You are on page 1of 385

EUROPEAN COST A.R.BO.LOR.

FUNDAMENTALS OF WOOD
DRYING

edited by
PATRICK PERRÉ
Fundamentals of wood drying
Edited by Patrick Perré
ISBN

Copyright © 2007 by A.R.BO.LOR.


All rights reserved
No part of this publication may be reproduced or transmitted in any form or
by any means, electronic or mechanical, including photocopy, recording, or
any information storage and retrieval system, without permission in writing
from the publisher.

A.R.BO.LOR
ENGREF,
14, rue Girardet
F-54 042 Nancy, France

Printed in France

Cover

Foreground :
Liquid water menisci in spruce tracheids, ESEM microphotograph, P. Perré.
Simulated moisture content fields using a heterogeneous drying model, P. Perré and
I.W. Turner (cf. p. 233).
Checks in a spruce disk after drying due to the shrinkage contrast between normal
wood and compression wood, P. Perré.
Background :
Oak (Quercus pedunculata), ESEM microphotograph, P. Perré.
COST A.R.BO.LOR.

FUNDAMENTALS OF WOOD
DRYING

edited by

PATRICK PERRÉ
To my wife
To my children
Preface

Wood is an integral component of all standing trees, playing many important roles
such as acting as a mechanical support and vascular system; and providing
structural durability. This is possible as a result of the complex structure of organic
macromolecules, which are well organised at different spatial scales, from the
nanometer to the meter (molecules, cell wall, cellular structure, anatomical pattern,
annual rings, logs). As a material, wood is used because of its unique combination
of properties, which include low density, high longitudinal rigidity, high impact
strength, dimensional stability in the longitudinal direction, low thermal
conductivity, low thermal effusivity (pleasant to touch whatever the wood
temperature) and finally, its aesthetical and textural aspects.

Nowadays, in the context of sustainable development, wood is seen as a renewable


resource, which is able to prolong the storage of carbon dioxide fixed during the
growth of forests (use as a material) and to avoid the emission of carbon dioxide
from fossil energy (use as an energy).

However, this material of biological origin also has some drawbacks: it is a


hygroscopic material subject to swelling/shrinkage and is prone to biological
attacks, especially when it contains water with high activity. Therefore, drying is an
operation that is absolutely essential in the chain of wood processing. This
operation takes place after the first cutting operation (for dimensional reasons,
drying of logs would be difficult) and, due to shrinkage, before any structural
assembly or finishing operations.
ii

At this point we know why wood has to be dried, but not how to dry it effectively. In
fact, although wood has been used by man since pre-history, what is today called
“conventional drying” started only around one century ago. Prior to this, wood was
just dried using “natural drying”. One therefore notes that the present industrial
drying knowledge benefits from around one century of empirical feedback and
roughly half a century of scientific research in the field of drying. However, wood
drying is a complex industrial operation that involves coupled heat and mass
transfer, shrinkage induced stress and chemical reactions that might take place
during the process. Each single point of this list can be the subject of experimental
and theoretical investigation. Just the problem of mechanical behaviour of wood
subject to load, moisture and temperature history is of dramatic complexity,
involving anisotropy, elasticity, viscoelasticity, mechanosorption, thermo and hygro
activation, plasticity, cell collapse, thermal degradation, wood variability…

This complexity is probably the main reason why wood drying remains, at the
industrial level, mainly based on empirical knowledge. Just by looking only at what
happens in industry, one quickly concludes that the situation evolves very slowly.
Computers now control kilns and the continuous evolution of drying parameters
replaced classical schedules with a discrete set of drying conditions. On the other
hand, the principle of wood drying, the optimisation of drying conditions and the
nature of information collected during drying are basically the same as they were 30
to 40 years ago. In the Principles of wood science and technology (Kollman and
Côté 1968), we see that most techniques used nowadays are already mentioned,
including heating by Joule’s effect or by high-frequency, high temperature drying…
Concerning the mechanisms, several subtle phenomena are clearly indicated such
as the effect of steaming before drying, the change of colour with temperature level,
or the required limited temperature level during the liquid domain to avoid collapse.

Obviously, numerous improvements to dryers and dryer controls have been


implemented during the last four decades (use of computers to control kilns, a wider
range of temperature levels, improved sticker design to avoid stain, improved
quality of the dried products, development of vacuum drying - including high vac -,
better control of dryers - including the case of airflow reversal -, improvement of
kiln insulation, new kiln instrumentation and so on…). Clearly, most of these
improvements have come directly from industrial experience and very few
originated from scientific research work. However, the amount of scientific
knowledge in the field of wood drying is nowadays impressive and all is ready for a
rapid input of science in the industrial field in the future, especially through
modelling and simulation, which provide powerful tools able to synthesise the
present knowledge and which are capable of prediction for management decision
making processes.

Filling the gap between industrial practices and scientific knowledge was one of the
goals of the COST Action E15 (Advances in the drying of wood), namely of Working
Group 2 (WG2: Fundamentals aspects of wood drying) of this action, which ran
iii

from 1999 to 2004. COST is an intergovernmental network that is scientifically


completely self-sufficient with nine scientific COST Domain Committees formed by
some of the most outstanding scientists of the European scientific community. Scien-
tists can propose a Memorandum Of Understanding for a new action in a specified
field. If accepted by the COST committee, participating countries are invited to sign
a new action. Basically, a COST action allows travelling costs to be paid for meet-
ings, workshops and short term scientific missions. Our COST action E15 has been
accepted in the domain FPS (Forests, their Products and Services). It allowed
European scientists in the field of wood drying to meet on average twice a year dur-
ing 5 years.

The present book results from the work done during this period of time in WG2
(Fundamentals aspects of wood drying). It should be considered as a state of the art
of fundamental knowledge on phenomena involved in wood drying. Its major aim is
to provide sufficient knowledge to scientists and engineers involved in this field in
order to analyse and improve existing situations. The book also may be a first step
in the development of innovative procedures that hopefully will enable drying
practitioners to imagine, simulate, improve, and implement any new idea...

For the editor, writing this preface to the book means the end of a long process.
When I proposed first, as the chairman of WG2, to prepare a publication devoted to
the state-of-the-art of fundamental knowledge in wood drying, many scientists in the
audience told me that, without any money except travelling expenses, the project
was just utopia. However, we advanced stepwise, sometimes too slowly, but we all
kept going forward. A major step in the project, at least in motivating people to
undertake the project, has been the analysis of the questionnaire I proposed to WG2
members at the end of 2001. This questionnaire not only allowed the “ideal” content
of the book to be defined, the answers were more than encouraging. In fact out of 37
answers, I received 23 excellent or good responses regarding the relevance of such
a project and 26 answers excellent or good regarding the capacity of our group to
write this book.

From this point we found enough volunteers as chapter authors or co-authors, so


the present content is very close to the ideal list as defined by our questionnaire: one
useful chapter has been added (psychometrics) and only one, although important,
topic is missing, due to the lack of a suitable author (energy consumption in wood
drying).

Then, we allocated time for writing the text and we just had time, before the end of
COST Action E15, to organise a collective reviewing of most chapters. From this
time, I had almost all of the material and a long lonely work was in front of me to
read and edit all the material in such a way to ensure that the layout, the list of
symbols, the section, figure and equation numberings, the reference list were
consistent. Doing this work, I corrected some errors here and there, but the content
of each chapter basically remains under the responsibility of the authors. During
this work, I had serious health problems, that changed my life priorities for a while
iv

and that delayed the printing of this book considerably. I’m aware of this and I
would like to apologise here to all authors.

Before allowing the reader shift to science, I would like to thank all people that
allowed this project to be achieved:
- the COST Action which allowed us to meet regularly,
- Helmut Resch, member of the FPS Committee and Alpo Ranta-Manus,
Chairman of Action E15 who encouraged me in this project,
- all members of WG2 for their participation,
- all authors of the book chapters,
- ARBOLOR, the association for wood science in Lorraine (France), which
paid the printing fees,
- all colleagues of my team at ENGREF who supported me.

In addition, I would like to thank more specifically my colleague and friend Ian
Turner, from QUT in Australia, who helped me a lot in correcting my English and
Wiesław Olek from Agricultural University of Pozna for his invaluable help in
reviewing the entire book.

I dedicate this book to all my family, whose support and comfort was determinant.

Nancy, July 2007.

Prof. Patrick Perré


Chairman of WG2: Fundamentals aspects of wood drying (COST E15)
Vice-president of ARBOLOR
Table of contents

Preface .............................................................................................................i
Table of contents ...........................................................................................v
Contributors .................................................................................................xi
1 From fundamental to practice: the interaction chain...........................1
1.1 The need for timber drying .......................................................................... 2
1.2 The raw product: density, anisotropy and variability................................... 3
1.3 Wood drying: a complex system of interfering processes ........................... 5
1.4 Physics and mechanics of timber drying ..................................................... 7
1.5 Measures to obtain a fast and good drying process ................................... 10
1.5.1 Drying time ................................................................................... 10
1.5.2 Mechanical quality ........................................................................ 11
1.5.3 Recipes for obtaining both a fast and good process ...................... 12
1.6 Innovations in the industrial drying process .............................................. 14
1.7 About this book ......................................................................................... 16
1.7.1 Wood and air ................................................................................. 16
1.7.2 Equilibrium states.......................................................................... 17
1.7.3 Single phenomena ......................................................................... 17
1.7.4 Coupled phenomena ...................................................................... 18
1.7.5 Miscellaneous................................................................................ 18
1.8 Conclusion ................................................................................................. 19
2 Wood anatomy - an introduction..........................................................21
2.1 Introduction ............................................................................................... 22
2.2 Chemical composition ............................................................................... 22
2.3 Cell wall structure...................................................................................... 27
2.4 Cell wall sculpturing.................................................................................. 30
2.4.1 Pits................................................................................................. 30
2.4.2 Spiral thickenings .......................................................................... 33
2.4.3 Warty layer.................................................................................... 33
2.5 Microscopic structure of wood .................................................................. 34
2.5.1 Softwood structure ........................................................................ 34
2.5.2 Hardwood structure ....................................................................... 37
2.5.3 Reaction wood............................................................................... 43
2.5.4 Juvenile wood................................................................................ 45
2.6 Macroscopic structure of wood.................................................................. 46
3 Psychrometrics .......................................................................................51
3.1 Introduction ............................................................................................... 52
3.2 Moist air..................................................................................................... 52
vi Fundamentals of wood drying

