong>of</strong> s<strong>of</strong>twood timber, Ph.D. thesis, University<strong>of</strong> Canterbury, New Zealand.Ashworth, J.C. and Keey, R.B., 1979. The kiln seasoning <strong>of</strong>s<strong>of</strong>twood boards, Chem. Eng., 347(8): 593–598, 607.Audebert, P. and Temmar, A., 1997. Vacuum drying <strong>of</strong>oakwood: moisture strains and drying process, <strong>Drying</strong>Technol., 15: 2281–2302.Banks, W.B., 1968. A technique for measuring the lateralpermeability <strong>of</strong> wood, J. Inst. <strong>Wood</strong> Sci., 4(2): 35–41.Behnke, C. and Militzer, K.-E., 1996. Vacuum dryer model,Proceedings <strong>of</strong> the Fifth International <strong>Wood</strong> <strong>Drying</strong>Conference, Québec, Canada, pp. 139–145.Bolton, A.J. and Petty, J.A., 1978. The relationship betweenthe axial permeability <strong>of</strong> wood to dry air and to a nonpolar solvent, <strong>Wood</strong> Sci. Technol., 12: 111–126.Booker, R.E., 1994. Collapse or internal checking, whichcomes first? Proceedings <strong>of</strong> the 4th IUFRO <strong>Wood</strong><strong>Drying</strong> Conference, 133–140, Rotorua, New Zealand.Booker, R.E., 1996. New theories for liquid water flow inwood, Proceedings <strong>of</strong> the 5th IUFRO <strong>Wood</strong> <strong>Drying</strong>Conference, 437–445, Québec, Canada.Boone, R.S., Kozlik, C.J., Bois, P.J., and Wegert, E.M.,1988. Dry kiln schedules for commercialwoods temperate and tropical, USDA FS ForestProducts Laboratory, General Technical ReportFRL-GTR-57.Bosshard, H.H., 1984. Holzkunde, Band 2, Zur Biologie,Physik und Chemie des Holzes (<strong>Wood</strong> Science: Biology,Physics and Chemistry <strong>of</strong> <strong>Wood</strong>, Vol. 2), BirkhäuserVerlag, Basel.Brandaõ, A. and Perré, P., 1996. The ‘‘flying wood’’: aquick test to characterise the drying behaviour <strong>of</strong>tropical woods, Proceedings <strong>of</strong> the Fifth InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec, Canada,pp. 315–324.Carlquist, S., 2001. Comparative <strong>Wood</strong> Anatomy, Springer,Berlin.Chen, G., Keey, R.B., and Walker, J.C.F., 1997a. Thedrying stress and check development on hightemperaturekiln seasoning <strong>of</strong> sapwood Pinusradiata boards, Holz Roh-Werks., 55: 59–64.Chen, G., Keey, R.B., and Walker, J.C.F., 1997b. Stressrelief for sapwood Pinus radiata boards by coolingand steam-conditioning processes, Holz als Roh-Werks., 55: 351–<strong>36</strong>0.Chen, P.Y.S., Helmer, W.A., and Rosen, H.N., 1982. Experimentalsolar-dehumidifier kiln for drying lumber,Forest Prod. J., 32(9): 35–41.Chen, P.Y.S., Zhang G., van Sambeek J.W., 1998. Relationshipsamong growth rate, vessel lumen area,and wood permeability for three central hardwoodspecies, Forest Products Journal, 48:87–90.Choong, E.T. and Kimbler, O.K., 1971. A technique <strong>of</strong>measuring water flow in woods <strong>of</strong> low permeability,<strong>Wood</strong> Sci., 4(1): 32–<strong>36</strong>.Choong, E.T. and Tesoro, F.O., 1989. Relationship <strong>of</strong> capillarypressure and water saturation in wood,<strong>Wood</strong> Sci. Technol., 23: 139–150.Choong, E.T., Tesoro, F.O., and Manwiller, F.G., 1974.Permeability <strong>of</strong> twenty-two small diameter hardwoodsgrowing on southern pine sites, <strong>Wood</strong>Fiber, 6(1): 91–101.Cloutier, A. and Fortin, Y., 1991. Moisture content–water potential relationship <strong>of</strong> wood from saturatedto dry conditions, <strong>Wood</strong> Sci. Technol., 25:263–280.Clair, B., 2001. Etude des propriétésmécaniques et du retraitau séchage du bois à l’échelle de la paroi cellulaire:essai de compréhension du comportement macroscopiqueparadoxal du bois de tension à couche gélatineuse,PhD dissertation, ENGREF, Montpellier.Comstock, G.L., 1967. Longitudinal permeability <strong>of</strong> woodto gases and nonswelling liquids, Forest Prod. J.,17(10): 41–46.Comstock, G.L., 1970. Directional permeability <strong>of</strong> s<strong>of</strong>twoods,<strong>Wood</strong> Fiber, 1: 283–289.Comstock, G.L. and Côté, W.E., 1968. Factors