Basics of Wood Drying
Basics of Wood Drying
net/publication/369768547
CITATIONS READS
10 4,058
3 authors:
Sohrab Rahimi
University of British Columbia
30 PUBLICATIONS 354 CITATIONS
SEE PROFILE
All content following this page was uploaded by Stavros Avramidis on 14 April 2023.
Contents
13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 679 specially designed chambers where under controlled
ambient conditions most of the water present in the
13.2 Heat and Mass Transfer in Wood . . . . . . . . . . . . . . . . . . . . . . 680
pores and cell walls of wood is removed. This water
13.3 Wood Drying Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683 extraction is necessary for further utilization, performance
13.3.1 Drying Air Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683
of wood, and longer service life. Wood drying is classified
13.3.2 Moisture Profiles Within Drying Wood . . . . . . . . . . . . . . . . . . 684
13.3.3 Drying Curve and Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 686 as a separation operation and water needs to be removed
from a multiphase system consisting of a complicated
13.4 Drying Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 686
13.4.1 Free Shrinkage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 686 solid structure. Therefore, the drying operation can be
13.4.2 Restrained Shrinkage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687 elucidated by treating it as a simultaneous heat, mass,
13.4.3 Drying Stress Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687 and momentum transfer set of phenomena involving
13.5 Drying Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 688 phase change of water within the complex material that
13.5.1 Conventional Kiln Drying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689 is wood. This chapter provides a basic and applied
13.5.2 Vacuum Drying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690 description of wood drying from the fundamental mois-
13.6 Modeling of the Drying Process . . . . . . . . . . . . . . . . . . . . . . . . 695 ture and thermal energy transfer to modeling and thereaf-
13.7 Drying Quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 699
ter, a description of the most common practices related to
13.7.1 Moisture Variability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 700 this followed by the wood products industry worldwide.
13.7.2 Shape Distortions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702
13.7.3 Checking (Cracking) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 702 Keywords
13.7.4 Casehardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
13.7.5 Discolorations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703 Timber Drying · Kinetics · Modeling · Methods ·
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704 Moisture profiles · Stresses · Defects · Variability ·
Quality
Abstract
13.1 Introduction
Drying is one of the most important processing steps in the
production of timber and many of the wood composites. One of the most important unit operations (basic process
During this process, freshly cut timber, veneer, strands, or steps) in the production of timber and many of the wood
wood particles of various shapes and sizes are placed in composites is drying. During this process, freshly cut tim-
ber, veneer, or wood particles of various shapes and sizes
are placed in specially designed chambers where under
controlled ambient conditions most of the water that is
S. Avramidis (*) · S. Rahimi naturally present in the pores and cell walls of wood is
Department of Wood Science, University of British Columbia,
removed. This water removal is necessary for further utili-
Vancouver, BC, Canada
e-mail: [email protected]; [email protected] zation of wood. Water removal or otherwise referred to as
“reduction of moisture content” is paramount for better
C. Lazarescu
FPInnovations, Vancouver, BC, Canada performance, longer service life, and transportation cost
e-mail: [email protected] reduction [1].
The drying of wood is classified as separation operation. 13.2 Heat and Mass Transfer in Wood
To obtain the desired solid phase and water content mix,
water needs to be removed from a multiphase system Removing water from a wood is a process that requires three
consisting of a complicated solid structure. Therefore, the controllable variables when the material is placed inside a
drying operation can be elucidated by treating it as a simul- chamber (dryer) and atmospheric conditions (to distinguish
taneous heat, mass, and momentum transfer set of phenom- from vacuum drying). The first is thermal energy that is
ena involving phase change of water within the complex expressed as temperature (T ) needed to excite the water
material that is wood. Drying involves simultaneous transport molecules and provide them with the kinetic energy required
of mass (water and water vapor), heat, and momentum and is to evaporate (heat of vaporization – phase change), but also
described by “basic equations” that take the form of simulta- the energy to move (mass transfer) within wood toward the
neous differential equations representing water and heat surfaces of the board. The second is reduced concentration of
fluxes. Mass transport has to overcome forces related to water molecules in the air expressed as relative humidity (H )
capillarity at the liquid phase level in the porous that will allow the development of the pressure or concentra-
sub-structure of wood and molecular forces in the cell wall tion differential in water molecules between the insides of the
nano-level. Multiphase flows in porous wood are more com- board and the air which will drive mass transfer. The third is
plex in description when anisotropy is considered. Wood is a air circulation expressed as air velocity (u) that improves the
highly anisotropic material where properties are directly evaporation rate by reducing the boundary layer’s thickness
affected by orientation of fibers and microfibrils within its and the external resistance of molecular flow. These three
solid matrix [2]. variables together can increase or decrease drying rate of
Wood diversity in structure and water content differ- wood like the throttle can increase or decrease the speed of
ences between sapwood and heartwood that are mostly a car. Researchers and practitioners interested in wood drying
prominent in softwoods could easily increase drying com- are always trying to find out the optimum combination of
plexity. Imagine wood boards with a 50 by 100 mm in those three that will allow maximum product quality for
cross section that are half sapwood and half heartwood and minimum drying time and energy [3].
with an average moisture content of 150% and 60%, Mass transfer can be considered at two scale levels: micro
respectively. That is a typical scenario with species such and macro. At the macro-level, the process is called bulk flow
as western hemlock (Tsuga heterophylla) on the North or capillary flow and it involves the flow of water and/or
American Pacific coast. Imagine also that there are two vapor under a pressure differential through the interconnected
such boards and one is a perfect flat-sawn, namely, annual void volume of wood, that is, lumens and pits. It is important
rings running parallel to top and bottom surfaces, whereas to emphasize here that although porosity is a measure of the
the other is quarter-sawn with annual ring orientation per- scale of void volume in wood, high porosity does not suggest
pendicular to top and bottom surfaces. Since boards in a high flow rates. The voids must be connected to each other
dry kiln release most of their water from the top and for this to happen; in other words, the pits should be “open”
bottom surfaces, that means that in the former case we to allow fluid flow from one lumen to the next. There are
have imbalanced transfer and evaporation due to half of many examples where wood species with very high porosity
it containing much more water than the other. In the latter value have a very low tolerance to fluid flow (refractory
case, since the two areas are side-by-side, the geometry of species) [3].
evaporation is quite different. Without further consider- Permeability is the measure of the ease with which fluids
ation of these two extreme situations, we can easily see are transported through a porous solid under the influence of
that by exposing these two boards to the same ambient a pressure gradient. A solid must be porous to be permeable,
temperature, humidity, and air flow conditions, the end but not all porous solids are permeable. Since permeability
result will be different moisture variation and stresses can only exist if the void spaces are interconnected by open-
within and between them thus contributing to one of the ings. Softwoods are permeable because pit pairs with open-
major problems of drying – moisture variability [1]. ings in the membranes connect the lumens of the tracheids in
Solving practical problems related to wood drying that softwoods. Above the fiber saturation point and at tempera-
usually involve multidimensional effects, two water phases tures above the boiling point, flow results from vapor pres-
that are time-space dependent with quite irregular interfaces sure generated within the wood. Under this condition,
move and a highly diverse material, require sophisticated capillary forces and wood permeability have dominant influ-
modeling and numerical procedures that represent quite a ence on moisture movement [4, 5].
demanding task. It is the objective of this chapter to provide Darcy’s law governs the flow of liquids and gases through
a brief description of the challenges persons responsible for wood in drying and steaming and the equations that describe
optimizing wood drying face [3] cf. ▶ Chaps. 6, ▶ 7, ▶ 8, this process under steady-state conditions are
▶ 9, and ▶ 23.
13 Basics of Wood Drying 681
VL m3 ðliquidÞ=m Pa s
tyloses that are very pronounced in white oak and greatly
kl ¼ ð13:1Þ
tAΔP reduce fluid flow. Aspiration and embolism are two phenom-
ena that also reduce permeability. The former is the result of
moving the torus/margo system of the bordered pits due to
capillary forces during drying and permanently attaching
VL P m3 ðgasÞ=m
kg ¼ Pa s ð13:2Þ them by hydrogen bonds to the cell wall thus closing any
tAΔP P
capability of communication between two adjacent lumens.
where kl and kg are the superficial permeability coefficients to The latter is the result of air bubbles plugging the pits thus 13
liquids and gases, respectively; V is the volume of flow (m3); not allowing further fluid flow through them. Both signifi-
L is wood length (m); t is time (s); ΔP is the pressure drop cantly reduce the value of K and hence reduce the uptake of
across wood (Pa); and P is the average pressure between the liquid preservatives and pulping chemicals during wood
two ends of wood (Pa). The superficial permeability coeffi- processing [4, 5].
cient is a measure of how fast fluid will flow through various The aforementioned equations apply to only steady state
wood species, but as it can be seen in Eqs. (13.1) and (13.2), or static conditions. These are the cases where the pressure
it is highly affected by the compressibility of the fluid (gases differential causing this fluid flow remains the same over
vs. liquids). A better way of expressing permeability levels is time. In many real-life situations such as timber drying or
by excluding the effect of the fluid and converting this prop- steaming, pressure conditions within the piece of wood
erty to one that is only affected by the structure of wood, that and/or within the air drying or steam heating medium do
is, specific permeability (K ) calculated by change with time. That is because drying schedules (combi-
nations of air T and H as a function of time of M) are dynamic
K ¼ μl kl ¼ μg kg ð13:3Þ and, therefore, boundary conditions are dynamic too. As a
result, these equations tend to “break down” and cannot
measured in m3m1, where μ is the viscosity of the measur- accurately predict unsteady state or dynamic fluid flow. In
ing fluid (liquid or gas in Pa s). It is obvious from Eq. (13.3) such case, a new dynamic equation is required to predict fluid
that K is independent of the type of measuring fluid and once fluxes such as
the superficial permeability is measured by, for example, dry
air or any fluid, the specific permeability can be then calcu- @P2 @2P
lated as a product of the latter and the viscosity [4, 5]. ¼ Dp 2 ð13:4Þ
@t @x
Although Darcy’s law is manly used to measure the per-
meability coefficient of various species so that fluid flow where Dp is the average diffusivity for hydrodynamic flow
under different pressures can be predicted, it must be empha- (m2/s) and equal to kg P=va and vα is wood porosity. Eq. (13.4)
sized here that the law applies to laminar flow that is not is useful for very approximate calculations because the coef-
always the case in wood. Further to laminar or viscous flow, ficient is a strong function of pressure with only the porosity
in wood with such small pore sizes, we may have turbulent term being constant [4, 6].