3.2.1 Humidity ....................................................................................... 53


a) Relative humidity (H).................................................................... 53
b) Absolute humidity (Y)................................................................... 54
c) Moisture concentration (c) ............................................................ 54
d) Classification of moist air states.................................................... 55
3.2.2 Enthalpy (h)................................................................................... 55
3.2.3 Moist air density ( ) ...................................................................... 56
3.2.4 Empirical formulas for calculations of moist air properties .......... 57
3.3 Psychrometric chart (Mollier h-Y diagram for moist air) .......................... 58
3.3.1 Basic processes of moist air and their interpretation in the
psychrometric chart ....................................................................... 60
a) Heating – process “0” “1” ........................................................ 60
b) Cooling – process “0” “2” ........................................................ 60
c) Cooling with dehumidification – process “0” “3” .................... 60
d) Humidifying with steam – process “0” “4” .............................. 60
e) Humidifying with water – process “0” “5” .............................. 60
f) Drying – process “0” “6”.......................................................... 61
g) Dehumidification – process “0” “7”......................................... 61
h) Mixing moist air with moist air ..................................................... 61
3.3.2 Measurements of moist air relative humidity ................................ 63
a) Hygrometric methods .................................................................... 63
b) Dew point method ......................................................................... 63
c) Psychrometric method................................................................... 63
4 Sorption isotherms of wood...................................................................67
4.1 Introduction ............................................................................................... 68
4.2 Definitions ................................................................................................. 68
4.3 How to measure Xe?................................................................................... 70
4.4 Factors affecting wood equilibrium moisture content ............................... 72
4.4.1 Influence of species ....................................................................... 72
4.4.2 Influence of sorption hysteresis..................................................... 74
4.4.3 Influence of temperature ............................................................... 75
4.5 Sorption models ......................................................................................... 76
4.6 Thermodynamics of sorption ..................................................................... 83
4.7 Testing of sorption theories ....................................................................... 84
4.8 Conclusions ............................................................................................... 85
5 Shrinkage, swelling and warp caused by moisture changes...............87
5.1 Introduction ............................................................................................... 88
5.2 Definitions ................................................................................................. 88
5.3 How to measure shrinkage/swelling and warp?......................................... 89
5.4 Shrinkage and swelling.............................................................................. 90
5.4.1 Wood – water relationships ........................................................... 90
5.4.2 The anisotropic character of shrinkage and swelling in wood....... 92
5.4.3 Shrinkage/swelling for various wood species ............................... 94
5.4.4 The influence of various wood properties ..................................... 95
5.4.5 Shrinkage and deformations during drying of full size timber ...... 96
a) Shrinkage in full size timber ......................................................... 97
Table of contents vii

b) Twist, bow, spring and cupping .................................................... 98


c) Economic aspects of shrinkage and warp.................................... 102
5.4.6 Dimensional stabilisation by modification .................................. 103
6 Bound water migration in wood .........................................................105
6.1 Introduction ............................................................................................. 106
6.2 Observation/Formulation ......................................................................... 106
6.3 How to measure diffusion coefficients .................................................... 107
6.4 Isothermal water diffusion in wood ......................................................... 110
6.4.1 Moisture movement in wood....................................................... 110
6.4.2 Isothermal diffusion equations .................................................... 111
6.4.3 The diffusion coefficient ............................................................. 114
6.4.4 Non-Fickian diffusion ................................................................. 119
6.5 Nonisothermal water diffusion in wood .................................................. 121
6.5.1 General considerations ................................................................ 121
6.5.2 Theoretical equations .................................................................. 122
7 Fluid migration in wood ......................................................................125
7.1 Introduction ............................................................................................. 126
7.2 Darcy’s law: single-phase flow................................................................ 126
7.2.1 Observation, formulation............................................................. 126
7.2.2 How to measure permeability...................................................... 128
7.3 Specific features of wood regarding permeability ................................... 130
7.3.1 Orders of magnitude: variability and anisotropy of wood........... 130
7.3.2 Simple pore models applicable to wood: from fluid viscosity to
Darcy’s law ................................................................................. 133
7.3.3 Pit aspiration................................................................................ 134
7.4 Deviation from the linear law .................................................................. 135
7.4.1 Laminar flow, turbulent flow ...................................................... 135
7.4.2 Slip flow ...................................................................................... 137
7.4.3 Effect of sample length on permeability...................................... 139
7.4.4 Percolation models ...................................................................... 140
7.4.5 Multiscale effects ........................................................................ 141
7.5 Capillary pressure .................................................................................... 144
7.5.1 Surface tension: observation, formulation................................... 144
7.5.2 The case of three-phase system ................................................... 145
7.5.3 Vapour pressure at the surface of a meniscus.............................. 147
7.5.4 The capillary pressure in porous media....................................... 148
7.6 Generalised Darcy’s law: multiphase flow .............................................. 152
7.6.1 Observation, formulation............................................................. 152
7.6.2 Migration mechanisms during drying ......................................... 154
7.7 Conclusion ............................................................................................... 156
8 Creep deformation in drying wood ....................................................157
8.1 Introduction ............................................................................................. 158
8.2 Importance of creep for drying crack formation ...................................... 158
8.3 Definition of creep and creep mechanisms .............................................. 160
8.3.1 Different mechanisms of creep.................................................... 162
8.4 Review of experiments ............................................................................ 163
viii Fundamentals of wood drying

8.4.1 Experimental techniques for measuring creep in drying wood ... 163
a) Measurements in constant conditions.......................................... 163
b) Measurements in changing state.................................................. 164
8.4.2 Viscoelastic creep........................................................................ 164
8.4.3 Mechanosorptive creep................................................................ 164
8.4.4 Relative magnitude of the creep mechanisms ............................. 168
8.5 Multidimensional effects ......................................................................... 168
8.6 Modelling of creep................................................................................... 168
8.6.1 Viscoelastic creep........................................................................ 168
8.6.2 Mechanosorptive creep................................................................ 169
a) General ........................................................................................ 169
b) First generation (Maxwell type) models...................................... 170
c) Dependence of hygroexpansion on strain.................................... 171
d) Kelvin type models...................................................................... 172
e) Generalized Kelvin and Maxwell type models............................ 172
f) Models of combined activation ................................................... 173
9 External heat and mass transfer.........................................................175
9.1 Introduction ............................................................................................. 176
9.2 Basic formulation..................................................................................... 176
9.3 How to measure external drying resistances............................................ 177
9.4 Heat transfer coefficient .......................................................................... 178
9.4.1 Centre part of normal kiln stack .................................................. 178
a) Fully developed flow between parallel plates ............................. 179
b) Entry region................................................................................. 180
c) Plate(s) in parallel flow ............................................................... 180
d) Influence of gap between boards and other irregularities............ 183
9.4.2 End part of kiln stack................................................................... 185
a) Single rectangular cylinder in cross flow .................................... 186
b) Multiple rectangular cylinders in cross flow ............................... 186
9.5 Mass transfer coefficient.......................................................................... 188
9.5.1 Analogy between heat and mass transfer .................................... 188
9.5.2 Apparent deviations from the analogy ........................................ 188
9.5.3 External mass transfer above FSP ............................................... 189
9.5.4 External mass transfer below FSP ............................................... 194
9.6 The surface emission concept .................................................................. 196
The surface emission approach to the external mass transfer is based on the
following single equation. ....................................................................... 196
9.7 The influence of the drying method......................................................... 196
9.8 Some pitfalls regarding external heat and mass transfer.......................... 198
10 Coupled heat and mass transfer .........................................................203
10.1 Introduction ............................................................................................. 204
10.2 Description of heat and mass transfer during drying ............................... 205
10.2.1 Low temperature convective drying ............................................ 205
a) The constant drying rate period................................................... 205
b) Effect of external drying conditions ............................................ 206
c) The decreasing drying rate period ............................................... 208
Table of contents ix

d) The effect of internal pressure on mass transfer .......................... 211


10.2.2 Drying with internal vaporisation................................................ 212
a) Convective drying at high temperature ....................................... 213
b) Vacuum drying............................................................................ 215
c) Microwave heating at atmospheric pressure ............................... 216
10.3 Physical formulation................................................................................ 218
a) The Comprehensive three equation model – (Model 1) .............. 219
b) The two equation model for heat and mass transfer – (Model 2) 220
c) The one equation model for mass transfer only – (Model 3)....... 221
10.4 Case studies in wood drying .................................................................... 230
10.4.1 Case Study 1 – The heterogeneous model................................... 230
10.4.2 Case Study 2 – Dielectric Drying ................................................ 236
10.4.3 Case Study 3 – Vacuum drying with radiative heating ............... 238
10.5 Conclusion ............................................................................................... 240
11 Stress development...............................................................................243
11.1 Introduction ............................................................................................. 244
11.2 A brief overview of mechanical aspects .................................................. 244
11.3 Stress development in a board during drying........................................... 246
11.4 Experimental evidence of drying stress ................................................... 248
11.5 Observation on micro-samples ................................................................ 250
11.5.1 Shrinkage: the driving force for stress......................................... 250
11.5.2 Checking: the end result of stress................................................ 251
11.6 The “Flying-wood” test........................................................................... 254
11.7 Drying stress formulation ........................................................................ 257
11.8 Some simulation examples ...................................................................... 262
11.9 Conclusion ............................................................................................... 271
11.10Acknowledgements.................................................................................. 271
12 Discoloration of wood during drying..................................................273
12.1 Introduction ............................................................................................. 274
12.2 Discolorations in freshly felled and stored round wood before kiln-drying274
12.3 Discolorations during kiln-drying............................................................ 276
12.3.1 Mechanisms of colour reactions during kiln-drying of hardwoods276
12.3.2 Mechanisms of colour reactions during kiln-drying of softwoods278
12.4 Discolorations during pre-steaming and steaming................................... 280
12.5 Remedies ................................................................................................. 282
12.5.1 Before sawing.............................................................................. 282
12.5.2 After sawing ................................................................................ 283
a) Fungal discoloration .................................................................... 283
b) Bacterial discoloration................................................................. 284
c) Chemical discoloration................................................................ 284
d) Sticker staining............................................................................ 286
e) Mineral discoloration .................................................................. 287
f) Alkaline discoloration ................................................................. 288
12.6 Drying schedules for light-coloured woods ............................................. 288
12.7 During storage (after drying) ................................................................... 289
12.8 Conclusions ............................................................................................. 289
x Fundamentals of wood drying

13 Airflow within kilns .............................................................................291


13.1 Introduction ............................................................................................. 292
13.2 Observed airflow phenomena in a kiln .................................................... 294
13.2.1 Kiln-wide flow phenomena ......................................................... 294
13.2.2 Stack flow phenomena ................................................................ 295
13.3 Measurement of flow field....................................................................... 297
13.3.1 Typical Measuring Equipment .................................................... 297
13.3.2 Selected measuring techniques for airflow in kilns..................... 298
a) Pressure probes............................................................................ 298
b) Hot-wire anemometry.................................................................. 299
c) Laser-Doppler anemometry......................................................... 300
d) Flow visualisation ....................................................................... 301
13.4 Airflow simulation................................................................................... 303
13.4.1 Conservation equations ............................................................... 303
13.4.2 Turbulent flows ........................................................................... 304
13.4.3 Notes on use of CFD ................................................................... 305
13.4.4 Examples of airflow simulation................................................... 306
a) Flow in an overhead fan kiln ....................................................... 306
b) Flow in an end-mounted fan kiln................................................. 310
13.5 Airflow analysis....................................................................................... 311
13.5.1 Airflow in the stack ..................................................................... 311
a) Flow between the board rows...................................................... 311
b) Stack pressure drop ..................................................................... 316
13.5.2 Airflow in the kiln ....................................................................... 318
a) Branch loss ratio of manifolds..................................................... 318
b) Airflow uniformity along the wood stack height......................... 320
13.5.3 Airflow design using engineering rules....................................... 323
a) Airflow and kiln design............................................................... 324
b) Improving airflow in existing kilns ............................................. 325
13.6 Influence of airflow on wood moisture distribution ................................ 326
13.6.1 Airflow induced drying non-uniformities.................................... 326
13.6.2 Temperature drop across the stack during the first drying period 327
a) One-dimensional heat transfer..................................................... 327
b) Drying rate uniformity across the width of the stack .................. 329
c) Concluding remarks .................................................................... 331
13.7 Acknowledgements.................................................................................. 332
14 References.............................................................................................333
15 List of symbols......................................................................................363
15.1 Latin letters .............................................................................................. 363
15.2 Greek Symbols ........................................................................................ 365
15.3 Superscripts and Subscripts ..................................................................... 365
15.4 Mathematical operations.......................................................................... 366
Contributors
Stavros Avramidis
Department of Wood Science
The University of British Columbia
#4028-2424 Main Mall, Vancouver, BC, V6T 1Z4, Canada
stavros.avramidis@ubc.ca