affectingpermeability and pit aspiration in coniferous sapwood,<strong>Wood</strong> Sci. Technol., 2: 279–291.Cown, D.J., 1992. New Zealand radiata pine and Douglasfirsuitability for processing, FRI Bull., 168,NZFRI, Rotorua.Cown, D.J. and McConchie, D.L., 1983. Studies on theintrinsic properties <strong>of</strong> new-crop radiata pine II:wood characteristics <strong>of</strong> 10 trees from a 24-year-oldstand grown in the Central North Island, FRI Bull.,37, NZFRI, Rotorua.Culpepper, L., 1990. High temperature drying-enhancingkiln operations, Miller Freeman, San Francisco, CA.Dahlblom O., Petersson H. and Ormarsson S., 1994.Numerical simulation <strong>of</strong> the development <strong>of</strong> deformationand stress in wood during drying, proceedings<strong>of</strong> the 4th IUFRO <strong>Wood</strong> <strong>Drying</strong>Symposium, 165–180, Rotorua, New Zealand.Dahlblom O., Ormarsson S. and Petersson H., 1996. Simulation<strong>of</strong> wood deformation processes in drying andother environmental loading, Annals <strong>of</strong> Forest Science,53: 857–866.Davis, C.P., 2001. <strong>Drying</strong> Pinus radiata boards in dehumidifierconditions, Ph.D. thesis, Otago University,New Zealand.Davis, C.P., Carrington, C.G., and Sun, Z.F., 2001. <strong>Drying</strong>rate curves for the dehumidifier drying <strong>of</strong> Pinusradiata boards, Proceedings <strong>of</strong> the Seventh InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Tsukuba,Japan, pp. 216–221.Davis, C.P., Carrington, C.G., and Sun, Z.F., 2002. Theinfluence <strong>of</strong> compression wood on the drying curves<strong>of</strong> Pinus radiata in dehumidifier conditions, <strong>Drying</strong>Technol., 20: 2005–2026.ß 2006 by Taylor & Francis Group, LLC.


Doe, P.E., Oliver, A.R., and Booker, J.D., 1996a. A nonlinealstrain and moisture content model <strong>of</strong> variablehardwood drying schedules, Proceedings <strong>of</strong> theFourth IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Rotorua,New Zealand, pp. 203–210.Doe, P.D., Booker, J.D., Innes, T.C., and Oliver, A.R.,1996b. Optimal lumber seasoning using acousticemission sensing and real-time strain modelling,Proceedings <strong>of</strong> the Fifth International IUFRO <strong>Wood</strong><strong>Drying</strong> Conference, Québec, Canada, pp. 209–212.Dullien, F.A.L., 1992. Porous Media: Fluid Transport andPore Structure, 2nd ed., Academic Press, London.Fengel D. and Wegener, G., 1984. <strong>Wood</strong>: Chemistry, Ultrastructure,Reactions, de Gruyter, Berlin.Forest Products Laboratory, 1999, <strong>Wood</strong> handbook—<strong>Wood</strong> as an engineering material. General TechnicalReport FPL-GTR-113, U.S. Department <strong>of</strong>Agriculture, Forest Service, Forest Products Laboratory,Madison, WI, 463 pp.Fumoto, H., Perng, W.R., Sebastian, L.P., 1984. Studies onFlow in <strong>Wood</strong> VIII. Air permeability and polymerpenetration <strong>of</strong> jack and red pines, MokuzaiGakkaishi, 30:646–652.Genevaux, J.-M., 1989. Le fluage à température linéairementcroissante: Caractérisation des sources de viscoélasticitéanisotrope du bois (Creep tests at increasingtemperature: determination <strong>of</strong> the sources <strong>of</strong> viscoelasticanisotropy in wood), Thèse de Doctorat del’Institut Nationnal Polytechnique de Lorraine,Nancy.Ge<strong>of</strong>frey, M.I., 1984. The glass transition <strong>of</strong> lignin andhemicellulose and their measurement by differentialthermal analysis, <strong>Wood</strong> Chem., 67(5): 118–121.Goring, D.A.I., 1963. Thermal s<strong>of</strong>tening <strong>of</strong> lignin, hemicelluloseand cellulose, Pulp and Paper Mag. Canada,T517–T527.Grace, C., 1996. <strong>Drying</strong> characteristics <strong>of</strong> heartwood, M.E.