flow, nonlinear flow due to kinetic energy losses at the The micro-level of mass transfer involves water molecules
entrance of short capillaries (i.e., pits) and slip flow or Knud- and it is commonly called molecular flow or diffusion. In
sen diffusion due to mean free path of gases being larger than wood drying below the fiber saturation point, diffusion is the
the size of a pore (i.e., margo membranes). Total levels of mechanism of mass transfer through the cell walls and the
volumetric fluid transfer can then be a combination of two or empty lumens. Actually, in the grand scheme of wood struc-
more of the above types of flow [4, 5]. ture complexity and uneven moisture distribution in wet
Permeability is also a physical property that is greatly pieces of wood, diffusion starts at very early stages of drying.
affected by the anisotropic nature of wood. Longitudinal Regardless, molecular flow is divided in two types based on
permeability coefficients are 15 to 50,000 larger than the the medium through the water molecules migrate under a
transverse ones due to the orientation of the wood fibers. concentration gradient, namely, inter-gas through all pore
Not much difference exists between radial and tangential sizes and bulk through cell walls [4, 6].
direction; however, the K values of the former are slightly The flow of molecules is by activation and random and
higher due to the extra flow path provided by the rays. spontaneous “jumps” from one sorption site to another under
Sapwood is normally more permeable than heartwood and the influence of a concentration (or moisture) gradient.
the species anatomy will also affect K. A very common Bound diffusion through the cell walls is quite pronounced
example cited is the differences between red oak and white during drying and sorption, but diffusion through the lumens
oak. Both have the same density and porosity, but red oak is (inter-gas) is also an important component of the total
more permeable by a factor of 10,000 due to the lack of diffusive flux.
682 S. Avramidis et al.
Fick’s law describes the process of steady-state temperature (K), a2 is the volume fraction of cell wall (1 – va)
diffusion as where va is wood porosity. DT is a composite of the bound
and the intergas diffusion coefficients and it only refers to
wL
D¼ ð13:5Þ water molecules inside the drying wood piece. When the
tAΔC
molecules evaporate, they have to also diffuse through the
where D is the gross wood water diffusion coefficient so-called boundary layer that provides an extra resistance
(m2s1), w is the weight of fluid transferred (kg), and ΔC is (commonly named “external”) additionally to the “internal”
the driving force for diffusion, namely, the concentration resistance to molecular flow [3, 4–6].
difference (kg m3) and the rest are the same as in It must be emphasized here that the process of diffusion is
Eq. (13.1). Since moisture diffusion is probably the most very slow when compared to bulk flow, and thus, it is the
important for wood processing, ΔC can be substituted in controlling mechanism in drying when total processing time
Eq. (13.5) by the moisture content differential, ΔM, which is concerned. Unfortunately, there are no reliable methods of
in turn will make more sense from the view point of calcula- improving the diffusion coefficient of wood at an industrial
tions. After this, Eq. (13.5) changes to scale so that the drying speed of timbers can be increased.
Attempts have been made with pre-steaming, hot water and
100 w L surfactant solution dipping, ultrasonic field exposure, and
D¼ ð13:6Þ
t A ρw GM ΔM freezing as possible methods with some success mostly in
improving permeability than diffusion.
where GM is specific gravity of wood and ρM is water density In drying, total thermal energy is paramount for water flow
(kg m3). The gross wood diffusion coefficient is a compos- and evaporation. This is the sum of thermal energy absorbed
ite of bound water and inter-gas vapor diffusion, and it is and transferred through wood and is directly affected by
highly affected by moisture content, temperature, species, material thermal properties, namely, specific heat, and ther-
and type of wood, and, last but not least, direction. Specifi- mal conductivity. The latter refers to thermal energy transfer
cally, D is directly proportional to moisture content and from the wood surface to its geometric center. Furthermore,
temperature since higher moisture levels will swell the cell the convective part of the process that involves the thermal
walls thus creating new paths of flow, and the higher temper- energy transfer from the hot air in drying or steam to the
ature will provide the extra thermal energy required for faster surface of wood is also paramount in the entire thermal
molecular movement. The wood species and type (heartwood balance. As a result, convective plus conductive transfers
vs. sapwood) have an effect on D through density and chem- are the two mechanisms controlling the total thermal energy
ical composition that affect accessibility to sorption sites. in wood drying. The former is affected by the hot fluid
Finally, longitudinal D values are higher that transverse characteristics such as temperature, viscosity, and velocity
ones at low moisture content values in the hygroscopic whereas the latter is affected by wood structure and density.
range, but the difference decreases by increasing moisture Transfer of thermal energy through wood by conduction
content toward the fiber saturation point [4, 6]. takes place by molecular interaction in the cell walls.
As in the case of bulk flow, diffusion is also a dynamic Fourier’s law describes steady-state flow of heat as follows:
process in wood drying for the simple reason that moisture
content decreases with time and in within the drying timber. Qh L
As a consequence, the need of an unsteady state diffusion Kq ¼ ðW=mKÞ ð13:9Þ
tAΔT
equation is paramount.
where Qh is the quantity of heat transferred (J), and ΔT is the
temperature differential driving heat transfer (K). For wood,
@M2 @2M
¼D 2 ð13:7Þ Kq can be calculated by an empirical equation
@t @x
Again, Eq. (13.7) is only suitable for approximate calcu- K q ¼ Gð0:2 þ 0:0038MÞ þ 0:024 ð13:10Þ
lation because of the assumption of a constant coefficient. A
good approximation of the calculation of transverse DT of a where G is the specific gravity of wood at a particular mois-
drying piece of timber at moisture content within the hygro- ture content (M ). This equation is valid for moisture contents
scopic range is given by below 40%. For above 40%, the same equation can be used
with the only exemption of the coefficient of M that in this
7E 5 exp ðEb=RT Þ case is 0.0052 [4].
DT ¼ ð13:8Þ Eqs. (13.9 and 13.10) refer to a steady-state situation
ð 1 a 2 Þ ð 1 aÞ
where temperatures of wood surface and core remain the
where Eb is the water molecule activation energy (¼38,500- same with time, something that is not the case in real drying
290 M) (J/mol), R is the gas constant (8.31 J/mol K), T is situations. The same equation format applies to convective
13 Basics of Wood Drying 683
heat transfer with the only difference of Kq being replaced by 13.3.1 Drying Air Variables
hh that refers to the convective heat transfer coefficient. The
external resistance to heat transfer is also affected by the same Removal of water from wood in a controlled way is another
parameters as the diffusion one and faster transfer can happen approach of describing the separation of the liquid (water)
when air is replaced by steam and the air flow velocities are from the solid (wood) phase with the help of a gaseous
increased especially when the wood moisture content is phase (air). It is a complex process that takes time and
above the fiber saturation point (presence of liquid water energy and because wood is a highly hydrophilic and aniso-
in wood). tropic material that shrinks when it loses water, the predic- 13
As in the case of diffusion, in wood drying, temperatures tion of the final product quality is quite challenging
change with time and therefore Eq. (13.9) cannot accurately (Fig. 13.2). The whole wood drying process can be explored
describe the process. In that case, the unsteady-state or
dynamic format of Fourier’s law has to be used that takes
the form of
Heat transfer via
@T @2T convection from
¼ Dq 2 ð13:11Þ
@t @x air to wood surface
Kq
Dq ¼ ð13:12Þ Air flow
ρcpw
from two different viewpoints, namely, dryer and wood. In air flow direction every fixed number of hours is beneficial to
other words, the process needs to be analyzed from the the reduction of final moisture content variability between the
gaseous phase and the solid/liquid phase. The process is pieces of wood inside a dryer [1, 3].
regulated by three air variables: temperature, relative Dry kilns are units that are designed to accurately and
humidity, and velocity [3]. continuously control T, H, and u during a drying cycle. By
Air temperature is an indication and measure of the devising a plan or schedule that accurately sets the “T-H-u”
thermal energy in the air. The higher the thermal energy, combination as function of time or average moisture content
the larger the flux of it to the surface and center of the (rarely since there is no technology to monitor that) the
drying piece of wood. A greater flux will result in faster drying speed of wood can be controlled like the speed of a
thermal excitation of the water molecules in the lumens car when we press on the throttle. Since drying is partly
(free water) and cell walls (bound water). Enough energy science and partly art, the art at this point is to create a
to increase their kinetic energy and thus facilitate the proper schedule that will minimize total drying time and
evaporation of the former and breaking of hydrogen bond- maximize final quality of wood. That is mostly achieved
ing and diffusion of the latter. The net result is the loss of through individual experience, trial-and-error and with
moisture as a function of time and space and thus the some engineering “intuition.” An optimized drying sched-
initiation and continuance of the drying process. Greater ule will benefit the “bottom-line” of a wood processing
levels of temperature will accelerate drying measured by facility [3].
the change or moisture content with time and normally
called, drying rate (dM/dt). Caution should be exercised
with the levels of temperature used in dry kilns for 13.3.2 Moisture Profiles Within Drying Wood
research and experience have revealed that wood species
are very sensitive to those levels and many defects can In an ideal world, a piece of freshly cut wood will contain
develop because of that [3]. the same amount of moisture in every 3-D spatial point and
Air relative humidity is a measure of the amount of water the water will have the same capacity of escaping along all
in the air under a particular temperature level. It is normally three axes. Unfortunately, neither the former nor the latter is
expressed as percent relative humidity (H ) that is current the case in wood drying because of the water content var-
amount related to maximum amount the air can hold at a iability inside a tree and the anisotropic structure of wood. If
particular temperature and ranges between 0 and 100% or we examine a piece of log that is converted to boards, as
absolute humidity that refers to the amount of water per unit seen in Fig. 13.3, many possibilities of cross-sectional
volume of moist air (AH, kg/m3). The former is more com-
mon to designing ways or schedules for drying particular
species and sizes of wood products. Circulating air inside a
kiln that is very humid will result in a slow drying process.