Marián Babiak
Faculty of Wood Science and Technology
Technical University in Zvolen
T.G.Masaryka 24, 960 53 Zvolen, Slovak republic
babiak@vsld.tuzvo.sk

Peder Gjerdrum
Norwegian Forest Research Institute
Høgskoleveien 12, N-1432 Ås, Norway
peder.gjerdrum@skogoglandskap.no

Ryszard Guzenda
Faculty of Wood Technology
Agricultural University of Pozna
ul. Wojska Polskiego 38/42, 60-627 Pozna , Poland
rguzenda@au.poznan.pl

Antti Hanhijärvi
VTT Building and Transport,
P.O.Box 1806, FIN-02044 VTT, Finland
Antti.Hanhijarvi@vtt.fi

Gerald Koch
Institute of Wood Biology and Wood Protection
Federal Research Centre of Forestry and Forest Products
Leuschnerstr. 91, D-21031 Hamburg, Germany
gkoch@holz.uni-hamburg.de

Stefan F. Ledig
Technische Universität Dresden
Institut für Verfahrenstechnik und Umwelttechnik
Mommsenstrasse 13, D-01069 Dresden, Germany
Stefan.Ledig@mailbox.tu-dresden.de

Wiesław Olek
Faculty of Wood Technology
Agricultural University of Pozna
ul. Wojska Polskiego 38/42, 60-627 Pozna , Poland
olek@au.poznan.pl
xii Fundamentals of wood drying

Bart Paarhuis
TNO TPD, Models & Processes Division
P.O. Box 155, 2600 AD Delft, The Netherlands
Paarhuis@tpd.tno.nl

Joëlle Passard
Université Henri Poincaré, IUT, Nancy-Brabois,
F-54 601 Villers-lès-Nancy cedex, France
Joelle.Passard@iutnb.uhp-nancy.fr

Patrick Perré
AgroParisTech
ENGREF 14, rue Girardet F-54 042 Nancy cedex, France
perre@nancy-engref.inra.fr
Present: Université Paris-Saclay, CentraleSupélec patrick.perre@centralesupelec.fr

Jörg B. Ressel
University of Hamburg, Dept. of Wood Science
Leuschnerstr. 91, D-21031 Hamburg, Germany
ressel@holz.uni-hamburg.de

Michel Riepen
TNO TPD, Models & Processes Division
P.O. Box 155, 2600 AD Delft, The Netherlands
Riepen@tpd.tno.nl

Jarl-Gunnar Salin
Swedish Institute for Wood Technology Research
P.O.B. 5609, SE-114 86 Stockholm, Sweden
JarlGunnar.Salin@tratek.se

Knut M. Sandland
Norwegian Institute of Wood Technology
Forskningsveien 3B, N-0371 Oslo, Norway
knut.sandland@treteknisk.no

Michalis Skarvelis
N.AG.RE.F./Forest Research Institute of Athens
Terma Alkmanos, 115 28 Athens, Greece
Skmi@fria.gr

Ian Turner
School of Mathematical Sciences,
Queensland University of Technology,
GPO Box 2434, Brisbane, Q4001, Australia
i.turner@qut.edu.au
1 From fundamental to practice:
the interaction chain
Patrick Perré1 and Wiesław Olek2
1
LERMAB
Laboratoire d´Etude et de Recherche sur le Matériau Bois
UMR1093 INRA, ENGREF, UHP,
ENGREF 14, rue Girardet 54 042 Nancy Cedex
France
perre@nancy-engref.inra.fr
2
Faculty of Wood Technology
Agricultural University of Pozna
ul. Wojska Polskiego 38/42
60-627 Pozna
Poland
olek@au.poznan.pl

ABSTRACT
Timber drying is depending on many different parameters. In general, practical objectives are
related to low drying cost, short drying time and reasonable drying quality. To achieve an
overall acceptable result of these mutual interfering targets compromises are required. The
complex system of intricate physical and mechanical processes during timber drying includes
external heat and mass flow, coupled heat and mass transfer within wood, shrinkage induced
stress and deformation, mechanical memory behaviour of wood. Additional effects are
chemical alterations leading to discolorations and physical phenomena, as cell collapse.
Based on these intricate phenomena, practical rules to obtain short drying time and good
drying quality are discussed in detail separately and mutually.
This introductory chapter points out the importance of fundamental aspects of wood drying,
which is precisely the objective of this book. Consistently, this chapter ends with a reading
guide of all following chapters that include equilibrium states, basic physical and mechanical
phenomena and the intricate coupling between these phenomena.

KEYWORDS: kiln drying, drying time, drying quality, drying cost, heat and mass transfer,
stress, deformation, moisture content gradient, drying defects, drying conditions, process
optimization.

Fundamentals of wood drying, Patrick Perré editor, pages 1 to 19.


2 Fundamentals of wood drying

1.1 The need for timber drying

To induce the ascent of sap in tree, the menisci present in the leaf stomata pull up
water. Because most trees are more than 10 m high, the absolute liquid pressure in
the sap column must be negative. No gaseous phase can exist in such conditions.
The vascular system developed in trees has many other implications for the drying
process (Perré and Keey 2006):
• Because the system is designed for longitudinal sap flow from the roots to the
canopy, the wood material is strongly anisotropic.
• Because of negative pressure, the vascular system must be able to support a gas
invasion due to injury or cavitation. This is the role of bordered pits, or vessel-
to-vessel pits. These anatomical features may inhibit dramatically the fluid
migration in the wood.
• In heartwood, due to metabolite deposition, aspiration or closure of bordered
pits and/or tylose development, the permeability is often reduced by one or
several orders of magnitude.
• The wood is fully saturated in the sapwood part of logs (an air-free sap column
is required to obtain negative pressures), whereas the heartwood zone is
generally only partly saturated.

The moisture content of wood (dry basis) is defined by the following mass ratio:

mass of water
X=
oven-dry mass
[−] (1.1)

This dimensionless quantity is often expressed in percent. In practice, MC (for


Moisture Content) is commonly used instead of X. Moisture in wood is
distinguished between the liquid sap present unattached in the cell lumens, called
free water, and the water molecules held in the cell walls, called bound water, with
an activity less than 1, whose departure produces shrinkage (see Chapters 4 and 5).

Before sawing, due to the large size of logs, the moisture content changes very
slowly. Therefore, unless natural drying has taken place since cutting, the initial
moisture content of timber at the beginning of the drying process, Xini, corresponds
to the moisture content of standing trees. Table 1.1 indicates some orders of
magnitude generally observed for the moisture content of greenwood. Indeed,
because the sapwood part is fully saturated, the maximum moisture content Xsat in
this zone can be calculated by assuming that the entire pore volume is filled with
water:

φ ρℓ ρ
X sat = with φ = 1 − 0 (1.2)
(1 − φ ) ρs ρs

where φ is the porosity, ρ0 the basic density (oven-dry mass/green volume), ρ s the
density of the cell-wall substance ( ρ s ≅ 1530 kg·m-3) and ρℓ the sap density
( ρ ℓ ≅ 1000 kg·m-3).
From fundamental to practice: the interaction chain 3

Table 1.1. Typical values of moisture content found in greenwood

Moisture content
Wood type Dry basis (%)
Sapwood Heartwood
Softwoods 150 to 200 40 to 80
Hardwoods 80 to 120 60 to 100

Under given conditions of temperature and relative humidity of the surrounding air,
the moisture content of wood varies until it reaches equilibrium (see Chapter 4). The
value at equilibrium is called equilibrium moisture content (Xe or EMC). This value
depends on the region, the season, the period of the day and the kind of usage. In-
use moisture content is defined as the average equilibrium moisture content that a
wooden structure attains in usage. Typical values are in the range of 8 to 12% for
indoor conditions and 12 to 15% for outdoor conditions.

To avoid dimensional changes and defects, e.g. warping, checking, splitting, due to
shrinkage or swelling in the finished products, timber has to be dried to a final
moisture content close to this average in-use moisture content. This is the main
motivation for timber drying. Note also that lowering the moisture content of a
piece of wood allows its water activity to be reduced, hence to prevent biological
attacks.

1.2 The raw product: density, anisotropy and variability

In the field of wood science, to define the raw product is a wide subject, out of the
purpose of this introductory chapter. In this section, the reader will just find a clear
definition of wood density and specific gravity. We also emphasise on the high
anisotropy ratios and wide variability found in wood.

Because wood mass as well as wood size change with the moisture content X, it is
necessary to specify the moisture content at which the density is determined.
Therefore, the wood density is defined as the mass of a sample mX divided by its
volume VX :

m X  kg 
ρX =   (1.3)
VX  m3 

At oven-dry conditions, index d is used instead of X.

The total mass mX obviously depends on X. In addition, within the domain of bound
water (see Chapters 4 and 5) swelling / shrinkage occurs, hence the sample volume
depends on the moisture content. This is why the apparent density of a wood sample
varies dramatically with the moisture content (Siau 1984). To avoid this problem
and to be able to compare different samples, different species and measurements
4 Fundamentals of wood drying

performed in different countries, the basic density ρ0 is a convenient way to


quantify the wood substance contained in a sample:

oven-dry mass  kg 
ρ0 =  3 (1.4)
green volume m 
The green volume is the volume of the completely swollen sample (moisture content
greater than the fiber saturation point).

The specific gravity is the ratio of the density of a material to the density of water
(dimensionless). Similarly to basic density, the basic specific gravity, BSG, is the
ratio of the basic density to the density of water (Wheeler et al. 1989):

ρ0
BSG =
ρℓ
[ −] (1.5)

Typical values of basic density are in the range of 300 to 450 kg/m3 for softwoods
and 400 to 750 kg/m3 for hardwoods. Note however that more than 4000 wood
species exist in the world and that extreme values range from about 100 to
1200 kg/m3. In softwoods, the basic density varies dramatically from earlywood
(250 kg/m3) to latewood (can be more than 1000 kg/m3).

As general trend, regardless of species, one may note statistical relationships


between basic density and most of the wood properties, which affects the drying
behaviour: shrinkage values, modulus of rigidity and rupture, hardness, thermal
conductivity are among properties that tend to increase with basic density. On the
contrary, permeability and mass diffusivity decreases as basic density increases
(Perré and Turner 2001a, b).

Finally, one have to keep in mind that, as a result of the secondary growth, wood is
formed to insure important roles in the living trees, such as sap ascent, mechanical
support, durability. Therefore, as a material wood is orthotropic and strongly
anisotropic. Table 1.2 indicates some order of magnitude for dimensionless
anisotropy ratios found in wood for the most important properties involved in
drying. These ratios together with the cylindrical geometry are of utmost importance
to explain the drying behaviour. For example, the ease of fluid migration in wood,
(i.e. the permeability), is by far the property that presents the highest anisotropy
ratio. The reduction in wood permeability from sapwood to heartwood affects
particularly the longitudinal direction in hardwoods (esp. for ring-porous species
developing tyloses) and all directions in softwoods.