(Chen) thesis, University <strong>of</strong> Canterbury, NewZealand.Guilmain, C., Baixeras, O., and Jomaa,W., 1996. Discontinuousand convective vacuum drying <strong>of</strong> oak: experimentaland modelling studies, Proceedings <strong>of</strong>the Fifth International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference,Québec, Canada, pp. 147–157.Hanhijärvi, A., 1999. Deformation properties <strong>of</strong> Finnishspruce and pine wood in tangential and radialdirections in association to high temperaturedrying, part II: experimental results underconstant conditions, Holz als Roh- Werks., 57:<strong>36</strong>5–372.Haslett, A.N., 1998. <strong>Drying</strong> radiata pine in New Zealand:research and commercial aspects, FRI Bull., 206,NZFRI, Rotorua, New Zealand.Hedlund, B.A., 1996. Temperature and final moisture contentdistributions in vacuum driers with differentfan configurations, Proceedings <strong>of</strong> the Fifth InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec,Canada, pp. 159–167.Hidayat, S. and Simpson, W.T., 1994. Use <strong>of</strong> green moisturecontent and basic specific gravity to group tropicalwoods for kiln drying, USDA FS Forest ProductsLaboratory Research Note RN-0263, Madison,WI.Hilderbrand, R., 1989. Die Schnittholztrocknung (The <strong>Drying</strong><strong>of</strong> Sawn Timber). Hilderbrand Maschinenbau,Nürtingen.Hoadley, R.B., 1980. Understanding wood, a craftsman’sguide to wood technology. The Taunton Press,Newtown, Connecticut, 256pp.Holmes, W.S. and Riley, S.G., 1996. Microwave method forin-kiln moisture content measurement, Proceedings<strong>of</strong> the Fifth International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference,Québec, Canada, pp. 255–259.Imré, L., 1995. Solar drying, in Mujumdar, A.S. (Eds.), <strong>Drying</strong>’84, Hemisphere, Washington, D.C., pp. 373–452.Irvine, G.M., 1984. The glass transitions <strong>of</strong> lignin andhemicellulose and their measurements by differentialthermal analysis, Tappi J., 67(5): 118–121.Joly, P. and More-Chevalier, F., 1980. Théorie, pratique etéconomie du séchage des bois (Theory, Practice andEconomics <strong>of</strong> <strong>Wood</strong> <strong>Drying</strong>), H.Vidal éditeur, Paris.Jomaa, W. and Baixeras, O., 1997. Discontinuous vacuumdrying <strong>of</strong> oakwood: modelling and experimentalinvestigations, <strong>Drying</strong> Technol., 15: 2129–2144.Josza, L.A. and Middleton, G.R., 1994. A discussion <strong>of</strong>wood quality attributes and their practical implications,Forintek Canada Special Publ. SP-34.Kamke, F.A., Casey, L.J., 1988. Gas pressure and temperaturein the mat during flakeboard manufacture,Forest Products Journal, 38:41–43.Keey, R.B., 1968. Batch drying with air recirculation,Chem. Eng. Sci., 23: 1299–1308.Keey, R.B., 1978. Introduction to Industrial <strong>Drying</strong> Operations,Pergamon, Oxford.Keey, R.B., 2003. Growing pains: developing strategies todry fast-grown s<strong>of</strong>twood, Proceedings <strong>of</strong> the SecondNordic <strong>Drying</strong> Conference, Copenhagen, Denmark,25–27 June.Keey, R.B. and Pang, S., 1994. The high-temperature drying<strong>of</strong> s<strong>of</strong>twood boards in a kiln-wide model, Trans.IChemE., 72A: 741–753.Keey, R.B., Langrish, T.A.L., and Walker, J.C.F., 2000.The Kiln-<strong>Drying</strong> <strong>of</strong> Lumber, Springer, Berlin.Kho, P.C.S., Keey, R.B., and Walker, J.C.S., 1990. Thevariation <strong>of</strong> local mass-transfer coefficients instreamwise direction over a series <strong>of</strong> in-line bluntslabs, Proceedings <strong>of</strong> the Chemeca Conference,Auckland, New Zealand, pp. 348–355.Kollmann, F.P. and Côté, W.A., 1968. Principles <strong>of</strong> <strong>Wood</strong>Science and Technology, Solid <strong>Wood</strong>, Vol. 1,Springer, Berlin.Kröll, K., 1978. Trockner und Trocknungsverfahren (Dryersand <strong>Drying</strong> Processes), 2nd ed., Springer, Berlin.Langrish, T.A.G., 1999. The significance <strong>of</strong> gaps betweenboards in determining the moisture content pr<strong>of</strong>ilesin the drying <strong>of</strong> hardwood timber, <strong>Drying</strong> Technol.,17: 1481–1494.Langrish, T.A.G. and Keey, R.B., 1992, A solar-heated kilnfor drying New Zealand hardwoods, Trans IPENZChem./Elec./Mech., 18: 9–14.ß 2006 by Taylor & Francis Group, LLC.