A
Therefore, by controlling the air humidity, the vapor pressure
B
differential (Δp) between the inside of wood and the sur- C
rounding air is controlled and thus, the drying rate is regu- D
lated [3, 4]. E
Stagnant air of low H and high T has good drying capacity,
but moving air has a much better one. Therefore, controlling
the air circulation or velocity (u) will result in increased dM/ F
dt levels and shorter drying times. High air velocities are
mostly beneficial when liquid (free) water in the lumens is
present, namely, when the moisture content of wood is above
the fiber saturation point (Mfsp). Research has shown that
below Mfsp high u does not increase much the dM/dt and
thus, it can be reduced for realizing some electricity cost
savings. Air circulation also affects the external resistance
to drying, namely, the flow of water molecules through a
boundary layer attached to the surface of wood. The resis-
tance is proportional to the thickness of the layer and greater
air velocities tend to reduce its size and, thus, increase exter-
nal heat and mass transfer [3].
Fig. 13.3 Cross section of a simple Douglas-fir log hypothetical break-
Air direction is also important for drying balance and in down to produce 40 140 mm boards
dryers where there is large special difference, and reversal of
13 Basics of Wood Drying 685
Zone α
Zone β Zone α
Zone β
13
Fig. 13.4 The formation of wet-front and the two moisture zones across the thickness of a drying board
13.3.3 Drying Curve and Rate the third phase of 2FRP, where all points within the board are
below Mfsp (t4 in Fig. 13.5) diffusion becomes the controlling
In any industrial drying process as well as any drying variable and there is a further slowdown of moisture loss rate
research endeavor aiming to optimize the process and [3, 8–10].
improve product quality, knowing the average moisture con- The dM/dt change with respect to time in drying is shown
tent (M) of wood as well the drying rate (dM/dt) as functions in Fig. 13.7. The reader should keep in mind that the above
of time is paramount. Figure 13.6 shows a typical drying curves refer to ideal conditions of homogeneous Mi distribu-
curve as a function of time doe wood drying from an initial tion and drying that for most cases and in industrial applica-
M of 80% to a target M of 15%. Initially, the drop in M is tions that is not the case.
quite fast, but furthermore in the drying process, there is a A third-degree polynomial can well represent the drying
notable slow down. As seen in the Fig. 13.7, there is a distinct curve and it will have the form of
partition of the curve into three periods (phases), namely,
constant rate period (CRP), first falling rate period (1FRP), 3 2
M ¼ αt þ βt þ γt þ δ ð13:13Þ
and second falling rate period (2FRP). As it happens, these
three periods coincide with the three aforementioned α and β The first derivative of it will allow the calculation of the
phases. CRP is when the board is well above Mfsp and the drying rate as a function of time
drying rate is almost constant and controlled by the perme-
2
ability coefficient of the wood. In 1FRP the effect of the dM=dt ¼ 3αt þ 2βt þ γ ð13:14Þ
diffusion coefficient becoming a second controlling variable
is evident by the slowdown of dM/dt. As explained before, in The latter equation immediately reveals the complexity
of the process and the many assumptions and the uncer-
tainty of the wood behavior plus the moisture distribution
within wood during drying because no CRP phase is pre-
Mi = 80% dicted (no linear component of the curve). Calculation of
Constant rate period (CRP) dM/dt values per few hours directly from the raw drying
curve data and then, plotting them as a function of time will
1st falling rate period (1FRP)
allow with the use of two-step curve fitting to create the
M (%) 2nd falling rate CRP and the FRPs [3, 8–11].
period (2FRP)
Target
M = 15%
13.4 Drying Stresses
0 t (hours) Water removal from wood cell walls below the fiber satura-
tion point comes with what is commonly referred to as
Fig. 13.6 Drying board curve showing the three drying rate periods shrinkage. That is because open space is created within the
wall matrix that results in cellulose chains coming closer. The
shrinkage is not of the same level within a piece of timber
thus resulting in internal stresses. This chapter deals with the
fundamentals of stress development.
Drying rate or dM/dt (%/hr)
Tensile strain
The structural positioning of cellulose chains inside the Shell
largest layer of the cell wall (S2) causes most of the shrinkage
to happen along a direction perpendicular to the tree stem and Core
just a small amount along its length (~0.3%). The fibrils
making up the primary part of the cell wall tend to be oriented 13
spirally, almost at right angles to the fiber length, fitting 0
tightly around a water-swollen fiber. Any gain or loss in
Compressive strain
moisture content will be restrained by these fibril wrappings.
During a desorption process, at high moisture contents the Core Shell
restraining action is very efficient but further drying deter-
mines the non-crystalline region of these fibrils to begin to
shrink. The restraining is proportional with the microfibril
angle [15]. It is a fortunate feature because it limits the
amount of swelling and shrinkage but in the same time
creates stresses at cell wall level [12]. Another interesting 0 1
Drying time
aspect of the shrinkage phenomenon is that minor shrinkage
values are recorded before the wood reaches the fiber satura-
tion point (Mfsp). Recent research in this domain explains this Fig. 13.8 Strain development in shell and core during a drying
early shrinkage either by a large range of free-bound water schedule
coexistence [16] or the slight change of cell shape during the
water removal process [17, 18]. The controversy of free-
bound water coexistence was approached by splitting the develops more slowly and, as drying proceeds and the ten-
term Mfsp into two parts: hygroscopicity limit and the cell sion stresses advance inward, a decreasing amount of mate-
wall saturation limit [19]. The former was defined as the rial is subjected to them. Once the core has reached the Mfsp,
equilibrium moisture content (Memc) of a relative humidity and it starts to shrink, a reversed restraining phenomenon
slightly lower than 100% (to avoid full saturation) while the happens – the shell is now restraining the core. During this
latter was determined to be proportional with the ratio of phase tensile stresses develop at the interior and compression
shrinkage and specific gravity of wood. stresses on the exterior. Similarly, the compression stress
determines the apparition of compression set which repre-
sents a higher than normal shrinkage value (Fig. 13.8).
13.4.2 Restrained Shrinkage The phenomenon is more complicated because the set
appears simultaneously in different parts of a single piece of
As previously stated, drying stresses are minimized when wood driven by moisture gradient changes [12].
small wood specimens are dried. Thick and long specimens
are affected by the moisture gradient set up between shell and
core. In this case, the surface fibers will start to shrink and the 13.4.3 Drying Stress Components
overall volume of the piece will be reduced even though the
average moisture content is above Mfsp. The shrinkage of the In a comprehensive stress model, total strain (ε) rate is
surface layers is restrained by the wet (more plastic) inner assumed to be composed of thermal expansion/contraction
layers which show no tendency to shrink; these conditions strain (εT), free shrinkage and swelling (εα), elastic strain (εe),
determine a less-than-free shrinkage value called tensile set. and viscoelastic strain (εve) with its two components “creep”
Cracks are generally occurring during this early drying (εc) and “mechano-sorptive” (εms). Strain components are
stage [20]. generally assumed to be conceptually separable and the
As internal forces are balancing, a compensating compres- total strain can be written as:
sion stress appears on the inner layers. As drying proceeds,
the tension stresses advance inward, while a decreasing
T α e c ms
amount of material is subjected to increasing compression @ε=@t ¼ @ ε þ ε þ ε þ ε þ ε =@t ð13:15Þ
stresses in the core. The tensile stresses developed in the
adjacent layers are smaller than the one developed in the Wood expands if the temperature increases and shrinks if
outer slices [21]. The compression stress from the interior the temperature decreases. During the drying process,
688 S. Avramidis et al.
Compression
be negative [22]. Also, if the temperature is held constant or
the variation interval is very small, thermal expansion/con-
traction may be neglected (εT ¼ 0).
Under short-term loading, stresses below a certain
Stress
limit (called the proportional limit) produce strains that
substantially disappear when the load is released
Tension
(directly recoverable). This strain is called the elastic
strain. The region of elastic behavior is determined
experimentally through direct evaluation of stress-strain
diagrams and can be modeled using spring elements Days drying 5 10 18 28 36 50
(Hooke method). Mean M (%) 77 64 50 35 17 10
Any stress extended over a period of time will develop an
Kiln Memc (%) 18 17 13 8 2 2
additional strain named creep. Wood, with its complicated
structure at macroscopic, microscopic and molecular level,
develops two types of creep strains during drying, namely, Fig. 13.9 Distribution of residual stresses at various stages of drying
time-dependent creep and mechano-sorptive creep. The time- throughout the 2-inch thickness of a red oak drying board, based on
McMillen 1958 [21]
dependent creep appears in materials subjected to a constant
load and it is typically characterized by a family of stress-
strain diagrams. For wood, these stress-strain diagrams can usually applied in green specimens at the beginning of
be developed under constant temperature and moisture con- the test.
tent. The mechano-sorptive creep, also interpreted as an Drying stresses measured either by slicing and curvature
accelerated creep due to moisture content changes, is the measurements were initiated by Peck (1940) [36] and latter
result of transient redistribution of stresses and it is associated perfected by Kuebler (1960) [37]. The released deformation
with moisture content changes that cause the rupture of was measured with transducers positioned against the end of
hydrogen bonds. These bonds will reform in a different the tested specimens [28]. The stress is calculated as the
location under the bias of the applied stress. The quantifica- product of the strain and Young’s modulus at a particular
tion of this mechanism can be done using mathematical temperature and moisture content level. A stress diagram
models of rheology (from the Greek “ρεoλoγία”) – the obtained by these methods will look, more or less, like the
study of the time-dependent stress-strain behavior of one illustrated in Fig. 13.9. After 5 days of drying the outer
materials. layers of the wood are under severe tensile stress. The long
Under sufficiently low moisture content and temperature, duration of drying process will develop a tensile set and wood
wood behaves much like a brittle material (linear elastic will shrink less-than-free shrinkage value. Often these
manner). Under intermediate moisture content and tempera- stresses create small surface checks, which may cause prob-
ture wood exhibits viscoelastic behavior. At higher stress lems if the surface is to be coated with paint or clear finishes.
levels, or in fluctuating environmental conditions wood will
have a nonlinear viscoelastic behavior with considerable
plastic deformation [23]. 13.5 Drying Processes
Recent research in developing constitutive models of
drying stresses was done by Rice and [24–35]. Most of Drying of wood boards can be done with two different ways,
the authors built Burger, Kelvin, N Kelvin, or Maxwell namely, in the presence of a hot fluid (convective drying) that
models based on series or parallel combinations of spring will transfer heat to the wood and take away the evaporating
and dashpot elements. An original approach was done by water molecules, and under partial atmospheric pressure
Moutee et al. (2007) [33] who built a rheological model of (vacuum drying).