Finally, even in the same species, wide variations can be found in wood properties.
Even though one part of these variations are under genetic control, the most
important variations regarding drying behaviour originate from the within-tree
variability: sapwood vs. hardwood, juvenile wood vs. mature wood, reaction wood
vs. normal wood (Perré and Keey 2006).
From fundamental to practice: the interaction chain 5

Table 1.2. Order of magnitude of the dimensionless anisotropy ratios encountered


in wood relative to tangential value (after Perré and Keey 2006)

Property Direction
T R L
Stiffness 1 2 20
Shrinkage 1 0.5 ≈0
Thermal conductivity 1 1.5 2
Mass diffusivity 1 1-2 20
Permeability 1 1-10 100-105

1.3 Wood drying: a complex system of interfering processes

Section 1.1 told us that drying is inevitable and requires a huge amount of water to
be removed from the wood. Understanding modern timber drying requires the
complex relations between many causes and effects to be understood. The causes are
related to process parameters, as e.g. drying temperature, relative humidity, drying
medium (air, water vapour, nitrogen…) velocity in convective drying, board
thickness,… as well as physical, mechanical and chemical wood properties, mostly
changing with wood moisture content and temperature.

Drying requires energy to be transferred to timber. Heat transfer, heat conduction,


mass transport, i.e., moisture movement in any form (liquid flow, gas flow, surface
diffusion etc.) and mass transfer (water evaporation), are coupled processes
interfering each other. Additionally theses processes are depending on wood
properties, time and environmental conditions. Change in moisture content and
related dimensional changes result in drying stresses, which involve the memory
effect of wood. Additional effects are chemical alterations leading to discolorations
and physical phenomena, like cell collapse.

The following chapters of this book - written by different authors - explain particular
aspects of the drying process and drying behaviour respectively in details. The
content of these contributions is more centred on scientific aspects than on practical
application, nevertheless also practitioners may draw some useful advantages. This
drying operation lasts a certain time and requires suitable equipment, energy, labour
and so on, which builds up the cost, but it also may induce some degrade (checking,
discoloration…) which can be summarised in terms of product quality. From an
industrial point of view, the goal is quite simple to define (Fig. 1.1).

However, in the right balance between time, cost and quality lies the complexity of
the wood drying operation (Perré 2001b). The three major aims have to fit together
to optimise the drying process in practice. However, none of these aims is
independent, they influence each other in a complex manner.
6 Fundamentals of wood drying

• Drying time is certainly the simplest factor to evaluate. However, it has


many important implications: fixed capital, possibility to satisfy demands of
the customers, storage capacity…
• Drying quality (or product quality) concerns mainly moisture content
specifications (average MC, MC variance within the kiln load and within
each single board, MC gradient across the board thickness…) and mechanical
quality (straightness, stability, absence of casehardening, and absence of
checking). Depending on the species, additional aspects can be of essential
importance too: discoloration, sticker staining, collapse…
• Drying cost results from the choice of the equipment and it is also the
compromise chosen between the drying time and quality of the dried product.
Additionally some more topics have to be considered here:
- labour input (sticking, loading and unloading, process supervision…),
- maintenance, repair, tear and wear of the overall drying plant,
- losses due to timber distortion and degradation during the drying process,
- thermal energy, electricity and water or steam consumption,
- availability of thermal energy (e.g. shut down of the boiler during
weekends),
- kiln capacity and utilization of its capacity,
- interest on timber value and technical investment.

THE DREAM THE NIGTHMARE

1$
1 $1 $

LOW
!# SHORT "
COST

J TIME

HIGH
COST L LONG
TIME

GOOD QUALITY BAD QUALITY

Figure 1.1. The wood drying process: the industrial dream (after Perré 2001b)

Obviously, it is very difficult to optimise all three aims at once. In practice, a


compromise has to be found between cost, quality and time. This is a complex
optimisation, depending on individual circumstances and limitations, e.g.,
determined by the specific kiln and its utilisation, the experience of the operator,
market conditions and demands, etc… So, for a factory already equipped with kilns,
a satisfactory compromise has to be found from practical experience between drying
time, cost and product quality. For example, in the case of overloaded equipment,
some losses due to low quality may be tolerated to reduce drying time. On the
From fundamental to practice: the interaction chain 7

opposite, despite longer drying times, carefully chosen drying schedules will be used
in order to make the most of valuable species.

However, certain situations may break this way of presenting the compromise
between the three major criteria. The overall satisfaction is sometimes very poor:
inappropriate drying schedules always increase the drying cost because of a poor
balance between product quality and drying time. Other situations may occur, in
which both the drying time and the product quality can be improved for an
equivalent cost. This is probably a very pleasant configuration, for which the overall
satisfaction exceeds 100%. Such a situation has two possible explanations:

• The given situation is really bad. In that case know-how transfer is required,
i.e. training and technical formation. Experienced consultants, often situated
midway between research and practice are in charge of this effort (in many
European countries, this is a typical task for scientists, employed, e.g. at the
Technical Centre for Wood and Furniture, CTBA in France, or at the Federal
Research Centre for Forestry and Forest Products and the University of
Hamburg, Centre for Forest Products in Germany).
• The previous situation is already rather good. Then the superordinated
objective lies in innovation. Theoretically, this is the job of fundamental and
applied research. Fundamental research deals with understanding and hence
predicting, while applied research provides the industry with tools being able
to implement and adjust new procedures and processes. For example, using
the concept of glass transition temperature with tropical hardwoods, the
drying time has been cut by a factor two in existing industrial kilns, keeping
the same quality (Aguiar and Perré 2000b).

In the following, some basic measures will be drawn from the physical and
mechanical phenomena to ensure a fast and satisfying drying process. Some
experimental proofs of these rules will then be proposed and explained with an
industrial implementation.

1.4 Physics and mechanics of timber drying

In terms of science (physics, chemistry, mechanical and chemical engineering…)


wood drying involves very complex and coupled phenomena. A detailed description
of these phenomena, including transfer processes and mechanics, together with the
relevant physical and mathematical formulation, is the primary aim of this book. The
most important phenomena involved in wood drying and their interrelated
complexity are outlined in Fig. 1.2, namely:

• coupled heat and mass transfers in porous media,


• development of moisture content and temperature fields,
• shrinkage field and induced drying stresses,
• mechanical behaviour of wood leading to stress reversal and casehardening,
• specific effects such as casehardening, discolorations, collapse…
8 Fundamentals of wood drying

Among all couplings between these mechanisms, some are of utmost importance to
understand drying problems:

• stress development and deformation originate from the moisture and


temperature fields (i.e. gradients across thickness, width and length of the
board being dried) through shrinkage,
• the mechanical behaviour of wood is dramatically affected by temperature
and moisture content levels, as well as by the time dependent change of
moisture content (mechano-sorptive effect),
• large displacement fields change the initial geometry significantly and
change the transfer conditions,
• stress values exceeding the modulus of rupture lead to checking that changes
significantly the transfer parameters,
• chemical degradation is strongly tied to the moisture content level, the
temperature level and the duration of these drying conditions,
• with high moisture flux values, extractive deposition may occur close to the
exchange surface, which in turn becomes impervious, leading to dramatically
long drying time and high stress levels,
• collapse is primarily due to capillary forces, but is promoted by high
temperature levels and peripheral zones under tension,
• discoloration also requires oxygen and particularly temperature and moisture
content levels.

Figure 1.2 proves the difficulty to optimise drying of a single board, but how to
handle an entire stack of boards? Indeed, a stack is built from pieces cut from
different parts of a single log:

• A flat sawn board mainly dries along its radial direction involving shrinkage
in its tangential direction: it dries fast on a high stress level with larger defor-
mation,
• A quarter sawn board mainly dries along its tangential direction involving
shrinkage in its radial direction: it dries more slowly on a low stress level
with lesser deformation,
• A sapwood piece has high initial moisture content with easy moisture
migration: it can tolerate high drying rates, in particular in the beginning of
the drying process, without any mechanical degradation,
• A heartwood piece has low initial moisture content with difficult moisture
migration. The drying process has to be carefully controlled already from the
beginning of the drying.
From fundamental to practice: the interaction chain 9

EXTERNAL CONDITIONS
SHRINKAGE FIELD
(Dry bulb, wet bulb, air flow,
(results from temperature and
pressure)
moisture content fields)
INTERNAL TRANSFER
CONSTITUTIVE EQUATION
(conductivity, diffusivity,
(elasticity, viscoelasticity,
permeability, anisotropy
mechanosorption)
ratios)
DIMENSIONS
DIMENSIONS
(length, thickness)
(length, thickness)
ANISOTROPY
HETEROGENEITY

COUPLED HEAT AND


MASS TRANSFER SOLID MECHANICS

+ + +
+ + +
+ - - - - - - -- - +
-
+ + + -+ +

TEMPERATURE AND SPECIFIC PHENOMENA DEFORMATION AND


MOISTURE CONTENT Chemical degradation, STRESS FIELDS
Discoloration, VERSUS TIME
FIELDS VERSUS TIME Cementation,
Collapse

FSP

Thickness

QUALITY OF THE
DRYING TIME
DRIED PRODUCTS

Energy Moisture Product


Content value
consumption variations

COST

Figure 1.2. Some of the phenomena influencing drying quality, but also drying time
and drying cost (after Perré 2001b)

Beside this, a stack or a complete kiln load is built from boards coming from
different trees (a kiln with a capacity of 100 m3 contains boards coming from 50 to
200 different trees). Due to the well-known variability of biological products, a wide
range of physical properties does exist. Even for the same species and the same
density, e.g. the shrinkage coefficients may vary by a factor two. Same levels of
10 Fundamentals of wood drying

variability are found for the mass diffusivity. Concerning permeability the situation
is even worse!

Consequently, wood variability is certainly one other major obstacle to the


improvement of wood drying. The great variability of the decisive properties
encountered within a stack requires a compromise to be found:

• a drying process suitable for the worse pieces leads to a very low amount of
losses to the detriment of the process duration,
• on the opposite, the process can be faster if a lower quality is accepted for the
dried products.

1.5 Measures to obtain a fast and good drying process

Understanding the mechanisms in terms of physics during wood drying processes,


partially summarised in Fig. 1.4, allows the derivation of basic rules for a fast and
good process.

1.5.1 Drying time

In order to reduce the process duration, only heat and mass transfer have to be
considered (Fig. 1.3). Some parameters depending on the kiln load are important
too, but they can not be controlled:

• thickness is a very important parameter (roughly speaking, the duration of the


drying process increases as the thickness squared),
• transfer and transportation properties of the wood (diffusivity, permeability,
capillary pressure, thermal conductivity, etc. are given by the species being
dried, and could be influenced partially by the sawing patterns during log
cutting).