Langrish, T.A.G., Keey, R.B., Kho, P.C.S., and Walker,J.C.F., 1993. Time-dependent flows in arrays <strong>of</strong>timber boards, flow visualisation, mass-transfermeasurements and numerical simulation, Chem.Eng. Sci., 48: 2211–2223.Langrish, T.A.G., Booker, A.S., Davis, C.L., Muson, H.E.,and Barton, G.W., 1997. An improved drying schedulefor Australian ironbark timber: optimisationand experimental verification, <strong>Drying</strong> Technol.,15: 47–70.Ledig, S. F. and Militzer, K.-E., 1999. Measured gas velocityand moisture content distribution in a convectivevacuum kiln, Proceedings <strong>of</strong> the SixthInternational IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference,Stellenbosch, South Africa, pp. 31–<strong>36</strong>.Lide, D.R., 1995. Handbook <strong>of</strong> Chemistry and Physics, 76thed., CRC Press, New York.Lowery D.P., 1979. Vapor pressure generated in wood duringdrying, <strong>Wood</strong> Science, 5: 73–80.Mackay, J.F.G. and Oliveira, L.C., 1989. Kiln operator’shandbook for Western Canada, Forintek CanadaCorp Spec. Publ. SP-31, Vancouver.Mark, R., 1967. Cell wall mechanics <strong>of</strong> tracheids, YaleUniversity Press, New Haven and London.Martensson, A., 1992. Mechanical behaviour <strong>of</strong> wood exposedto humidity variations, Doctoral dissertation,Lund Institute <strong>of</strong> Technology.Martin, M., Perré, P., and Moser, M., 1995. La perte detempérature à travers la charge: intérêt pour le pilotaged’un séchoir à bois à haute température(Temperature loss across the load: interest to controla high-temperature wood dryer), Int. J. HeatMass Transfer, 38(6): 1075–1088.Mauget, B. and Perré, P., 1999. A large displacement formulationfor anisotropic constitutive laws, Eur.J. Mech. A/Solids, 18: 859–877.McDaniels, D.K., 1984. The Sun, Our Future EnergySource, 2nd ed., Wiley, New York.McDonald, K.A. and Wastney, S., 1995. Analysis <strong>of</strong> volatileemission from kiln drying <strong>of</strong> radiata pine, Proceedings<strong>of</strong> the Eighth International Symposium on<strong>Wood</strong> Pulp Chemistry, Vol. 3, pp. 431–4<strong>36</strong>.McLean, V., 2003. Kiln drying: challenges still to meet, NZForest. Industry, 34(2): 32–<strong>36</strong>.Meyer, R.W., 1971. Influence <strong>of</strong> pit aspiration on earlywoodpermeability <strong>of</strong> Douglas-Fir, <strong>Wood</strong> Fiber, 2:328–339.Milota, M.R., 2001. Emissions from small-scale kilns: testconsiderations and comparison to large-scalekilns, Proceedings <strong>of</strong> the Seventh InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Tsukuba, Japan,pp. 3<strong>36</strong>–341.Mohager, S. and Toratti, T., 1993. Long-term bendingcreep <strong>of</strong> wood in cyclic relative humidity, <strong>Wood</strong>Sci. Technol., 27: 49–59.Morén, T., 2001. Adaptive kiln control systems based onCT-scanning and industrial practice, Proceedings <strong>of</strong>the Seventh International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference,Tsukuba, Japan, pp. 48–53.Nijdam, J.J., 1998. Reducing moisture-content variations inkiln-dried timber, Ph.D. thesis, University <strong>of</strong> Canterbury,New