wood cantilever for modeling the creep behavior and stress Convective drying can also be subdivided into low/high
in various moisture content conditions and various load temperature conventional and dehumidification drying with
levels. Others carried their research by starting from some variations, whereas vacuum drying can be subdivided
models made on small specimens, which were later used into conductive, superheated steam, and dielectric [3, 38, 39].
to characterize the drying stresses in full boards. This Last, there is the category of solar drying that is a hybrid
approach implied the study of the drying process under between natural air drying and conventional kiln drying with
an imposed stress, either tensile or compression, which is minor control over the drying variables. In the following
sections, specific features of the different drying processes
13 Basics of Wood Drying 689
φ = 1.0
13.5.1 Conventional Kiln Drying
h=
40
20
0K
13
h=
The drying is done in a specially designed chamber called
J/k
10
A
0K
g
kiln or dry kiln or dryer. Although there are differences in
J/k
detail between kiln manufacturers, essentially, it is a box 0
g
made with thermally insulated aluminum panels attached to 0 40 x 80
a steel frame and fitted with overhead hardware specially g/kg dry air
designed to control air temperature, relative humidity, and
circulation (speed and direction). Until recently, the majority
of the kilns were of the batch type, namely, doors will located Fig. 13.10 Mollier-h,x-diagram for moist air at atmospheric pressure
(simplified). Example: A conventional kiln drying process and
either in one end or both, kiln will be filled with the stickered air-exchange from outside is indicated by the thick black line, starting
board packages, and the drying run will commence by fol- with the state of the air from outside (A), heating up the air and the
lowing a time-based drying schedule. When the run is over, indoor climate (X) – compare box above
the wood is taken out and a new back for green wood enters
the kiln for processing [3, 8].
The main principle of conventional kiln drying in order 239–26 ¼ 213 kJ has to be provided via the heat-coils to hold the
to exhaust moisture of the circulating air is based on the chamber temperature (without considering other heat losses).
Mollier h-x diagram (Enthalpy-Humidity Mixing Ratio) This heat amount includes the heat for heating up the air
diagram, developed by Mollier (1923) [40]. The following (38 kJ) and the heat for evaporation of the wood moisture.
practical example (see box below) helps to understand the Progressive (or continuous) kilns are much longer chambers
principle of the dehumidification of the kiln-drying cham- where the air conditions are also precisely controlled but are not
ber. The Mollier diagram is a graphic representation of the the same from one end to the other. The kiln is divided into
relationship between air temperature, air moisture content, zones where conditions are different. The wood sits on carts that
relative humidity, and specific enthalpy and therefore in turn sit on rails that enter the kiln from one end and slowly
describes the properties and the state of the moist air, move toward the other (exit) like a train in a tunnel. As the
which is used as a circulating medium for transport both wood moves, it goes through the various zones where condi-
heat and vapor during the drying of timber. The diagram tions in the beginning are low T and high H and as the wood
helps us to understand how air changes state when we cool moves through the system and consecutive zones, conditions
and heat it. become harsher, namely, higher T and lower H. Wood moves at
A specific kiln drying schedule as shown in ▶ Chap. 23 a calibrated speed so when it exits, theoretically speaking, the
“Sawn-timber Steaming and Drying” is the controlling basis Mf is at target. The problem with this method is that it is good
of the variables such as kiln air temperature, relative humidity for species of Mi that does not differ much plus, and it will not
(RH), and velocity as shown in Fig. 13.10. allow for implementing stress relief by conditioning and Mf
In Fig. 13.10, an example for a theoretical path of air in- and standard deviation reduction by equalizing, something that
outside a drying kiln is given. As parameters are chosen: outdoor batch kilns will allow the mill to carry out. Batch kilns will
air 10 C/80% RH, actual chamber climate 60 C/50% RH. The also provide much more flexibility to production logistics for
support air, entering the kiln dryer via the roof vents, exhibits a allowing to dry various species and thicknesses [1].
moisture content x of approx. 6 g/kg. After heating up the air A third drying system that falls under this category is the
(without mixing with the circulating air) the RH would decrease so-called dehumidification drying. Although it is very similar
to 8% (ϕ ¼ 0.08); after mixing with the circulating air, the to a batch kiln, upon closer look there are couple parts that are
exhaust air is released from the chamber with 60 C/50% RH missing, namely, heat exchangers and roof vents. This is
(x ¼ 68 g/kg). For each kg air transported via the valves (corre- because the H in the circulating air inside the unit is con-
sponds to approx. 1.2 m3 of air) a Δx of 68–6 ¼ 62 g water as trolled by a dehumidification (air conditioning) unit that is
vapour per 1 kg dry air is released from the chamber (exhaust air). based on the.
The incoming support air has an enthalpy h of 26 kJ/kg, the Carnot cycle, and it removed the extra moisture picked
exhaust air 239 kJ/kg; for an air exchange, a heat amount of Δh from wood by the circulating air in this closed system that is
690 S. Avramidis et al.
a
100
Evaporator
superheated steam, and dielectric. Their difference is in the
way the thermal energy is transferred to wood [38, 39].
Fig. 13.11 Schematic of a dehumidification kiln (a), and the dehumid-
ification unit (b). Cold and humid air (dash) over cold evaporator will
Conductive Vacuum Drying
lose moisture and reduce H, then, over hot condenser will pick up
thermal energy [22] In conductive vacuum drying (CVD) the wood is in conduct
with heating elements (hot plate) that are normally made out
of aluminum and stacks of wood are laid between the hot
plates that are normally heated by a hot fluid flowing through
then heated and re-sent into the dryer as seen in Fig. 13.11.
them (Fig. 13.13). This system provides uniform heating of
Because of this, no vents are needed to control excess air
the lumber and good control of the temperatures used. How-
humidity and no heat exchanger to control air T. Therefore,
ever, kiln loading and unloading are time-consuming, if done
the main difference between a conventional batch kiln and
manually, and plates require periodic maintenance or replace-
dehumidification one is the fact that in the former the vapor
ment, adding to the cost. Some kiln-manufacturing compa-
differential between wood and air that drives wood moisture
nies offer automatic systems for stacking the boards and hot
out is created mostly by reducing partial vapor pressure in the
plates. In convective drying, moisture moves along the thick-
air through dehumidification whereas in the latter by raising
ness of the boards and evaporates from the top and bottom
the temperature of water in the wood. Dehumidification
surfaces. However, in conductive vacuum drying, free water
dryers are all electrical, and they take longer to dry wood
in lumens above Mfsp and bound water in cell walls below
due to their lower drying Ts and have no conditioning capac-
Mfsp move along the length of the boards through the
ity due to the lack of steam producing unit. However, they are
interconnected capillary structure of wood. Longitudinal per-
simple to operate, have no need for a boiler, and the capital
meability becomes paramount factor concerning drying
costs plus energy costs are lower [3, 8]. ▶ Chap. 23 will
speed and quality. Refractory species do not perform well
provide the reader with much more information about these
and this process is more fit for drying high or medium
three systems of convective drying.
permeability species [38, 41–43].
Generally speaking in vacuum drying of wood, the boiling
point of water is reduced by drawing a vacuum via a high
13.5.2 Vacuum Drying
power vacuum pump, allowing for drying at lower tempera-
tures than conventional drying. As in dehumidification dry-
As the name implies, the drying in this case is carried out at
ing, vacuum reduces the water vapor pressures to extremely
reduced atmospheric pressure conditions. The basic principle
low levels thus creating a very high pressure gradient from
of a vacuum dryer is the fact that the boiling point of water is
the center toward the surface of wood. Thus, the benefits of
below 100 C when the air pressure in the autoclave is below
high-temperature drying are achieved, that is, less warp and
the atmospheric air pressure of about 1000 mbar (Fig. 13.12).
much reduced drying time, but at lower temperatures. Similar
There are three major vacuum drying methods: conductive,
to the “wet-front” described in 13.3.2, here there is a
13 Basics of Wood Drying 691
y
Electric
Timber
Ti
Tim
im ber
heating
plates
Vacuum
E H
chamber
E0
x
13
Wood boards Conductive vacuum dryer H0
Moisture and temperature Heat transfer z
Autoclave measurement
Heating
plate Fig. 13.14 Diagrammatic illustration of a plane electromagnetic wave.
Timber E and H represent the electrical and magnetic components of the wave;
Eo and Ho are their respective amplitudes
RF Dryer Drop
Green Conventional Planer
sort
lumber kiln
wets
RF voltage
Generator
RFV
13
[49]
+ Electrode –
Wood
a b
– Electrode +
c d
Dielectric index (DI)
eliminate oxidative reactions and potential surface Superheated Steam Vacuum Drying
discolorations A hybrid between conventional and vacuum drying is the
• Can dry very thick timbers (Fig. ▶ 10.15a-d) and can also so-called superheated steam vacuum drying or SSV. In this
dry short pieces for finger-jointing since no stickering is case, superheated steam (water vapor at temperature higher
allowed than the boiling point) is used under low-pressure conditions
• Provides a dried product with significantly less variation and forced through stickered layers of lumber, heating by
in final moisture content within and between boards convection, and a continuous vacuum-drying process can be
• Is purely electrical with high efficiency and low drying achieved. This process is known as SSV or convective vac-
costs that in many cases can be up to 50% of the drying uum. Superheated steam has better heat transfer properties
costs in conventional dry kilns than hot air at the same temperature; however, steam under
• Fully automated process with excellent drying run termi- vacuum has lower heat capacity (due to lower density) and
nation point and minimum labor involvement drying rates are lower than with hot moist air as in conven-
• Friendly to the environment with low carbon footprint tional drying. This can be compensated by circulating air at
since the collected water from wood can be treated before high speeds, of about 10 m/s, and by frequent fan reversals.
discarded and if electricity is produced by green means, SSV has been credited for faster drying times and no surface
drying is also green discoloration.