In conventional drying, the controlled parameters are the dry and wet bulb
temperature (drying climate) as well as the velocity of the air flow, i.e. the drying
medium. All together these parameters determine the conditions for external
transfers:

• The “drying potential” of the air flow is defined by the heat transfer
coefficient (increases with air velocity) times the difference between dry and
wet bulb temperatures.
• The air velocity plays also an important role in the homogeneity (of wood
MC distribution) within the stack. However, its effect becomes less important
as the drying progresses and moisture migration is mostly controlled by
internal transfer and transport respectively.
From fundamental to practice: the interaction chain 11

Additionally some more subtly effects have to be kept in mind:

• The internal transport (diffusion, liquid migration) becomes easier with


increasing temperature level.
• Above the boiling point of water, an additional driving force, the total
pressure gradient, acts with a dramatic efficiency (Perré 1995). Drying is
mostly controlled by internal vaporisation under such conditions.
• The internal transport mechanisms depend on the local moisture content
(liquid migration is usually much more efficient than bound water or vapour
diffusion). In addition, diffusion becomes very slow when the bound water
content decreases towards zero.

PARAMETERS LOW VALUE HIGH VALUE

EXTERNAL
Velocity
LL J
TRANSFER Drying potential
(dry bulb – wet bulb)
LL JJ
Thickness
JJJ LLL
INTERNAL Mass diffusivity
LL JJ
TRANSFER
Thermal diffusivity
L J
Temperature
LL JJJ

Figure 1.3. Guidelines on how to obtain a fast drying operation (from J good to
JJJ excellent and from L poor to LLL disastrous) (after Perré 2001b)
In conclusion, drying time is reduced with high velocity, high temperature and low
relative humidity of the air. However, a too low relative humidity could lead to a
surface zone with low moisture content (due to a low value of Xe), and hence, a too
low moisture migration close to the surface. All high temperature configurations
(convective drying at high temperature, vacuum drying, conductive drying…), are
processes that accelerate internal moisture migration due to the overpressures
generated within the product, and thus an absolute pressure gradient.

Finally, drying in electromagnetic fields (microwave and radio frequency heating)


offers an entirely new possibility: any internal temperature can be attained without
resorting to thermal diffusion.

1.5.2 Mechanical quality

Drying stresses originate from shrinkage: as soon as the shrinkage field within the
board is not geometrically compatible, a stress field develops in the material. From
then, the mechanical behaviour of wood is the key to understand stress development,
up to the end of the process. The non-elastic part of this behaviour, mainly
12 Fundamentals of wood drying

viscoelastic and mechanosorptive effects, requires the history of the dried product to
be known for evaluating the stress field (Perré 1996). This is the reason why these
effects are usually denoted by the general expression “memory effect”. This
behaviour is responsible for the stress reversal encountered during the drying
process; it has to be taken into account for a good understanding of strain, stress and
mechanical degradation during the process. In order to reduce the stress level
throughout the entire process, hence surface checking, internal checking and the
residual stresses, several conditions should be fulfilled (Fig. 1.4):

• low shrinkage coefficients,


not under control
• small board thickness,
• low MC values between surface and centre of the board(s),
• significance of viscoelastic creep, which allows the stress level to be
released, hence the mechanosorptive creep, responsible for stress reversal, to
be lowered; such an effect is obtained through a high value of temperature,
without a too low moisture content value, which would foil the effect of
temperature (Irvine 1984).

Note that a low temperature level is sometimes desired (e.g. to avoid collapse),
while a too high temperature level may lead to thermal degradation and
discoloration.

SECOND DRYING PERIOD

PARAMETERS LOW VALUE HIGH VALUE


Tension

Compression
Thickness
JJJ LLL
Tension MC (bound) gradient
(centre - surface)
JJ LL
END OF DRYING Shrinkage
JJ LL
Compression
Surface moisture
content
LLL JJJ
Tension Temperature
J L JJ L
Compression

Figure 1.4. Guidelines on how to obtain a good product quality (from J good to
JJJ excellent and from L poor to LLL disastrous) (after Perré 2001b)
1.5.3 Recipes for obtaining both a fast and good process

Despite the large contradiction to the supervisory strategies mentioned above, a fast
and satisfying drying process should summarise these measures by determining
acceptable compromises (Fig. 1.5). In particular, three major rules have been
proposed as a guideline to obtain a good compromise (Perré 2001b):
From fundamental to practice: the interaction chain 13

1. High relative humidity: the only way to ensure a low MC gradient is the
reduction of the drying potential (dry bulb – wet bulb) as much as possible.
Additionally this precaution imposes a high EMC value (only a part of
shrinkage is expressed and the effect of temperature on the viscoelastic creep
is not inhibited by a too low MC level). However, a high relative humidity
could support the growth of mould, and/or, discolorations, e.g. blue stain.
2. High temperature: particularly a high temperature level has a positive
effect. It accelerates the internal moisture transfer and activates the
viscoelastic creep. However, care should be taken with sensitive species: high
temperature levels could increase the risk of collapse, discoloration, or even
thermal degradation of the wood constituents.
3. High air velocity: a high air velocity ensures a good homogeneity of MC
distribution throughout the stack. However, this increases the electricity
consumption and may produce, via the heat transfer coefficient, a too high
external transfer flux, which is opposite to the effect intended in rule 1.

Related to moisture transfer, rules 1 and 2 lead to an accelerated internal transport


while external transfer should be reduced. Exceeding the boiling point of water is
decisive for internal transport. However, note that these measures impose the
temperature to be high with a high relative humidity value: such conditions may be
difficult to ensure for certain kilns. This is the reason why innovative drying
procedures may need new kiln technologies. Finally, it has to be mentioned that, too
often, the cumulated effect of temperature and moisture content on the viscoelastic
behaviour is disregarded in the optimisation of drying schedules.

BAD GOOD
Sharp profiles Smooth profiles with low drying time

FSP
FSP

Thickness Thickness

PARAMETERS VALUE WHY ? PROBLEMS

Drying potential LOW Unique way obtain a low MC May be difficult to ensure for
(dry bulb – wet bulb) gradient certain dryers
Imposes a high Equilibrium MC Other problems (coloration, fungi)
(few shrinkage, high creep)
Temperature HIGH Easy internal transfer + high Can increase the risk of collpase
viscoelastic effect
Velocity HIGH Good homogeneity throughout the Electricity consumption
stack

Figure 1.5. Guidelines on how to obtain a fast drying operation together with a
good product quality (after Perré 2001b)
14 Fundamentals of wood drying

Obviously, the situation strongly differs between individual species. Usually,


softwood species are quite easy to be dried. Among all existing species, no doubt
that oak is among the worst (depending on their country, other scientist would say
that the worst one is eucalyptus, and they would be right!) At least, in the case of
oak, we know why drying is so difficult:

• the fibre zones with high density together with tyloses, blocking the vessel
lumina, produce an anatomical pattern with very difficult moisture migration
behaviour,
• its shrinkage coefficients and shrinkage anisotropy are rather high,
• on the anatomical level, the existence of zones with high shrinkage and zones
with low shrinkage, induces an important micro-stress field, which promotes
checking (Perré 2001a),
• the parenchyma cells of oak are prone to collapse,
• discoloration easily develops in oak,
• oak properties present a dramatic variability…

Consequently, the industrial drying of oak remains a big challenge. Conventional


kiln drying lasts several weeks up to several months, primarily depending on the
timber thickness. Drying of green boards is economically almost impossible: the
timber has to be pre-dried in a timber yard for months or even years before a final,
technical drying operation could be carried out to obtain suitable final moisture
content.

1.6 Innovations in the industrial drying process

Due to the difficulties in drying species like oak, it offers an outrageous role for
innovations of the drying process. On the other hand, species prone to fast and good
drying allow any audacious technology to be invented. Several technological
innovations are currently proposed. Most of them originate from ideas described and
also patented several decades ago:

• development of pre-drying kilns offers a better process control and enables


the replacement of air drying under uncontrolled conditions,
• use of superheated steam/vacuum dryers (good results were obtained with
thin boards, but larger dimensions are still problematic - Ressel 1994, Joyet
and Meunier 1996),
• use of radio frequency/vacuum dryers. This approach allows to cumulate two
major advantages a) to enhance the internal moisture transport by exceeding
the boiling point of water (reduced environmental pressure, vacuum) and
b) to supply the wood pieces with energy, whatever its thickness. Several
prototypes are used in laboratories. Some industrial plants exist, for softwood
but also for hardwoods as, e.g. oak (Smith et al. 1994, Avramidis 1999,
Kobayashi et al. 1999, Resch and Gautsch 2000),
From fundamental to practice: the interaction chain 15

• development of very fast microwave based drying processes able to drop the
conventional batch process for a continuous drying method (Antti and Perré
1999),
• development of a glass-transition based drying schedule applied in industrial
kilns for tropical hardwoods (Aguiar and Perré 2000b),
• use of total pressure cycles to enhance the departure of liquid and vapour
from the wood pieces (Hayashi et al. 1994, Schill et al. 1994).

In the domain of wood drying, we often heard that fundamental research and applied
development are opposed by their skills, their motivations and their expectations
(Welling 2000). However, for all of these new and innovative processes, it is
interesting to observe that we gradually shifted from empirical knowledge towards
scientific knowledge based on universal laws of science and invariable facts of the
physical world. This is important to notice than innovative processes can not be
imagined by thinking about principles behind wood drying: heat and mass transfer,
mechanical aspects, chemical problems and so on. This is probably why we can say
that wood drying became a mature field (Milota 1999).

Nevertheless, we have no choice but to notice that all these new and innovative
processes come from experimental knowledge, even though they are based on
sophisticated experimental devices. Indeed, comprehensive tools able to catch the
whole physical, chemical and mechanical complexity of drying are still under
construction.

Among promising tools, numerical simulation projects from others. A numerical


model does not produce knowledge, but it synthesises a large among of knowledge
to provide prediction possibilities. The simulation of the wood drying process has
been the subject of numerous works, including physical and mechanical formulation
and numerical solutions. Nowadays, thanks to the increased performance of
computers, the numerical simulation becomes a very promising tool to cope with the
tricky conditions encountered with drying (Perré 1996, 1999). These models can
now predict the mechanical quality and help in imagining new procedures over a
large range of drying conditions (Perré and Passard 2004).

However, very few results are found in the literature where numerical modelling has
been able to improve or optimise drying schedules. Among the reported works in
this area, Salin (1999) could improve industrial timber drying from a numerical
model. In order to be predictive, some parameters have been fitted from industrial
trials. Although this approach is limited to drying conditions very close to those
employed to tune the model, this work is a good success. Some other works propose
a comprehensive algorithm that could optimise drying schedules by minimising
certain criteria (Carlsson et al. 1996, Carlsson and Arfvidsson 1999).

This research is very interesting, however, still a lot has to be done: the optimised
drying schedules computed at present would lead to a disaster concerning the
product quality. The reason for that is obvious; optimising the drying of an industrial
16 Fundamentals of wood drying

kiln imposes the wood variability, as well as the change of drying conditions
throughout the stacks to be accounted for. Therefore, multiscle approaches are very
promising in addressing this question of drying optimisation by using numerical
simulation (Perré and Rémond 2006, Perré et al. 2007). Finally, the numerical
simulation is the only tool available to analyse phenomena too complex for intuitive
understanding (Antti et al. 1999, Ledig and Militzer 1999a, Bucki and Perré 2003).

1.7 About this book

This introductory chapter emphasised phenomena involved in wood drying as well


as the ways of their coupling. The goal of this book is to present the fundamental
aspects of wood drying. It should allow the reader to learn about each single aspect
and understand their interactions during the drying process. Including the present
chapter, the book comprises 13 chapters giving an overview of the most important
features of wood drying, ranging form sorption equilibrium concepts to the complex
problem of air flow in actual industrial kilns (Table 1.6).