Zealand.Nijdam, J.J. and Keey, R.B., 1996. Influence <strong>of</strong> local variations<strong>of</strong> air velocity and flow direction reversals onthe drying <strong>of</strong> stacked timber boards in a kiln, Trans.IChemE, 74A: 882–892.Nijdam, J.J. and Keey, R.B., 1999. Airflow behaviorin timber (lumber) kilns, <strong>Drying</strong> Technol., 17:1511–1522.Nijdam, J.J. and Keey, R.B., 2000. The influence <strong>of</strong> kilngeometry on flow maldistribution across timberstacks in kilns, <strong>Drying</strong> Technol., 18: 1865–1877.Nijdam, J.J. and Keey, R.B., 2002. New timber kiln designsfor promoting uniform airflows within the woodstack, Trans. IChemE, 80(A): 739–744.Nijdam, J.J., Langrish, T.A.G., and Keey, R.B., 2000. Ahigh-temperature drying model for s<strong>of</strong>twood timber,Chem. Eng. Sci., 55: 3585–3598.Ormarsson, S., 1999. Numerical analysis <strong>of</strong> moisturerelateddistortions in sawn timber, PhD thesis,Chalmers University <strong>of</strong> Technology, Sweden.Ostberg, G., Salmen, L., and Terlecki, J., 1990. S<strong>of</strong>teningtemperature <strong>of</strong> moist wood measured by differentialscanning calorimetry, Holzforschung, 44:223–225.Packman, D.F., 1960. The acidity <strong>of</strong> wood, Holzforschung,14(6): 178–183.Pandey, C.N., 2001. Systemization <strong>of</strong> kiln drying schedulesfor commercial Indian woods, Proceedings <strong>of</strong> theSeventh International IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference,Tsukuba, Japan, pp. 78–83.Pang, S. and Haslett, A.N., 2002. Effects <strong>of</strong> sawing pattern ondrying rate and residual drying stresses <strong>of</strong> Pinus radiatalumber, Maderas: Ciencia Tecnologia, 4(1): 40–49.Pang, S. Keey, R.B. and Walker, J.C.F., 1994. Modelling <strong>of</strong>the high-temperature drying <strong>of</strong> mixed sap and heartwoodboards, Proceedings <strong>of</strong> the 4th IUFRO <strong>Wood</strong><strong>Drying</strong> Conference, 430–439, Rotorua, New Zealand.Pang, S., Keey, R.B., and Langrish, T.A.G., 1995. Modellingthe temperature pr<strong>of</strong>iles within boards duringthe high-temperature drying <strong>of</strong> Pinus radiata timber:the influence <strong>of</strong> airflow reversals, Int. J. HeatMass Transfer, 38: 189–203.Panshin, A.J. and de Zeeuw, C., 1980. Textbook <strong>of</strong> <strong>Wood</strong>Technology, 4th ed., McGraw-Hill, New York.Passard, J. and Perré, P., 2001. Creep tests under watersaturatedconditions: do the anisotropy ratios <strong>of</strong>wood change with the temperature and time dependency?Proceedings <strong>of</strong> the Seventh InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Tokyo, Japan,pp. 230–237.Passard, J. and Perre, P., 2005. Viscaqelastic behaviour <strong>of</strong>greenwood across the grain. Part II A temperaturedependent constitutive model defined by increasemethod. Annals <strong>of</strong> Forest sec., 62: 823–830.Perré P., 1987. Measurements <strong>of</strong> s<strong>of</strong>twoods permeability toair: Importance upon the drying model, InternationalCommunicable Heat and Mass Transfer, 14: 519–529.ß 2006 by Taylor & Francis Group, LLC.