As is seen in Fig. 13.19, there is a significant reduction of
However, the main disadvantage of the RFV dryers is the the drying times between convective and SSV drying for
very high initial investment to design and build an RFV
system. These systems are much smaller (e.g., capacity of
75 m3), than conventional large-size kilns; however, drying a
speeds are much greater so that comparable volumes can be
m
mm
m
27 m
80
27 m
processed in the same time, but with all the quality benefits
m
52
65
discussed above. Large commercial dry kilns can reach m
70 m
capacities almost ten times that size; however, costs of drying 52
mm
Mg (%)
m
65 mm
65 mm
90 mm
90 m
hardwoods and softwoods. For example, oak boards of thickness of lumber do not greatly affect the drying rate.
52 mm in thickness will dry from an initial moisture content When moisture drops below Mfsp, moisture movement is
of 60% to 20% in about 61–62 days whereas in SSV it will dominated by diffusion. This mechanism also works in the
take about 15 days, a significant drying time reduction. Same sapwood of impermeable species, in which moisture
trend can be observed for softwoods. The time reduction passway is blocked by aspirated pits and so on. It is worth
becomes larger as the thickness of wood increases [38, 39, mentioning that lumber drying is intricate and seldom con-
69–72]. ▶ Chap. 23 will provide more information about the trolled by a single process.
SSV process and hardware. Many empirical and graphical-analytical methods have 13
been proposed to model wood drying over last decades.
Bramhall (1976) [84] proposed one of the first purely
13.6 Modeling of the Drying Process empirical lumber drying models as follows:
were sensible, though the predictions of the drying model model of drying time at the low temperature in batch
were not trialed clearly. kilns is
Milota and Tschernits (1990) [87] developed a model for
drying of loblolly pine (Pinus taeda) at an elevated temper- 0:233 π2 ðMs Memc Þ þ 0:81ðM0 Ms Þ
ature that included Tdb of 82 C to 132 C and Wbd of 60 C to t¼ 1þ ln B
Dm 4Bi 2
M Menc
93 C. This empirical drying model includes two stages, that
ð13:28Þ
is, initial constant rate drying period and falling rate period.
Equations 13.18 and 13.21 are offered to model the former
where Dm is mean moisture diffusion coefficient (m2/s),
and latter stages, respectively.
Bi is mass-transfer Biot number (βb/De), B is coefficient,
Ms is surface moisture (%) when the moisture mean is M at
Fcr ¼ ½0:01208 þ 0:00679ðT db T wb Þ 2:482 time t, Me is corresponding equilibrium value (%), and M0
5 2 is moisture value in core at the start of drying (%).
10 ðT db T wb Þ f v ð13:23Þ
In case of a three-step drying schedule for beech (Fagus
where Fcr is constant drying rate (kg/h.m2), Tdb is dry-bulb sylvatica), Eq. 13.23 changes into a practical form as
temperature ( C), Twb is wet-bulb temperature ( C), and fv is a represented in Eq. 13.24:
parameter related to air velocity (m/min), which is evaluated
through either of the following equations:
0:81ðM0 Me1 Þ M Me2 M Me3
t¼C ln 02 þ ln 03
0:3 M1 Me1 M2 Me2 M3 Me3
v
fv ¼ for v < 6:8 m=s ð13:24Þ ð13:29Þ
1329
0:5
v where M1 ¼ mean moisture (%) at the ith stage of drying
fv ¼ for v > 6:8 m=s ð13:25Þ
1329 program, M01 is moisture value (%) in the core at the
Drying rate is evaluated for the drying rate by fitting to the start of that step, Mei is corresponding equilibrium mois-
curves of drying rate as a function of moisture through the ture (%) for the specified relative humidity, and C is
next equation: coefficient (s1) depends upon the board thickness, the
width of the stack, and air speed in the kiln. Using
n n 1=n monograms, the width of the stack and air speed in the
Fm ¼ ½ST ðM Memc Þ þ ðFcr Þ ð13:26Þ
kiln are determined.
where Fm is falling drying rate (kg.m2/h), ST is the slope of Normalization of drying-rate curves has extensively used
the asymptote to the curve of drying rate as a function of to grasp the characteristics of industrial drying plants. Van
moisture at equilibrium, and n is a parameter that deter- Meel (1958) [90] put an analysis for the behavior of convec-
mines the shape of the drying curve at different conditions. tive batch drying and presumed that there is a corresponding
Olek et al. (1994) [88] came up with a new empirical particular drying rate relevant to unhindered drying rate in the
model for drying of European beech (Fagus sylvatica) and first drying period, which is independent of the external
proposed the following equation: drying conditions. Relative drying rate is clarified by the
following equation:
dM a2
¼ a1 þ ð13:27Þ
dt ð100MÞn þ a3 f ¼
Nv
ð13:30Þ
bv
N
where a1, a2, a3, and n are fitted parameters to the experi-
mental data and they are diverse with drying conditions. where Nv is drying rate (% s1) and Nbv is the rate in the first
Aside from empirical methods, several researchers have period of drying limited by the external convection (% s1).
worked with graphical-analytical methods for wood drying Characteristic moisture is defined as follows:
modeling. In 1971, Moscow State Forestry University
developed methods upon which, onward standards for M Memc
ɸ¼ ð13:31Þ
lumber kiln drying have been approved. Two versions Mcr Memc
were put up. One is based on graphical-analytical solutions
of the theoretical equations and the other is based on using where M is volume–averaged moisture (%), Mcr is
nomograms and tables for specific drying conditions moisture at a critical point (%), and Memc is moisture at
described in the standards [89]. As an example, the equilibrium (%).
Ashworth (1977) [91] conveyed that characteristic drying
curve is appropriate to describe kiln drying behavior of
13 Basics of Wood Drying 697
softwoods, given that the curve is according to small-scale relative drying rate ( f1) are expressed, in turn, in terms of
drying tests with a sample of the same thickness. Likewise, absolute moisture (M) and actual drying rate ( jw), as follows
Nijdam (1998) [92] did a detailed analysis of the mass-
transfer processes on softwood boards and reached the iden- M M2
ϕ1 ¼ , M2 < M < M1 ; ð13:35Þ
tical conclusion. Keey and Walker (1988) [93] claimed that M1 M2
the appearance of the characteristic drying curve ( f as a
jw jw2
function of ɸ) displays the wood permeability. The curve is f1 ¼ , jw2 < jw < jw1 : ð13:36Þ
jw1 jw2 13
concave-downwards in permeable wood while concave-
upwards in impermeable wood. The former shape exhibited The second falling drying-rate period applies to the step of
often in the drying of sapwood of softwood with fast moisture drying in which, lumber reach equilibrium moisture and
movement, whereas the latter shape exhibited in hardwoods drying comes to the end. Normalized moisture (ɸ2) and
with a high amount of extractives, which hinder water relative drying rate ( f2) are expressed as follows:
movement.
Kayihan (1993) [94] defines an empirical model that M Memc
ϕ2 ¼ , M2 > M > Memc ; ð13:37Þ
includes the impacts of the external conditions (through a M1 Memc
constant-rate drying parameter) and the internal moisture-
jw
transport process (through a diffusion parameter b) to esti- f ¼ , jw2 > jw > 0: ð13:38Þ
jw2
mate the rate of the change of the average moisture (M) as
below: The parameters M1, M2, and jw2 are characteristics func-
tions for each species.
dM bM Although the pass-ways for diffusion and capillary flow
¼h 3 i1=3 ð13:32Þ
dt (permeability) of water in wood are the same, their contribu-
1 þ bM=a
tion in these two mechanisms are varied because the rate of
diffusion is proportionate to its cross-sectional (πr2) area
In Eq. 13.32, the activation energy dependence of the while the rate of permeability for laminar flow in tubes is in
drying behavior on temperature is involved in parameter b proportion to πr4, based on the Hagen-Poiseuille equation.
as expressed in the following equation The effect of small pores (tubes) has a greater contribution to
diffusional processes than to permeability. Consequently, the
2700 diffusion of water in the swollen cell wall is more significant
b ¼ b0 exp ð13:33Þ
T than across the pits as the area available in the cell wall itself
is much bigger than the openings in the cell wall. Figure 13.20
where T is temperature ( K). presents a general overview of different pass-ways through
Keey and Pang (1994) [95] described the drying kinetics softwoods [5].
of radiata pine (Pinus radiata), using a characteristic drying In actuality, there is no single “best” model and there is a
curve, and analyzed the mechanisms of moisture movement considerable variability among the predictions of most
by which three steps in the drying of sapwood and two steps models mainly owing to the uncertain coefficients for the
in the drying of heartwood were identified, respectively. A
dual characteristic curve represents drying kinetics, covering
the two falling-rate phases, in order to reach simplified
expressions for the drying kinetics. The constant-rate period Cell walls
(Bound/liquid)
is only for sapwood. This period starts from initial moisture
content (M0) down to the first critical point (M1) and Fiber cavities Pit pores
expressed as follows: (Vapor/liquid) (Vapor/liquid)
Pit chambers
0
jw1 ¼ φK 0 Hs Hg ð13:34Þ (Vapor/liquid)
models, various degrees of simplifications, and different hydraulic conductivity (m s1), and μ is vapor viscosity
methods to solve heat and mass transfer problems. In this (Pa s).
regard, Kamke and Vanek (1994) [96] concluded that sophis- Various coefficients in these equations are not constant
ticated models do not essentially perform better than simple with respect to temperature and moisture. Even if these
ones do unless the required physical property data is ade- coefficients are considered as constant, solving the equations
quate. In other words, a simple model may work quite satis- is intricate.
factorily, provided that sufficient physical property data is Hardwoods and the heartwood of softwoods show quite
obtainable for both the species and drying conditions within different moisture profiles compared to sapwood of softwood
the range of which, a model is developed. Furthermore, the does, owing to the reflection of the relative permeability and
models that discrete the transport mechanisms are preferred ultrastructure of the different wood species. Shubin (1990)
to predict stress and strain behavior or for other purposes that [89] pointed out that moisture distribution in drying beech
need detailed heat and mass transfer information. (Fagus sylvatica) and the heartwood of pine (Pinus sp.)