Table 1.3. Book organisation

Transfers Mechanics Other


Overview 1 From fundamental to practice: the interaction chain
3 Psychrometrics 5 Shrinkage, 2 Wood anatomy -
Equilibrium 4 Sorption swelling and warp an introduction
isotherms in wood caused by moisture
changes
6 Bound water 8 Creep -
Single migration in wood deformation in
phenomenon 7 Fluid migration drying wood
in wood
9 External heat and 11 Stress 12 Discoloration of
Coupled mass transfer development wood during drying
phenomena 10 Coupled heat 13 Airflow within
and mass transfer kilns

1.7.1 Wood and air

Chapters 2 describes the structure of the material (wood) at different levels, i.e.
macro-, micro- and submicroscopic. The crucial differences occurring in softwood
and hardwood species are discussed. The significant influence of the wood structure
variation is emphasized in the context of the drying behaviour. The basic data on the
structure are given for better understanding heat and moisture transfer mechanisms
in wood. The important features of the chemical composition are also related to the
properties describing the wood-water system.
From fundamental to practice: the interaction chain 17

The important properties of moist air, being the most common drying medium of
wood, are presented in Chapter 3. The basic definitions and empirical formulas are
delivered. The relations for relative and absolute humidity, enthalpy, density etc. are
derived. The processes of moist air, as related to the drying of wood, are discussed
and illustrated.

1.7.2 Equilibrium states

The physics of water sorption by wood is discussed in Chapter 4. The concept of


the equilibrium moisture content is given together with the factors influencing it.
The methods for determining sorption isotherms are presented. Different sorption
models are reviewed. The procedure of fitting sorption models to the results from
sorption experiments is given. The integral and differential heats of sorption are also
discussed.

Changes of wood dimensions due to shrinkage and swelling in the hygroscopic


range are presented in Chapter 5. The influence juvenile wood, compression wood,
low and high density of wood was discussed in the relation to different shrinkage
potential. The wood deformations during drying were extensively discussed. The
influence of the shrinkage anisotropy on the deformations is emphasized. Spiral
grain is pointed out as the main factor of twist deformations. The methods for
reduction the dimensional changes of wood have been presented.

1.7.3 Single phenomena

The bound water movement within wood is presented in Chapter 6. The water
diffusion is discussed for steady- and unsteady-state under isothermal and
nonisothermal conditions. Fick’s first and second laws are used in the description of
the diffusion. The steady- and unsteady-state methods for the diffusion coefficient
measurements were presented. Non-Fickian behaviour of wood was mentioned. The
phenomenon of thermal-diffusion was explained and its importance was showed in
developing coupled heat and mass transfer models in wood drying.

Chapter 7 presents fluid (i.e. gas and liquid) transport mechanisms in wood. The
explanations are given for the phenomena of fluid migration in the case of single
fluid phase in the porous medium. The most deviations from Darcy's law are
described. Finally, the concepts of capillary pressure and multiphase migration, two
major mechanisms involved in unsaturated porous media hence during drying, are
presented and analysed.

Chapter 8 reviews the significance of creep deformation to the wood drying


process. Creep in drying wood is explained as a consequence of the internal stresses
as caused by shrinkage. The importance of creep is pointed out because of the
release of drying stresses and reduction of cracking. Two different mechanisms of
creep in wood are described, i.e. the simple viscoelastic creep and the
mechanosorptive creep. The influence of high temperatures on creep is showed. The
18 Fundamentals of wood drying

importance of the phenomenon is related the effectiveness of high temperature


drying.

1.7.4 Coupled phenomena

The interaction between the wood surface and the moist air is discussed in
Chapter 9. The heat and mass transfer coefficients are applied for the description of
the interaction. The prediction of the heat transfer coefficient is presented for
various situations of timber stacking. It also was showed that the mass transfer
coefficient may be estimated from the same correlations as the heat transfer
coefficient after taking into account the analogy between heat and mass transfer. The
concept of the surface emission was mentioned.

Chapter 10 presents the coupled heat and mass transfer phenomena involved in
wood drying. The physical mechanisms as well as related coupling are explained.
The physical description of the coupled transfer is presented. Different formulations
of coupled transfer modelling are compared. Specific case studies have been shown
to present the potential as well as limitations of the different formulations.

Stress development during wood drying as the main reason for wood mechanical
degrade during drying is discussed in Chapter 11. The mechanical aspects of wood
drying are used to explain the stress development during drying. The mathematical
description of the stress development is given. The numerical model is presented
and examples of predictions of the stress development are depicted. The results of
the analyses were related to earlier mentioned problems of drying time and quality.

1.7.5 Miscellaneous

Chapter 12 discusses and classifies various types of discolorations in green and kiln
dried wood. The physiological reactions of living parenchyma cells are the
dominating factor responsible for the discolorations of green and stored round wood.
The discolorations during kiln drying and steaming are primary caused by chemical
reactions of some wood compounds and cell wall tissue. It was showed that the
discolorations depended on drying parameters and chemical composition. There are
also reviewed measures of the discolorations.

The airflow within a kiln is extensively discussed in Chapter 13. The flow pattern
inside and outside the timber stack was presented. The special attention was paid on
pointing out the causes and effects of flow maldistribution. There were also
suggested methods for obtaining sufficiently uniform flow field in the kiln. There
were also discussed theoretical and experimental approaches.
From fundamental to practice: the interaction chain 19

1.8 Conclusion

This first, and introductory chapter, presented timber drying as an optimisation


between three major aims: drying time, drying quality and drying cost. The
complexity of the wood drying process is described as seen from different
viewpoints, i.e. industrial or customer requirements on one hand and fundamental
knowledge of all physical and mechanical processes involved on the other hand. It
also proposes a reading grid of all chapters of this book, which includes equilibrium
states, basic physical and mechanical phenomena and the intricate coupling between
all these phenomena.
2 Wood anatomy - an introduction
Jörg B. Ressel
University of Hamburg
Dept. of Wood Science
Section of Mechanical Wood technology
Leuschnerstr. 91
D-21031 Hamburg
Germany
ressel@holz.uni-hamburg.de

ABSTRACT

Wood as a natural material is described on the various biological or structural levels, i.e.,
chemical compounds, cell wall structure, wooden tissue and its elements as well as the
macroscopic morphology. Also reaction wood and juvenile wood is discussed. The aim is to
understand this unique material, in particular in relation to its behaviour during drying. This
paper is limited to wood only, the structure of phloem, wood formation, the tissues of roots
and branches are not considered.

Also physical and mechanical properties, e.g., moisture content and density variation within
the growing stem is ignored. Nevertheless these are characteristic features, which have to be
taken into account for satisfying drying results. Here, the reader is referred to the literature.

KEYWORDS: wood, chemical compounds, cellulose, hemicelluloses, lignin, cell wall, middle
lamella, primary wall, secondary wall, wall sculpturing, pit, fibre, tracheid, vessel, softwood,
hardwood, reaction wood, juvenile wood, earlywood, latewood, resin canal, ray, xylem,
phloem, bark, pith, cambium.

Fundamentals of wood drying, Patrick Perré editor, pages 21 to 50.


22 Fundamentals of wood drying

2.1 Introduction

To understand the behaviour of wood during drying as well as the resulting drying
quality, it is very useful to have some basic knowledge of the structure of this
natural raw material. Final moisture content and its variation, moisture movement,
internal stresses, different types of distortion, as cup, twist and bow, checking and
varying discolorations are complex reactions of the wood being dried.

The specific properties of wood and its behaviour depends significantly on the
chemical compounds as well as on its microscopic and macroscopic structure. In the
literature five structural levels could be distinguished.

• Integral level (stem structure)


• Macroscopic level (tissue structure)
• Microscopic level (cell structure)
• Ultrastructural level (cell wall structure)
• Biochemical level (biochemical composition of the cell wall)

On each level structural variations influence the physical and mechanical properties
of wood. These differences are considerably important for different applications of
wood (Seeling 2002).

This chapter will start with the chemical composition of the material, followed by
the submicroscopic structure of the cell wall, the microscopic structure of the
wooden tissue and finally end with the macroscopic structure of both softwoods and
hardwoods. Bark or phloem is not considered. Also primary and secondary growth
due to meristem1 activities, e.g., cell division in the cambial region, cell
differentiation and cell growth, are not considered in detail in this paper.

2.2 Chemical composition

Wood is a typical organic material (i.e., it contains carbon). With very small
variations between different wood species the material is formed by three main
elements carbon, oxygen and hydrogen. Nitrogen as well as some additional
inorganic elements, e.g., sodium, potassium, calcium, magnesium and silicon are
also essential material compounds, mostly involved in the metabolism of the living
cells during wood formation and growth.

1
A plant tissue consisting of actively dividing cells giving rise to cells that differentiate into
new tissues of the plant. The most important meristems are those occurring at the tip of the
shoot and root (see apical meristem) and the lateral meristems in the older parts of the plant
(see cambium; cork cambium) (acc. to Dictionary of Science, Oxford University Press).
Wood anatomy - an introduction 23

Table 2.1. Elementary composition of wood (Haygreen et al. 1982)

Elements Content
Carbon C 49 %
Hydrogen H2 6%
Oxygen O2 44 %
Nitrogen N2 <1%
Inorganic Elements Na, K, Ca, Mg, Si << 1 %

At a higher level these elements form typical macromolecules, i.e., polymers


representing the main cell wall compounds which are present in all wood species. To
a lesser extend low molecular-weight substances, e.g., extractives and inorganic
substances (ash) can also be found, representative for individual wood species. The
general scheme in Figure 2.1 summarises the molecular compounds of wood.

WOOD

Low-molecular-weight substances Macromolecular substances

Organic Inorganic Polysaccharides Lignin


compounds compounds (Holocellulose)

Extractives Ash Cellulose Polyoses WKG-053 / JBR

Figure 2.1. General scheme of molecular components of wood (acc.


to Fengel and Wegener 1989)

In contrast to the elementary composition of wood, there is a significant variation in


the main macromolecules composition – i.e., cellulose, hemicelluloses and lignin –
between hardwoods and softwoods as well as between individual wood species.

Table 2.2. Macromolecular substances of the wood cell wall (Haygreen and
Bowyer 2002)

Content [%]
Compounds
Softwoods Hardwoods
Cellulose 40...44 40...44
Hemicelluloses 30...32 15...35
Lignin 25...32 18...25
24 Fundamentals of wood drying

Polysaccharides, i.e., carbohydrate polymers, are based on monosaccharide


molecules, e.g., hexoses (glucose, galactose and mannose) and pentoses (arabinose
and xylose). These molecules form larger chains by glycosidic linkages. The
monosaccharides are all produced by photosynthesis and biosynthetic pathways.
Depending on the linkage, two different polymers are found in wood: amylopectine
starch, composed of α (1→4) linked glucose and cellulose - the most abundant poly-
saccharide - composed of β (1→4) linked glucose units. The size of the linear
polymer cellulose is expressed by the so called DP (degree of polymerisation =
molecular weight of cellulose / molecular weight of one glucose unit), which varies
between 3500 and 15000 for plant celluloses.

The DP of hemicelluloses is much lower than that of cellulose (DP ≈ 150...200).