Perré P., 1992. Transferts couplés en milieux poreux nonsaturés.Possibilités et limitations de la formulationmacroscopique, Habilitation à Diriger des Recherches,INPL, Nancy.Perré, P., 1995. <strong>Drying</strong> with internal vaporization: introducingthe concept <strong>of</strong> identity drying card, <strong>Drying</strong>Technol. J., 13(5–7): 1077–1097.Perré, P., 1996. The numerical modelling <strong>of</strong> physical andmechanical phenomena involved in wood drying: anexcellent tool for assisting with the study <strong>of</strong> newprocesses, Tutorial, Proceedings <strong>of</strong> the Fifth InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec,Canada, pp. 9–38.Perré P., 1997. Image analysis, homogenization, numericalsimulation and experiment as complementary toolsto enlighten the relationship between wood anatomyand drying behavior, <strong>Drying</strong> Technology Journal,15: 2211–2238.Perré P., 1998. The use <strong>of</strong> homogenisation to simulate heatand mass transfer in wood: Towards a double porosityapproach, Keynote lecture, 11th International<strong>Drying</strong> Symposium, published in <strong>Drying</strong> ’98, 57–72,Thessaloniki, Grèce.Perré, P., 1999. How to get a relevant material model forwood drying simulation? First COST Action E15<strong>Wood</strong> <strong>Drying</strong> Workshop, Edinburgh, 27 pp.Perré, P., 2001. The drying <strong>of</strong> wood: the benefit <strong>of</strong> fundamentalresearch to shift from improvement to innovation,Proceedings <strong>of</strong> the Seventh InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Tokyo, Japan,pp. 2–13.Perré, P. and Degiovanni, A., 1990. Simulations par volumesfinis des transferts couplés en milieu poreuxanisotropes: séchage du bois à basse et à hautetemperature (Finite-volume simulation <strong>of</strong> thecoupled transfer in anisotropic porous media:wood drying at low and high temperature), Int.J. Heat Mass Transfer, 33(11): 2463–2478.Perré, P. and Karimi, A., 2002. Fluid migration in twospecies <strong>of</strong> beech (Fagus silvatica and Fagusorientalis): a percolation model able to accountfor macroscopic measurements and anatomicalobservations, Maderas: Ciencia Tecnologia, 4(1):50–68.Perré, P. and Martin, M., 1994. <strong>Drying</strong> at high temperature<strong>of</strong> heartwood and sapwood: theory, experiment andpractical consequence on kiln control, <strong>Drying</strong> Technol.,12(8): 1915–1941.Perré, P. and Passard, J., 1995. A control-volume procedurecompared with the finite-element method for calculatingstress and strain during wood drying, <strong>Drying</strong>Technol. J., Special Issue Mathematical Modellingand Numerical Techniques for the Solution <strong>of</strong> <strong>Drying</strong>Problems, 13(3): 635–660.Perré, P. and Passard, J., 2002. A computational model <strong>of</strong>wood drying able to improve innovative processes:the importance <strong>of</strong> the mechanical constitutive law,Proceedings <strong>of</strong> the 13th International <strong>Drying</strong> Symposium,<strong>Drying</strong> 2002, Beijing, China, 10 pp.Perré, P. and Turner, I., 1996. The use <strong>of</strong> macroscopicequations to simulate heat and mass transfer inporous media, in Turner, I., and Mujumdar, A.S.(Eds.), Mathematical Modeling and NumericalTechniques in <strong>Drying</strong> Technology, Marcel Dekker,New York.Perré, P. and Turner, I., 1997. Microwave drying <strong>of</strong> s<strong>of</strong>twoodin an oversized waveguide, AIChE J., 43(10):2579–2595.Perré, P. and Turner, I., 1999a. The use <strong>of</strong> numerical simulationas a cognative tool for studying the microwavedrying <strong>of</strong> s<strong>of</strong>twood in an over-sizedwaveguide, <strong>Wood</strong> Sci. Technol., 33: 443–464.Perré, P. and Turner, I., 1999b. TransPore: a generic heatand mass TRANSFER computational model forunderstanding and visualising the