Notwithstanding the fact that describing moisture move- illustrated diffusion-like parabolic profiles. In contrast, the
ment based on diffusion are quite good in many cases, much drying of larch (Larix sp.) showed steep moisture gradients
attention has recently been paid to models by which various that are consistent with respect to the presence of an evapo-
transport mechanisms involved in the moisture movement rative plane near the surface. The moisture profiles in the core
can be considered. Such models are aimed to provide a better of a permeable species are rather flat at the beginning of the
understanding of lumber-drying process, thereby without the drying because assumingly the moisture moves as a liquid
need for extensive pilot-scale trials; we can improve drying under a pressure gradient. In the sapwood of softwood, the
schedules and maintain the lumber quality. Many recent permeability decreases gradually with a slight steepening of
studies described the lumber drying process by considering moisture gradients in the core. This trend occurs perhaps due
different transport mechanisms for bound water, free water, to the random aspiration of the bordered pits that leads to
and water vapor [97–101] and normally each piece of lumber blocking the liquid pathways [92]. McCurdy and Keey
taken into account as a homogeneous, hygroscopic, and (1998) [103] dried Pinus radiata at 120 C and reported
porous material. moisture at the core of the specimens to be relatively uniform.
From mass and energy balance over a volume element, a It is well known that the main goal of lumber drying is to
set of equations for the change of moisture (M), temperature reduce Mg to Mt. Practically, a large number of the pieces of
(T ), and pressure (P), with regard to time (t) is derived and lumber dry in a batch kiln simultaneously, which imposes a
represented as follows: great deal of variability of Mi to each batch of drying, which
consequently leads to an unacceptable (under-dried and over-
dried) lumber. For this reason, a systematic evaluation of Mf
@M 2 2
¼ Dρb ∇ M þ Dδρb ∇ T ð13:39Þ can help the kiln’s operator to make an objective and sensible
@t decision in terms of the optimization of the drying schedules
where D is diffusion coefficient (m2 s1), ρb is basic wood [104, 105]. For this purpose, stochastic simulation has been
density (kg/m3), and δ is thermal gradient coefficient used by several researchers to improve the prediction of Mf.
Stochastic simulation is a kind of simulation which is
designed to predict the outcome of systems (processes)
@T 2 εΔH LV @T containing many similar, but not identical components
¼ λ∇ T þ ð13:40Þ
@t CX @t [106]. In the probabilistic analysis, understanding of the Mf
is crucial on the grounds that the classical parametric statis-
where λ is thermal conductivity (kg.m/s3. K), ε is phase
tics are according to specific probabilistic distribution
transformation criterion (desorption and evaporation), ΔHLV
assumption [107]. Researchers have rigorously developed
is the heat of phase transformation (J mol1), and Cx is
numerical stochastic models to simulate M dispersion during
modified heat capacity (kg m2 K1 s2) of the unit mass of
and post-conventional drying [106, 108, 109]. In addition,
dry woody matter and its associated moisture
Cronic et al. (1997) [110], Cronin et al. (2002) [111], and
Cronin et al. (2003) [112] investigated on the probabilistic
analysis and design of the industrial timber drying process
@P RT K @M P @T
¼ ∇ ∇P þ ερb : ð13:41Þ and schedules. Elustondo and Avramidis (2005) [106] com-
@t ΨMv μ @t T @t
pared three methods for lumber drying simulation, namely,
where R is universal gas constant (J mol1 K1), Ψ is the Cronin’s method, Monte Carlo Method, and Proposed
fraction of tracheids or vessels occupied by moisture vapor method. After comparing these three methods, Elustondo
(%), Mv is the molar mass of the vapor (kg mol1), K is and Avramidis (2005) [106] claimed that proposed method
is able to be implemented to an iterative type of calculations,
13 Basics of Wood Drying 699
Piece count
way of the definition and value since it can be alleviated and 8 80
even, fully recovered, moisture variability between boards 6 60
(Mbb) as mentioned (Sect. 13.3.2) before is another quality-
4 40
related problem that needs to be considered firstly.
2 20
0 0
13.7.1 Moisture Variability 10 30 50 70 90 110 130 150 More
Initial moisture content (%)
Conventional wood kiln drying is a process of random
b 25 120
Piece count
80
more, ambient air conditions tend to not be completely 15
homogeneous throughout the kiln where T and H usually 60
oscillate in time and space due to control system limitations. 10
40
As a result, board positioning inside the kiln also affects
5 20
drying rate, and surface or internal checks that appear in
some or all of the boards could change their internal heat 0 0
and mass transfer characteristics. 0 4 8 12 16 20 24 28 More
Figure 13.2 shows the reasons why a large variability in Initial moisture content (%)
green moisture content (Mg) among boards of the same kiln
charge is always present, especially in softwoods, and it is Fig. 13.22 (a) Green moisture distribution for hemlock (average ¼
mainly due to the moisture differences between sapwood and 82%, StDev ¼ 38%, COV ¼ 46.3%) [108, 109]. (b) Moisture content
heartwood. Once the boards are in the kiln, random differ- distribution after kiln drying (average ¼ 14%, StDev ¼ 2.6%, COV ¼
18.6%) [108, 109]
ences in Mg and wood properties induce individual boards of
the same charge to dry at a different rates. As a result, board
M after drying fluctuates within a wide range of values. In Drying objective: move
industrial drying, this problem is further aggravated by non- the mean and compress
uniform kiln drying conditions due mainly to hardware cal-
Board piece count
the spread
ibration and malfunction issues. Typical sources of
nonuniformity in an industrial kiln involve poor board pack-
age arrangement; unbalanced hot air distribution from the
heaters; unbalanced fresh air distribution from the vents;
incorrect air flow distribution from the kiln fans; excessive
temperature drop across the load; and excessive cold air
Overdried
)
%
%
%
(%
80
Dried
Wets
Group A Group B Wets
13
13.7.2 Shape Distortions content, above 40% [1]. The incorrect stacking procedure can
be a contribution to bow. Both lumber and stickers must have
Shape distortions (warp) are any diversion of the face or edge a uniform thickness. Careful stickering practices, such as
of a lumber from the flatness or any edge that is perpendicular maintaining good vertical alignment and assuring no stickers
to the adjacent face or edge [1]. It is a result of the anisotropic are up on edge, are a contribution for decreases of bow.
shrinkage and presence of irregularity in wood structure as Foundations for green lumber piles, whether in the kiln,
high SOG, juvenile wood, or compression wood. It can occur pre-dryer, or air-drying yard, must be flat. Drying wood too
in one or more forms which are cup and diamonding, bow, slowly will exacerbate bow. Fast drying, especially at high
crook, and twist. Use of proper kiln schedule can help to moisture content, can reduce the amount of bow, but fast
reduce them (Fig. 13.27). drying results in checks.
Bow is warp along the length of the face of lumber. It Many authors [124–126], reported twist as the most severe
occurs when one face of lumber shrinks more in length than problem in the construction industry. Twist is the turning of
the other. Bow causes the lengthwise curvature of a piece of the four corners of timber cross-section, so that they are no
lumber, such that it resembles a bow used in archery. Crook, longer in the same plane. In the wood science literature
or side bend, is warp along the length of the edge of lumber, [127, 128] it is often related with the fibril orientation and
and in squared wood products, baby-squares for example, distance from the center of a log, but some authors [129, 130]
bow is a general term. A localized crook often due to a knot is showed no or little correlation between twist and grain angle.
called kink. Twist does not occur directly after sawing but arise during
Bow can be developed in the green timber immediately seasoning of the wood [120]. Lumber containing these grain
after sawing as a result of growth stress release [120]. Usu- characteristics can sometimes be dried better with the proper
ally, wood closer to the center of tree shrinks longitudinally stacking procedures [1].
more along the grain than mature wood [121]. For that
reason, the part of a lumber with juvenile wood, which is in
disagreement with the mature part, will cause the lumber to 13.7.3 Checking (Cracking)
try to bow excessively during drying. If bow is significant,
that type of pieces will be classified as degraded after drying. Surface checks develop on the surface of a piece of lumber
The critical lumber could also be sawn from a crooked log or during drying. They are caused by tension stresses that
in wood around large branches [122, 123]. The wood cells in develop in the outer part of lumber as the drying progresses.
the resulting lumber are oriented at an angle, causing longi- When the drying stresses exceed the tensile strength of wood
tudinal shrinkage and a tendency to bow. perpendicular to the grain they start to emerge. It means that
Uniformity of thickness in a drying charge can also pro- these stresses may become great enough to tear the outer
duce bow problems. Improper saw feed speeds or lapses in fibers apart which later will cause the surface checking
saw maintenance can result in lumber that is thinner on the [1, 114].
ends than in the middle. This leads to a type of bow, called Since the shrinkage in the direction of the growth rings is
“pile bend,” which appears in the upper layers of a pack. greater, so the checking is most likely to occur mainly on the
Proper and timely handling can help to reduce or eliminate flatsawn surface of lumber and on the edges of quarter-sawn
bow. Wood that is wet and warm can bend quite easily. The lumber. This type of check often occurs along the rays which
shrinkage that causes bow usually occurs at high moisture form planes of weakness [131]. Surface checks, however,
could be closes up when the stress reversal happens [48].
Surface checking is more likely during the first stage of
drying process, when the lumber is green and loses one-third
of its moisture. So, the moisture gradient is considerable in
the first steps of drying, though in some softwoods the danger
still exists through the drying. In general, the main cause of
surface checking is drying the lumber too fast in the first
steps, and the result is even worse when the H is also too low,
velocity is too high, and excessive temperatures are applied
[1, 3, 131, 132]. Accordingly, Keey et al. (2000) [3] suggest
that this can be minimized by maintaining high relative
humidity at the first steps of drying, which will later prevent
surface shrinkage and increase the plasticity of the wood to
Kink Twist Cup Crook Bow
accommodate the stresses.