Additionally these polymers (heteropolysaccharides) may be highly branched and
they may contain some other sugars than glucose, e.g., different hexoses and
pentoses. Depending on the basic sugar molecules galactan (→ galactose units),
arabinan (→ arabinose units), xylan (→ xylose units), and glucomannan (→ glucose
and mannose units) build up supramolecular structures in the wood cell walls. There
are also significant differences between softwoods and hardwoods related to the type
and the content of the various hemicelluloses in the cell walls (see Table 2.3).

The supramolecular structures (elementary fibrils and microfibrils, see below) built
up by cellulose and hemicellulose chains are responsible for the tension strength of
the wooden tissue. In contrast lignin, the third major cell wall polymer, is
responsible for compression strength and stiffness of the material. The complex
natural organic polymer lignin is formed by three basic monomers, i.e., p-coumaryl
alcohol, coniferyl alcohol and sinapyl alcohol. Theses monomers are based on a
phenyl ring linked with a propane side chain and with a hydroxyl group; additionally
one or two methyl groups are linked to the phenyl ring. Softwood lignin consists
mainly of phenyl propane units of the guaiacol type (coniferyl alcohol), whereas in
hardwood lignin both, coniferyl and sinapyl alcohol occur. In wood cell walls these
basic monomers join together to form a three dimensional polymer wrapping the
aggregated cellulose and hemicellulose chains and filling up most of the microvoids
between them.

The cellulose is held in larger aggregates, so called elementary fibrils, consisting of


approximately 37...42 parallel cellulose chains; microfibrils are composed of several
elementary fibrils shrouded by shorter hemicellulose chains. As a larger
morphological unit so called fibrils are composed of bundles of cellulose
microfibrils. In contrast to these amorphous aggregations the cellulose fibrils show
two different regions: crystalline regions with highly ordered parallel chain
arrangements and amorphous, i.e., less ordered chain regions without any parallel
arrangement. The crystalline regions account for 60...70% of the elementary and
microfibrils respectively. They are difficult to penetrate with water as well as
chemicals (some chemicals will break crystallites though).
Wood anatomy - an introduction 25

120@10-10m
Microfibril
Elementary fibril

30@10-10 m crystalline
region
30@10-10m

Amorphous Amorphous Fibril amorphous


region region region
Water
molecules
Crystalline
cellulose

Lenght approx. 300@10-10m

Figure 2.2. Basic cellulose morphology (elementary fibril – microfibril - fibril) and
water adsorption within the amorphous regions as well as on the surface of crystal-
line regions

The elementary fibril is the basic unit of cellulose morphology (about 30 x 30 10-10
m in transverse direction and about 300 10-10 m length between amorphous zones).
The cross section of microfibrils is approx. 120 x 120 10-10 m and of indefinite
length, composed of several elementary fibrils. Fibrils are composed of a bundle of
microfibrils (approx. 2000 10-10 m in width and indefinite length). When viewed in
cross-section they appear as organised blocks within a matrix of hemicelluloses and
lignin.

The chemical compounds of wood and their distribution is quite uneven across the
entire material as well as across the particular cell wall. On a larger scale within a
growth ring of wood, earlywood and latewood have to be distinguished. On a
smaller scale the cell wall is also a compound material consisting of several different
layers.

Another group of organic compounds in wood are the so called extractives, low
molecular, soluble substances. The typical extractive content is approx. 2...10% of
wood. Sometimes they are also classified as secondary metabolites, i.e., waste
products of plant metabolism, playing a non intrinsic role in physiological processes.
The extractive content varies within individual species, and it is also site specific.
These compounds are responsible for many useful and practical properties of wood.
Their occurrence could be advantageous, but they can also contribute to some
problems under certain circumstances: staining of heartwood, improved durability of
heartwood, enhanced flavour of wine and spirits, inhibition of resin curing and
gluing processes, discoloration of coatings or wood finishing, etc., just to mention a
few.
26 Fundamentals of wood drying

Table 2.3. Distribution of chemical compounds of spruce wood (Helm 2002)

Lignin content % of total % Poly-


[% by weight] Lignin saccharide
Whole
Wood 92 10 8
Middle lamella 55 13 45
Primary wall 22 77 78
Secondary wall
Early Wood
Middle lamella 85 12 15
Primary wall 50 16 50
Secondary wall 22 72 78
Late Wood
Middle lamella 99 9 1
Primary wall 60 10 40
Secondary wall 22 81 78

Table 2.4. Polysaccharide composition of the cell wall (Helm 2002). The * indi-
cates minor polysaccharides of the middle lamella and the primary wall, called
com pound middle lamella; the term pectins is related to arabinan and galactan

Secondary wall
Middle lamella and
Primary wall S1 S2 S3
Spruce [%] [%] [%] [%]
Galactan* 16 8 -- --
Cellulose 33 55 64 64
Glucomannan 8 18 24 24
Arabinan* 29 1 1 --
Xylan 13 18 11 13
Birch
Galactan* 17 1 1 --
Cellulose 41 50 48 60
Glucomannan 3 3 2 5
Arabinan* 13 2 2 --
Xylan 25 44 48 35

Most trees produce extractives to provide protection from predators (insects,


termites, bacteria, fungi) of the cell wall, i.e., trees produce their own natural
protectives, which differ in their ability to prevent degradation.
Wood anatomy - an introduction 27

Extractives are also present in bark, where the content ranges from 20 to 40%. The
bark can comprise up to 15% of the tree’s biomass, thus debarking might be
necessary for some processes, e.g., for pulping. Typically, bark is burned for fuel
value, but it could also be used for mulch production if the toxicity or pathogen
spread potential is not too high. Investigations on bark processing for medicines or
use in adhesives are a promising research field. Bark extractives may also influence
the pH-value and thus influence curing processes in wood based panel
manufacturing (Fengel and Wegener 1989, Helm 2002).

According to Helm (2002) wood extractives may be classified as follows.

Table 2.5. Major classes of wood extractives (Helm 2002)

Volatile Oils Wood resins Fats and Waxes


mainly softwoods mainly softwoods minor, less than 0.5%,
• Terpentines, based on 5- • Acidic diterpenes (20- • Suberin
carbon building block carbon molecules)
• Monoterpens are fragrant • Resins of softwoods,
(10 carbon molecules Basis of tall oil
• Turpentine, tropolenes

Tannins Lignans Carbohydrates


hardwoods and softwoods hardwoods and softwoods typically food reserves
• compounds for • optically active
tanning leather lignin dimers (controlled
• Hydrolyzable free radical coupling
tannins built upon glucose process)
and gallic acid
• Condensed
tannins, flavonoid-based

2.3 Cell wall structure

The wooden cell wall consists of several layers, which are more or less distinguished
from each other, not only by the macromolecular composition, but also by the
cellulosic fibril orientation, i.e., the angle between the fibril orientation and the
longitudinal axis of the cell.

Between two adjacent cells lies a highly lignified region, the so called middle
lamella. Both middle lamella and adjoining primary wall are sometimes referred to
compound middle lamella. In direction to the cell lumen the secondary cell wall is
following, subdivided into three zones, termed as S1, S2 and S3. Related to the
mechanical properties of wood the S2 layer is the most important part of the cell
wall. Due to its thickness most of the lignin is concentrated in this part of the wall –
but it does not have the largest percentage compared with other wall layers. The
varying fibril orientation in the particular layers causes a mechanical locking effect
28 Fundamentals of wood drying

leading to very high stiffness of the overall cell. This effective structure may be
compared with plywood or similar wood based composites. A transition zone lies
between adjoining layers, characterised by a slightly change of the microfibril angle;
even across the very thick S2 several sublayers have to be distinguished.

Table 2.6. Layered cell wall structure and average distribution of main macro-
molecular compounds (Helm 2002)

Primary cell Secondary cell wall


wall S1 S2 S3
-6 -6
Thickness -6
≈ 0.1 10 m 0.1 10 m 0.6 10 m 0.1 10-6 m
Microfibril Open network of 50...70° micro- 10...30° micro- 60...90° micro-
angle microfibrils fibril angle fibril angle fibril angle
Distribution Lignin 8.4 % Lignin 10.5 % Lignin 9.1 % Lignin --- %
of main Cellulose 0.7 % Cellulose 6.1 % Cellulose 32.7 % Cellulose 0.8 %
compounds Hemicell. 1.4 % Hemicell. 3.7 % Hemicell. 18.4 % Hemicell. 5.2 %

The figures 2.3. to 2.5. of the secondary wall are related to the formation of the
particular wall layer. According to Haygreen and Bowyer (2002) this layered
morphology is the key to the behaviour of wood. The S1 and the S3 layers of
softwoods are formed by 4...6 layers of clustered microfibrils (lamellae), while the
number lamellae comprising the S2 layer varies from 30...40 in thin-walled
earlywood cells to 150 and more in latewood cells (according to Kollmann and Côté
1968). Therefore the thickness of the overall S2 layer varies from 0.1 µm to 0.6 µm.

Figure 2.3. The structure of a cell


appears even more complicated if the
high powered magnification of a
transmission electron microscope is
used. The micrograph shows a cross
sectional view of a latewood pine
tracheid. The cell wall of the tracheid
is divided into various layers as
indicated on the picture. The S1, S2,
and S3 layers make up the secondary
wall. The Pr layer is the primary wall
S3 (SWST 2002)
S2
S1
Pr
Wood anatomy - an introduction 29

Figure 2.4. Mor-


Middle lamella Thickness approx. 2 nm, most- phology of the
ly consisting of lignin, pore dia-
meter approx. 10 nm wooden cell wall
layers
Primary wall aslo called cambial wall, net-
work of microfibrils, thickness
approx.0,1...0,2 µm, short
chains of holocellolose

S1 Threee secondary wall layers inclu-


ding particular transition zones,
varying orientation of fibrils within
each (sub)layer. S2 approx. 30...150
S1 / S2 lamellas from fibril bundles, overall
thickness 1...10 µm. Fibrils consit of
about 200...1000 cellulose chains,
embedded in a matrix of hemicellu-
loses and lignin.
S2
S1 thickness approx. 0,1...0,2 µm,
fibril angle 50...70E

S2 / S3 S2 thickness up to 0,6 µm,


fibril angle 5...20E

S3 slightly inclined fibrils, but


S3 = T without any strict formation as
within the S2 layer, also termed
as tertiary layer T

Warty layer adjoining the S3 or T layer, thin mem-


brane including incrustations, mostly
proteins and lignin, thickness approx.
0,1 µm

Figure 2.5. Average distribution of


organic components across a softwood
Lignin cell wall (Panashin and de Zeeuw 1980
cited in Haygreen and Bowyer 1982)
Hemicelluloses

Cellulose

S1 S2 S3
Secondary wall
Compound middle lamella (ML and Pr)
30 Fundamentals of wood drying

2.4 Cell wall sculpturing

2.4.1 Pits

Parenchyma cells fulfil two primary functions in the wooden tissue: storage and
conduction of food material. These cell type forms thin secondary walls. In contrast
other cells, i.e., tracheids and fibres, form thicker secondary walls. These cells serve
either as conduction paths (thin-walled earlywood tracheids) or they provide the ne-
cessary mechanical stiffness (thick-walled latewood tracheids) to the wood in which
they occur. In all cell types, the secondary wall layers show specific openings or
gaps, so called pits. These pits serve as canals, allowing liquids to flow between
adjoining cells (water flow to the top of the tree, controlled by the evaporation from
leaves or needles). Pit structure and shape varies with the type of the cells (see
below) they are connecting and also with the wood species concerned. Three
different types of pits have to be distinguished (Fig. 2.6):

• simple pits connecting two parenchyma cells


• bordered pits connecting to prosenchyma cells (tracheids)
• half-bordered pits connecting parenchyma cell to prosenchyma cell

The size of the pits is also significantly influenced by the connection between
earlywood and latewood cells - they disrupt the microfibril orientation. The pit
membrane is formed by the primary cell walls of both adjoining cells and the
sandwiched middle lamella. It is pectin rich and reinforced by cellulose microfibrils.
During cell wall growth in softwoods, the membrane centre is changed through
accumulation of dense packed microfibrils which are forming the so called torus,
held by radialy arranged microfibrils formed over the existing primary wall.