drying <strong>of</strong> porousmedia, <strong>Drying</strong> Technol. J., 17(7): 1273–1289.Perré, P. and Turner, I., 1999c. A 3D version <strong>of</strong> TransPore:a comprehensive heat and mass transfer computationalmodel for simulating the drying <strong>of</strong> POR-OUS media, Int. J. Heat Mass Transfer, 42(24):4501–4521.Perré, P. and Turner, I., 2002. A heterogeneous wood dryingcomputational model that accounts for materialproperty variation across growth rings, Chem. Eng.J., 86(1–2): 117–131.Perré, P., Moser, M., and Martin, M., 1993. Advances intransport phenomena during convective drying withsuperheated steam or moist air, Int. J. Heat MassTransfer, <strong>36</strong>(11): 2725–2746.Perré, P., Joyet, P., and Aléon, D., 1995. Vacuum drying:physical requirements and practical solutions, Proceedings<strong>of</strong> Vacuum <strong>Drying</strong> <strong>of</strong> <strong>Wood</strong> ’95, Zvolen,Slovak Republic, pp. 7–34.Pratt, G.H. and Turner, C.H.C., 1986. Timber <strong>Drying</strong> Manual,2nd ed., Building Research Establishment Garston,HMSO, London.Ranta-Maunus, A., 1975. The viscoelasticity <strong>of</strong> wood atvarying moisture content, <strong>Wood</strong> Science Technology,9: 189–205.Ressel, J.B., 1994. State-<strong>of</strong>-the-art report on vacuum drying<strong>of</strong> lumber, Proceedings <strong>of</strong> the Fourth IUFRO International<strong>Wood</strong> <strong>Drying</strong> Conference, Rotorua, NewZealand, pp. 255–262.Resch, H. and Gausch, E., 2001. High-frequency current/vacuum lumber drying, Proceedings <strong>of</strong> the SeventhInternational IUFRO <strong>Wood</strong> <strong>Drying</strong> Conference,Tsukuba, Japan, pp. 128–133.Riley, S.G. and Haslett, A.N., 1996. Reducing air velocityduring timber drying, Proceedings <strong>of</strong> the Fifth InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Québec,Canada, pp. 301–308.Riley, S.G. and Holmes, W.S., 2001. Development <strong>of</strong> amicrowave in-kiln moisture content measurementsystem that measures individual boards, Proceedings<strong>of</strong> the Seventh International IUFRO <strong>Wood</strong> <strong>Drying</strong>Conference, Tsukuba, Japan, pp. 348–353.Salin, J.-G., 1989. Remarks on the influence <strong>of</strong> heartwoodcontent in pine boards on final moisture contentß 2006 by Taylor & Francis Group, LLC.


and degrade, Proceedings <strong>of</strong> the Second InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Symposium, Seattle,WA, pp. 4–6.Salin, J.-G., 2001. Global modelling <strong>of</strong> kiln drying: takinglocal variations in the timber stack into consideration,Proceedings <strong>of</strong> the Seventh InternationalIUFRO <strong>Wood</strong> <strong>Drying</strong> Conference, Tsukuba, Japan,pp. 34–39.Salin, J.-G. and Öhman, G., 1998. Calculation <strong>of</strong> dryingbehaviour in different parts <strong>of</strong> a timber stack, Proceedings<strong>of</strong> the 11th International <strong>Drying</strong> Symposium(IDS ’98), Kalkidiki, Vol. B, pp. 1603–1610.Salmen L., 1984. Viscoelastic properties <strong>of</strong> in situ ligninunder water-saturated conditions, J. Mater. Sci.,19: 3090–3096.Schiffmann, R.F., 1995. Microwave and dielectric drying, inMujumdar, A.S. (Ed.), Handbook <strong>of</strong> Industrial<strong>Drying</strong>, 2nd ed., Marcel Dekker, New York,pp. 345–372.Schniewind, A.P., 1989. Concise Encyclopedia <strong>of</strong> <strong>Wood</strong> and<strong>Wood</strong>-based Materials, Pergamon, Oxford.Shubin, G.S., 1990. Sushka i templovaia orbabotka drevesiny(<strong>Drying</strong> and heat treatment <strong>of</strong> wood), LesnaiaPromyshlenmust, Moskva.Siau, J.F., 1984. Transport Processes in <strong>Wood</strong>, Springer, Berlin.Siau, J.F., 1995. <strong>Wood</strong>: Influence <strong>of</strong> moisture on physical properties.Department <strong>of</strong> <strong>Wood</strong> Science