Honeycomb is an internal check caused by a tensile failure
Fig. 13.27 Types of shape distortions (warping) due to drying across the grain of wood and usually occurs along the wood
13 Basics of Wood Drying 703
alleviate the problem in some cases. Drying in a reduced 18. Perré, P.: Holzforschung. 61(4), 419–429 (2007)
oxygen content environment might alleviate the problem. 19. Babiak, M., Kudela, J.: Wood Sci. Technol. 29(3), 217–226 (1995)
20. Schniewind, A.P.: Forest Prod. J. 13(11), 475–480 (1963)
A second common type discoloration is the sticker-stain. 21. McMillen, J.M.: Stresses in Wood during Drying, 174pp. Madison,
The contact area between wood and a sticker may result in U.S. Department of Agriculture, Forest Service, Forest Product
accumulation of chemicals from the wood and reaction with Laboratory, 1652 (1958)
sticker chemicals may result in a distinct line on the surface of 22. USDA. Wood handbook. Wood as an Engineering material
U.S. Department of Agriculture General Technical Report
wood which may be superficial or not. In some cases, this is a FPL-GTR-113. 486pp (1999)
superficial discoloration that is taken out in the planer, but in 23. Whale, L.R.J.: Deformations characteristics of nailed or bolted
others it runs deep into the thickness of the board and in this timber joints subjected to irregular short or medium term lateral
case it affect the grade of the product. The problem seems to loading. PhD dissertation, South Bank Polytechnic, London, 189p
(1988)
significantly reduce if wood-stickers are replaced by alumi- 24. Rice, R.W., Youngs, R.L.: Holz Roh-Werkst. 48, 73–79 (1990)
num, stainless steel or plastic ones at considerable increased 25. Salin, J.G.: Holz Roh-Werkst. 50, 195–200 (1992)
cost or their rectangular cross-section is replaced by an 26. Ranta-Maunus, A.: Mater. Struct. 26, 362–369 (1993)
i-beam or bone one thus reducing contact area and allowing 27. Wu, Q.: Rheological behavior of Douglas-fir as related to the
process of drying. PhD thesis, Oregon State University, 228p
water evaporation. Sticker staining is of a problem with (1993)
hardwoods and few softwoods that mostly end-up used in 28. Svensson, S.: Holzforschung. 51, 472–478 (1997)
products where are highly exposed to view (floors, furniture, 29. Ormarsson, S.: Numerical analysis of moisture-related distortions
cabinets, doors, windows, etc.). Most of softwoods are used in sawn timber. PhD thesis. Publication 99:7, Chalmers University
of Technology, Goteborg, Sweden, 214p (1999)
structurally where strength, not a List of symbols and 30. Haque, M.N., Langrish, T.A.G., Keep, L.B., Keey, R.B.: Wood Sci.
abbreviations [1]. Technol. 34(5), 447–457 (2000)
31. Dahlblom, O., Petersson, H., Omarsson, S.: Full 3-D
FEM-simulations of drying distortions in spruce boards based on
experimental studies. In: Proceedings of the 7th international
References IUFRO wood drying conference, Tsukuba, 246–251 (2001)
32. Pang, S.: Modeling of stresses and deformation of Radiata Pine
1. Simpson, W.T.: Dry Kin Operator’s Manual, p. 256pp. United lumber during drying. In: Proceedings of the 7th international
States Department of Agriculture, Forest Service Forest Products IUFRO wood drying conference. Tsukuba. pp. 238–245 (2001)
Laboratory (1991) 33. Moutee, M., Fortin, Y., Fafard, M.: Wood Sci. Technol. 41(3),
2. Simpson, W., TenWolde, A.: Physical Properties and Moisture 209–234 (2007)
Relations of Wood. USDA Forest Service, Forest Products Labo- 34. Eder, M., Arnould, O., Dunlop, J.W., Hornatowska, J., Salmén, L.:
ratory. General technical report FPL; GTR-113: pp. 3.1–3.24, Mad- Wood Sci. Technol. 47(1), 163–182 (2013)
ison (1999) 35. Clair, B., Alméras, T., Pilate, G., Jullien, D., Sugiyama, J., Riekel,
3. Keey, R.B., Langrish, T.A.G., Walker, J.C.F.: Kiln-Drying of Lum- C.: Plant Physiol. 152(3), 1650–1658 (2010)
ber. Springer (1999). https://doi.org/10.1007/978-3-642-59653-7 36. Peck, E.C.: Southern Lumberman. 161(2033), 136–137 (1940)
4. Siau, J.F.: Wood: Influence of Moisture on Wood Properties. 37. Kuebler, H.: Drying stresses and stress relief in thin sections of
Department of Wood Science and Forest Products, Virginia Poly- wood. Forest Products Laboratory Report No. 2164 (1960)
technic Institute and State University, Blacksburg (1995) 38. Espinoza, O., Bond, B.: Vacuum drying of wood state of the art.
5. Stamm, A.J.: Wood Sci. Technol. 1(2), 122–141 (1967a) Curr. For. Rep. 2, 223–235 (2016). https://doi.org/10.1007/s40725-
6. Stamm, A.J.: Wood Sci. Technol. 1(3), 205–230 (1967b) 016-0045-9
7. Skaar, C.: Wood-Water Relations. Springer-Verlag, 39. Mujumdar, A.S.: Handbook of Industrial Drying, 4th edn. CRC
Heidelberg/New York/Berlin (1988) Press (2014)
8. Perre, P.: Fundamental Wood Drying. European COST, Nancy 40. Mollier, R.: Zeitschrift des Vereines deutscher Ingenieure [Journal
(2007) of the Society of German Engineers] (in German). 67,
9. Perre, P., Martin, M.: Dry. Technol. 12, 1915–1941 (1994) 869–872 (1923)
10. Perre, P., Moser, M., Martin, M.: J. Heat Mass Transf. 36, 41. Kanagawa, Y., Yasujima, M.: Wood. 93, 292 (1993)
2725–2746 (1993) 42. Fohr, J.P., Chakir, A., Arnaud, G., du Peuty, M.A.: Dry. Technol.
11. Perre, P., Thiercelin, F., Aguiar, O.: Dry. Technol. 18, 13(8–9), 1675–1693 (1995)
1849–1863 (2000) 43. Defo, M., Cloutire, A., Fortin, Y.: Modelling vacuum Contact
12. Stamm, A.J.: Wood and Cellulose Science, p. 509p. Ronald Press, Drying of Wood: the Water Potential Approach in Drying of
New York (1964) Wood. Marcel Dekker Inc. (2000)
13. Stamm, A.J.: Ind. Eng. Chem. 1, 94–97 (1929) 44. Chen, Z., Lamb, F.M.: Wood Fiber Sci. 33(4), 639–647 (2001)
14. Stevens, W.C.: The shrinkage and expansion of wood. Forestry. 12, 45. Gerhards, C.C.: Wood Sci. Technol. 20(4), 349–360 (1986)
38–43 (1938) 46. Jung, H.S., Eom, C.D., So, B.J.: Dry. Technol. 22(5),
15. Abe, K., Yamamoto, H.: J. Wood Sci. 52, 15–19 (2006) 1005–1022 (2004)
16. Hernández, R.E., Bizon, M.: Wood Fiber Sci. 26(3), 47. Koumoutsakos, A., Avramidis, S., Hatzikiriakos, S.G.: Dry.
360–369 (1994) Technol. 19(1), 65–84 (2001)
17. Almeida, G., Assor, C., Perré, P.: The dynamic of shrinkage/ 48. Liu, F., Avramidis, S., Zwick, R.L.: For. Prod. J. 44(6),
moisture content behavior determined during the drying of 71–75 (1994)
micro-samples. In: Proceedings of the 10th IUFRO Interna- 49. Zwick, R.L., Avramidis, S.: Commercial RFV kiln drying: recent
tional Wood Drying Conference, Orono, 26–30 August, successes. In: 51st Western Dry Kiln Association Meeting. Western
2007: 16–22 Dry Kiln Association, Reno, NV (2000)
13 Basics of Wood Drying 705
50. Resch, H.: Maderas-Cienc. Technol. 8(2), 67–82 (2006) 89. Shubin, G.S.: Drying and heat treatment of wood. Lesnaia Pro-
51. Leiker, M., Adamska, M.A.: Eur. J. Wood Wood Prod. 62(3), myshlennost Moskva, 154pp (1990)
203–208 (2004) 90. Van Meel, D.A.: Chem. Eng. Sci. 9, 36–44 (1958)
52. Avramidis, S., Liu, F.: Dry. Technol. 12(8), 1963–1981 (1994) 91. Ashworth, J.C.: The mathematical of batch-drying of softwood
53. Koumoutsakos, A., Avramidis, S., Hatzikiriakos, S.G.: Dry. timber. PhD thesis. University of Canterbury. 2 vols (1977)
Technol. 19(1), 58–98 (2001) 92. Nijdam, J.J.: Reducing moisture-content variation in kiln-dried
54. Koumoutsakos, A., Avramidis, S., Hatzikiriakos, S.G.: Dry. timber. PhD thesis, University of Canterbury, Christchurch (1998)
Technol. 21(8), 1399–1410 (2003) 93. Keey, R.B., Walker, J.C.F.: The drying of impermeable timbers.
55. Koumoutsakos, A., Avramidis, S., Hatzikiriakos, S.G.: Maderas: New Zealand hard beech Proc CHEMECA’88, Sydney NSW 13
Ciencia y Tecnología. 4(1), 15–25 (2002) 2: 421–435 (1988)
56. Elustondo, D., Avramidis, S.: Wood Fiber Sci. 35(1), 49–55 (2003) 94. Kayihan, F.: Comput. Chem. Eng. 17, 265–273 (1993)
57. Elustondo, D., Avramidis, S., Oliveira, L.: Maderas: Ciencia y 95. Keey, R.B., Pang, S.: Chem. Eng. Res. Design. 72(6),
Technologia. 7(2), 65–78 (2005) 741–753 (1994)
58. Elustondo, D., Avramidis, S., Zwick, R.: For. Prod. J. 55(1), 96. Kamke, F.A., Vanek, M.: Comparison of wood drying models.
76–83 (2005) Proc 4th IUFRO Int Wood Drying Conf Rotorua, 1–21 (1994)
59. Avramidis, S., Zwick, R.L.: For. Prod. J. 42(7/8), 17–24 (1992) 97. Kayihan, F.: Simulation of heat and mass transfer with local three-
60. Lee, N.-H., Jung, H.-S.: For. Prod. J. 50(2), 69 (2000) phase equilibria in wood drying. Proc. 3rd Int. Drying Symp.
61. Leiker, M., Adamska, M.A., Gutter, R., Mollekopf, N.: Vacuum Birmingham. 1, 123–134 (1982)
micro-wave drying of beech: property profiles and energy effi- 98. Plumb, O.A., Spolek, G.A., Olmstead, B.A.: Int. J. Heat Mass
ciency. In: International Conference of COST Action E15 Wood Transf. 28, 1669–1678 (1985)
Drying, p. 10. European Co-operation in the Field of Scientific and 99. Stanish, M.A., Schajer, G.S., Kayihan, F.: AIChE J. 32,
Technical Research, Athens (2004) 1301–1311 (1986)
62. Simpson, W.T.: For. Prod. J. 37(1), 35–38 (1987) 100. Perre, P.: The convective drying of resinous woods selection vali-
63. Elustondo, D., Avramidis, S.: Wood Fiber Sci. 35(1), 49–55 (2007) dation and use of a model. These de Doctorate, I’Universite, Paris
64. Elustondo, D., Avramidis, S., Keery, R.H., H.: J. Inst. Wood Sci. (1987)
18(2), 68–74 (2008) 101. Turner, I.W.: The modelling of combined microwave and convec-
65. Koumoutsakos, A., Avramidis, S., Hatzikiriakos, S.G.: Maderas: tive drying of a wet porous material. PhD thesis, University of
Ciencia y Tecnologia. 4(1), 15–25 (2002) Queensland, St Lucia, 1990
66. Resch, H.: Maderas: Ciencia y Tecnología. 8(2), 67–82 (2006) 102. Chen, G., Keey, R.B., Walker, J.C.F.: Stress development and
67. Elustondo, D., Avramidis, S.: Dry. Technol. 20(9), permeability variation on drying sapwood above fiber saturation.
1827–1842 (2002) In: Proc 5th IUFRO International Wood Drying Conference, Que-
68. Elustondo, D., Avramidis, S., Shida, S.: Dry. Technol. 22(4), bec City, 455–462 (1996)
795–807 (2004) 103. McCurdy, M., Keey, R.B.: Determination of moisture saturation
69. Chen, Z., Lamb, F.M.: For. Prod. J. 51(10), 55 (2001) properties on high-temperature drying. Proc IPENZConf, Welling-
70. Chen, Z., Lamb, F.M.: Dry. Technol. 22(3), 577–595 (2004) ton (1998)
71. Kudra, T., Mujumdar, A.S.: Advanced Drying Technologies, 104. Horie, K., Nakamura, N., Iijima, Y.: Mokuzari Gakkaishi. 45,
p. 459. Marcel Dekker, New York (2002) 103–110 (1999a)
72. Neumann, R., Mielke, A., Gios, P.: Eur. J. Wood Wood Prod. 51(3), 105. Horie, K., Nakamura, N., Iijima, Y.: Mokuzari Gakkaishi. 45,
156–162 (1993) 455–460 (1999b)
73. Brook, R.C., Bakker-Arkema, F.W.: Trans. ASAL. 21(5), 106. Elustondo, D.M., Avramidis, S.: Comparative analysis of three
978–981 (1978) methods for stochastic lumber drying simulation. Dry. Technol.
74. Benin, R., Blazquez, M.: Dry. Technol. 4(1), 45–66 (1986) 23, 131–142 (2005). https://doi.org/10.1081/DRT-200047663
75. Vagenas, G.K., Marinson-Kouris, D.: Dry. Technol. 9(2), 107. Watanabe, K., Hayashi, T., Kobayashi, I.: J. Wood Sci. 62,
439–461 (1991) 479–486 (2016). https://doi.org/10.1007/s10086-016-1587-y
76. Parry, J.L.: J. Agri. Eng.Res. 32, 1–29 (1985) 108. Elustondo, D.M., Avramidis, S.: Dry. Technol. 20,
77. Mulet, A., Berna, S., Rossello, C.: Dry. Technol. 7, 537 (1989) 1827–1842 (2002)
78. Mulet, A., Berna, S., Rossello, C., Pinaga, F.: Dry. Technol. 7, 109. Elustondo, D.M., Avramidis, S.: J. Wood Sci. 49, 485–491 (2003)
641 (1989) 110. Cronin, K.P., Norton, B., Taylor, J.: Dry. Technol. 15(3–4),
79. Tong, C.H., Lund, D.B.: Biotechnol. Prog. 6, 67–75 (1990) 765–790 (1997)
80. Gekas, V., Lamberg, I.: J. Food Eng. 14, 317–326 (1991) 111. Cronin, K.P., Abodayeh, K., Caro-Corrales, J.: Dry. Technol. 20(2),
81. Zhang, D.Y., Sun, L.P., Cao, J.: J. For. Res. 17(2), 141–144 (2006) 307–324 (2002)
82. Čiegis, R., Starikovičius, V.: Math. Model. Anal. 7(2), 112. Cronin, K.P., Baucour, P., Abodayeh, K., Barbot Da Silva, A.: Dry.
177–190 (2002) Technol. 21(8), 1435–1458 (2003)
83. Pang, S.: Dry. Technol. 25(3), 421–431 (2007) 113. Wu, H., Avramidis, S.: Dry. Technol. 24, 1541–1545 (2006).
84. Bramhall, G.: Wood Sci. 8(4), 213–222 (1976) https://doi.org/10.1080/0737393060147584
85. Tschernitz, J.L., Simpson, W.T.: Wood Sci. 114, 202–208 (1979) 114. Bramhall, G., Wellwood, R.W.: Kiln Drying of Western Canadian
86. Helmer, W.A., Rosen, H.N., Chen, P.Y.S., Wang, W.S.: A theoret- Lumber. Vancouver: D. F. E., Information report/ Forest Products
ical model for solar-dehumidification drying of wood. In: Laboratory (no. VP-X-159), 112 pp (1976)
Mujumdar, A.S. (ed.) Drying’80, pp. 21–28. Hemispher, 115. Wengert, E.M., Lamb, F.M.: End coating of lumber to prevent end
Washington, DC (1980) checking. In: Proceeding 2nd IUFRO International Wood Drying
87. Milota, M.R., Tschernitz, J.L.: Wood Fiber. 22, 298–313 (1990) Conference, Seattle Washington, 164–168 (1994)
88. Olek, W., Guzenda, R., Dudzinski, J.: Influence of equilibrium 116. Ward, J.C., Simpson, W.T.: 1991. Dry Kiln Operator’s Manual.
moisture content and temperature on drying process of beech Chapter 8, Drying Defects. Forest Products Society, 274 p
Fagus sylvatica L. wood. Proc 4th IUFRO Int Wood Drying Conf 117. Esping, B.: Energy Saving in Wood Drying. Wood Technology
Rotorua NZ, 102–106 1994 Report No. 12. Svenska Traforsknings Institute, Stockhole (1982)
706 S. Avramidis et al.
118. Aune, J.E.: Kiln tests with species and moisture content sorted,
116mm square, hemfir merch timber. Final report prepared for the
Stability Work Group, ZAIRAI Timber Partnership Ltd., Vancou-
ver (2000)
119. Avramidis, S.: Evaluation of conventional and radio frequency
vacuum drying and re-drying of Pacific Coast hemlock Hashira
and Hirakaku timbers. Final report prepared for the Stability Work
Group, ZAIRAI Timber Partnership Ltd. and Coast Forest Timber
Association, Vancouver (2001)
120. Sandberg, D.: Holz als Roh-und Werkstoff. 63(1), 11–18 (2005)
121. Bendtsen, B.A.: For. Prod. J. 28(20), 61–72 (1978)
Dr. Ciprian Lazarescu is a Research Scientist in Smart Manufactur-
122. Larson, P.R., Kretschmann, D.E., Clark, A.I.I.I., Isebrands, J.G.:
ing, FPInnovations located in Vancouver. He joined the Drying and
Formation and Properties of Juvenile Wood in Southern Pines: a
Energy research group in 2014 and his research area is focusing on
Synopsis. General Technical Reports FPL-GTR-129, p. 42pp.
identifying new ways to reduce drying costs and degrade related to the
U.S. Department of Agriculture, Forest Service, Forest Products
drying process. Ciprian completed his undergraduate and master’s
Laboratory, Madison (2001)
degree in wood science in Romania at Transylvania University of
123. Burdon, R.D., Kibblewhite, R.P., Walker, J.C.F., Megraw, R.A.,
Brasov. A few years after graduation he came to UBC and conducted
Evans, R., Cown, D.J.: For. Sci. 50(4), 399–415 (2004)
research in wood drying particularly on hygro-stresses generated by the
124. Woxblom, L.: Quality variations in wall studs- A study conducted
drying process as part of a PhD program completed in 2009. After
at five sawmills in southern Sweden. (in Swedish) Swedish Uni-
graduation, he worked for UBC as a Postdoctoral Fellow and Research
versity of Agricultural Sciences, Department of Forest-Industry-
Associate on conventional and radio-frequency vacuum drying and
Market-Studies, Report No 28, Uppsala, Sweden (1993)
heating, near infraRed reflectance spectroscopy (NIR) of wood surface,
125. Danborg, F.: Scand. J. For. Res. 9, 91–98 (1994)
CT scanning of wood, and phytosanitary treatments for fungi, nema-
126. Perstorper, M., Pellicane, P.J., Kliger, I.R., Johansson, G.: Wood
todes, and insects.
Sci. Technol. 29, 157–170 (1995)
127. Balodis, V.: Wood Sci. 5, 44–50 (1972)
128. Forsberg, D., Warensjo, M.: Karst. Scandinavian J. For. Res. 16(3),
269–277 (2001)
129. Shelly, J.R., Arganbright, D.G., Birnbach, M.: Wood Fiber Sci.
11(1), 50–56 (1979)
130. Beard, J.S., Wagner, F.G., Taylor, F.W., Seale, R.D.: For. Prod.
J. 43(6), 51–56 (1993)
131. Pratt, G.H.: Timber Drying Manual, p. 152pp. McCorquodale
Printers Ltd, London (1974)
132. Denig, J., Wengert, E.M., Simpson, W.T.: Drying hardwood lum-
ber. Gen. Tech. Rep. FPL-GTR-118. Madison, WI: Dr. Sohrab Rahimi is a PhD student of wood drying at University of
U.S. Department of Agriculture, Forest Service, Forest Products British Columbia, Vancouver, Canada. He entered the Department of
Laboratory, 138pp (2000) Wood Science in 2017 and his research area is concentrating on model-
133. Fuller, J.: Dry. Technol. 18(1&2), 383–393 (2000) ing final moisture content of kiln-dried timbers. Sohrab has two master’s
134. Fuller, J.: Dry. Technol. 18(4), 1073–1080 (2000) degrees in forestry from University of Tehran, Iran, and West Virginia
University, USA.