and Forest Products,Virginia Polytechnic Institute and StateUniversity, Virginia, USA, 227pp.Simpson, W.T., 1992. Dry kiln operators’ manual, USDAFS Forest Products Laboratory, Agriculture HandbookAH-188, Madison, WI.Simpson, W.T. and Baah, C.K., 1989. Grouping tropical woodspecies for kiln drying, USDA FS Forest ProductsLaboratory Research Note RN-0256, Madison, WI.Skaar, C., 1988. <strong>Wood</strong>–Water Relations, Springer, Berlin.Smith, W.B. and Smith, A., 1994. Radio-frequency/vacuum drying <strong>of</strong> red oak: energy quality value,Proceedings <strong>of</strong> the Fourth International IUFRO<strong>Wood</strong> <strong>Drying</strong> Symposium, Rotorua, New Zealand,pp. 263–270.Spolek, G.A., Plumb, O.A., 1980. A numerical model <strong>of</strong>heat and mass transport in wood during drying’ 2ndInternational <strong>Drying</strong> Symposium, published in<strong>Drying</strong>’80: 84–92.Spolek, G.A. and Plumb, O.A., 1981. Capillary pressure ins<strong>of</strong>twoods, <strong>Wood</strong> Sci. Technol., 15: 189–199.Stamm, A.J., 1963. Permeability <strong>of</strong> wood to fluids, ForestProd. J., 13: 503–507.Stamm, A.J., 1964. <strong>Wood</strong> and Cellulose Science, RonaldPress, New York.Sun, Z.F., 2001. Numerical simulation <strong>of</strong> flow in an array <strong>of</strong>in-line blunt boards: mass transfer and flow patterns,Chem. Eng. Sci., 56: 1883–1896.Taylor, F.W. and Landoch, D., 1990. TDAL pr<strong>of</strong>iles <strong>of</strong>southern pine lumber during drying, Forest Prod.J., 40(10): 47–50.Tesoro, F.O., Kimbler, O.V., and Choong, E.T., 1972. Determination<strong>of</strong> the relative permeability <strong>of</strong> wood tooil and water, <strong>Wood</strong> Sci., 5(1): 21–26.Tesoro, F.O., Choong, E.T., and Kimbler, O.V., 1974.Relative permeability and the gross pore structure<strong>of</strong> wood, <strong>Wood</strong> Fiber, 6(3): 226–2<strong>36</strong>.Tetzlaff, A.R., 1967. An investigation <strong>of</strong> drying scheduleswhen kiln-drying radiata pine, B.E. Report, University<strong>of</strong> Canterbury, New Zealand.Timell, T.E., 1986. Compression <strong>Wood</strong> in Gymnosperms,Springer, Berlin, p. 1803 ff.Tremblay, C., Cloutier, A., and Fortin, Y., 2000. Determination<strong>of</strong> the effective water conductivity <strong>of</strong> red pinesapwood, <strong>Wood</strong> Sci. Technol., 34: 109–124.Trénard, Y., 1980. Comparaison et interprétation de courbesobtenues par porosimétrie au mercure sur diversesessences de bois (Comparison and interpretation <strong>of</strong>curves obtained by mercury porosimetry on differentwood species), Holzforschung, 34(4): 139–146.van Meel, D.A., 1958. Adiabatic convection batch dryingwith recirculation <strong>of</strong> air, Chem. Eng. Sci., 9:<strong>36</strong>–44.Vázquez, M.C.T., 2001. Tensiones de crecimiento en Eucalyptusglobulus de Galicia (España). Influencia de lasilvicultura y estrategias de aserrado (Growthstresses in E. globulus from Galicia (Spain). Influence<strong>of</strong> silviculture and sawing strategies), Maderas:Ciencia Tecnologia, 2(1): 68–89.Wagner, F.G., Gorman, T.M., Folk, R.L., Steinhagen, H.P.,and Shaw, R.K., 1996. Impact <strong>of</strong> kiln variables andgreen weight on moisture uniformity <strong>of</strong> wide grandfirlumber, Forest Prod. J., 48(11/12): 43–46.Walker, J.C.F., 1993. Primary <strong>Wood</strong> Processing Principlesand Practice, Chapman & Hall, London.Whitaker, S., 1977. Simultaneous heat, mass, and momentumtransfer in porous media: a theory <strong>of</strong> drying,Adv. Heat Transfer, 13: 119–203.Whitaker, S., 1998. Coupled transport in multiphase systems:a theory <strong>of</strong> drying, Adv. Heat Transfer, 31: 1–104.Zimmerman, M. H., 1983. Xylem Structure and the Ascent<strong>of</strong> Sap, Springer, Berlin.ß 2006 by Taylor & Francis Group, LLC.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!