Middle lamella and Middle lamella and


primary wall primary wall
Secondary wall
Secondary wall

Pitmembrane Torus Porus


Pit canal
Pit chamber
Margo

Pit canal
Pit chamber

Simple Pit Bordered Pit Half-borederd Pit

Figure 2.6. Various types of pit pairs (acc. to Grosser 1977, Fig. 7)
Wood anatomy - an introduction 31

Subsequently the pectine matrix of the so called margo is decomposed


enzymatically giving enough open voids for water flow between the adjoining
tracheids. The principal structure of hardwood bordered pits is similar, but the
membrane is quite different: no torus is developed and no dissolution of primary
wall components is visible. With further growth of the cell wall the secondary wall
is overarching the margo and torus, forming the pit chamber. Due to this structure
bordered pit pairs are like a valve: the elastic margo moves laterally sealing the
porus. When sealed, the pit is called an aspirated pit. The closing effect is caused by
injuring the tree or by heartwood formation. Once closed the torus will never return
back to the original position, any pathway through the porus is blocked for ever.
With some wood species, e.g., some pines, the warty layer covering the secondary
wall reduces the tightness of aspirated pits to some extend (for further details see
Figures 2.7 to 2.11).

Figure 2.8. Detail of a bordered pit pair in


a softwood showing margo and torus (acc.
to Bauch et al. 1972 in Sengenbusch et al.
2002)

Figure 2.7. Inside of a pit chamber of


a bordered pit pair, portions of three
longitudinal tracheid lumens (dark
areas) can be seen here (SWST 2002)

As already mentioned pits are connecting adjoining cells. Half-bordered pits


between parenchyma cell and prosenchyma cell, e.g., tracheid and ray-parenchyma
are of specific shape, which is a very useful feature for softwood identification on a
microscopic scale (see Figure 2.10). Similar differences are visible with vessel
pittings in hardwood species (see Figure 2.11).
32 Fundamentals of wood drying

Figure 2.9. Bordered pits, restricted to certain areas. Pits are concentrated on the
overlapping ends of tracheids and in the crossfield area, where tracheids are
adjoining rays. The left side of the left picture shows some late wood tracheids at
the end of an annual ring. On the bottom of the right hand picture cross field pitting
between tracheids and ray parenchyma is visible (Bergfeld et al. 2002)

Window-like pit in pine


(Pinus sylvestris L.)

Pinoid pits in pine


(Pinus ssp.)

Piceoid pits in spruce


(Picea abies (L.) Karst.)

Scaliform pitting Opposite pitting Alternating pitting

Cupressoid pits in juni-


per (Juniperus com- Figure 2.11. Typical shape of bordered pits
munis L.)
between adjoining vessel elements in
hardwoods (acc. to Grosser 1977)
Taxodiod pits in fir
(Abies alba Mill.)

Figure 2.10. Typical shape of half-bordered pits between axial tracheids and radial
ray parenchyma in softwoods (acc. to Grosser 1977)
Wood anatomy - an introduction 33

2.4.2 Spiral thickenings

These helically oriented thickenings on the inside of the cell wall are spirally
arranged ridges of microfibril bundles formed during the development of the S3
layer. They are distinctly separated from the S3 layer and only rarely parallel
arranged to microfibril orientation of this wall layer. They occur in some softwood
species, e.g., in Douglas-fir (Pseudotsuga menziesii (Mirb.) Franco) and in yew
(Taxus baccata L.), and more commonly in hardwood species, e.g., in maple, some
birches and cherry (Haygreen and Bowyer 1982).

Figure 2.12. Spiral thickenings on


a hardwood cell wall surface
(Arrighetti 2002)

2.4.3 Warty layer

Some softwood species, e.g., of the genus Callitris vent., (Cupressaceae), show a so
called warty layer on the surface of the S3 layer. The appearance is somehow
modulated or branched. The wart morphology could be a useful indicator for
determination of Callitris species.
Figure 2.13.
Warty layer in
Callitris (Heady
1997)
34 Fundamentals of wood drying

2.5 Microscopic structure of wood

On a larger scale a cross section of a wooden stem shows two concentric regions:
xylem and phloem. Between the two the so called cambium is located. The cambium
is just a few cells wide in the radial direction and it is responsible for the so called
secondary growth. It produces new phloem cells on the outside, and new xylem cells
on the inside. During their growth and maturation from the fusiform initial they
change in size, shape and structure to meet specific requirements of their genetically
predetermined function. Overall the wooden tissue consists of single elements, i.e.,
more or less long, slender and hollow cells. In the standing tree these cells are
mainly oriented in axial direction and only a smaller amount of them is radially
oriented.

Depending on their function and shape different cell types could be distinguished.
Upon a comparing examination, softwood tissue seems to be more simple and
uniform than hardwood tissue. The following Table 2.7 summarises the cell types of
wood.

Table 2.7. Cell types in wood (acc. to Haygreen and Bowyer 2002)

Softwood (Conifer) Hardwood (Angiosperm) Both

Cell type Tracheid Vessel Elements Fibers Parenchyma


conduction, mecha- conduction mechanical storage, defense
Function nical support support

Shape long, narrow short, wide short, narrow various


overlapping ends, connected end to connected with oriented radially in
Arrange- connected with pits end into long narrow pits rays, and in hard-
ment vessels woods, longitudi-
nally near vessels

The major botanical distinction between softwoods and hardwoods lies in the
structure of their wood. In softwoods, the cells that serve to transport water also
provide mechanical support for the stem. In higher developed hardwoods, some cells
are specialised in water transport, and others in providing mechanical support. In
hardwoods the water conducting cells are termed as pores or vessels. They are
commonly much larger in diameter than the cells, termed as tracheids, in softwoods.

2.5.1 Softwood structure

Softwoods basically consist of tracheids and parenchyma cells, oriented mainly


(approx. 90...95%) in the longitudinal direction, only a small amount is radially
oriented within the so called rays. The length to width ratio of the longitudinal
tracheids is about 100 : 1 or even greater (length approx. 3...4 mm, diameter approx.
Wood anatomy - an introduction 35

25...45 µm). Tracheids are closed at their ends; the overlapping regions of these
tracheids show specific bordered-pit patterns on their radial faces, allowing water
flow to the treetop. Earlywood tracheids are thin-walled, whereas latewood tracheids
are thicker-walled cells (see also Figure 2.9). The margin of a growing period or
annual ring is marked by an abrupt change in cell wall thickness. The transition bet-
ween earlywood and latewood regions within one annual ring varies from a
gradually change to sharply delineated zones, e.g., in Douglas-fir and in hemlock or
fir. Only few softwood species have thin-walled longitudinal parenchyma cells in a
smaller extent (approx. 1...2% of the overall tissue), e.g., redwood and some cedars.

In the tissue of some softwoods so called resin canals are found, e.g., in the genera
pine and spruce. These canals form a 3-D network in longitudinal and radial
direction in the tree. In radial direction these canals are embedded in rays. The
canals are intercellular spaces surrounded by specialised parenchyma cells, so called
epithelium cells, secreting resin into the canal system in response to injury or
harmful and traumatic events.

In all wood species particularly radially oriented cells could be found. In softwoods
these so called rays consist entirely either of parenchyma cells (homocellular ray),
or both ray parenchyma and ray tracheids (heterocellular ray). On tangential
surfaces the shape of the rays looks like tiny little spindles, just one or two cell(s) in
width (uniserate - biserate rays) and many cells in height. In some cases so called
fusiform rays contain a resin canal. Some of the ray tracheids show thickened walls,
sometimes teeth, slanting ends and nodular end-walls.

As already mentioned above, radial tissue is contacting longitudinal cells always on


the radial surface of the longitudinal tracheids, showing a particular crossfield
pitting, i.e., half-bordered pits as shown, e.g., in Figure 2.10. The size of the pits
varies depending on earlywood or latewood tracheids on one side. The shape of this
crossfield pitting is dependent on the species and thus a very useful indication for
anatomical identification (Haygreen and Bowyer 2002).
36 Fundamentals of wood drying

Collapsed sieve element


Phloem / bast parenchyma
Active sieve element
Cambium Cambium
Earlywood tracheids
Latewood tracheids
and resin canal

RC

Fusiform
ray and
resin canal

Ray

Uniserate ray acc. to Mägdefrau, 1951

Figure 2.14. Three-dimensional representation of xylem and phloem of larch (acc.


to Mägdefrau 1951; drawn from Grosser 1977, Fig. 6)

Figure 2.15. SEM 3-D picture of an


eastern spruce wood block. Most cells
run longitudinally, but some cells run
horizontally. The hole on the
transverse surface is called a resin
canal. The majority of the cells shown
here are called "longitudinal
tracheids". Take note on the wood
structure which appears differently on
each surface (SWST 2002). The rays
are mainly uniserate, a cut fusiform
ray is visible on the transverse surface.
Thickness growth is from left to right,
shown by changing size and wall
thickness of earlywood and latewood
tracheids
Wood anatomy - an introduction 37

Figure 2.16. Cross section of


redwood (Sequoia
sempervirens (D. Don) Endl.),
showing an abrupt transition
between earlywood (light
coloured area, larger diameter
cells) and the latewood (dark
coloured area, smaller
diameter cells) (SWST 2002).
Thin-walled longitudinal
parenchyma is marked by
deposits in their lumina

2.5.2 Hardwood structure

From an evolutionary view hardwood species are much younger than softwood
species. Thus tissue and cells respectively are much more adapted to meet specific
requirements most efficiently, as e.g., in relation to water transport, food storage and
mechanical support. There are also some transition cell forms, e.g., between fibres
and tracheids as well as between tracheids and vessel elements, representing
evolutionary steps. In hardwood xylem at least four different cell types could be
found, with a share of 15% or more each in the total xylem volume.

The most significant evolutionary step compared to softwood xylem is the


development of vessels, particular conducting pathways composed of single
elements, i.e., thin-walled cells with more or less open end plates. Distribution and
size of these distinguished longitudinal cells vary very much between hardwood
species, offering an useful feature for identification.

Rays in hardwoods are much larger in height and width compared to softwoods;
they can constitute up to 30% or more of the total xylem volume, with an average of
about 17%. There exist only homocellular rays in hardwoods, containing solely
parenchyma cells. They are not aligned in straight radial rows as in softwoods, in
particular in the vicinity of larger vessels distortions from the purely radial
orientation occur.

Fibres, fibre tracheids, vessels and parenchyma cells as longitudinal cells in


hardwoods vary considerably in size and shape. Fully developed after maturation
from the fusiform initial in the cambium they appear quite different.

Vessels are, as mentioned above, particular tubelike components of hardwood


structure. Due to the thin cell wall and the large diameter of these vessels, they
appear as holes in the transverse surface and thus also termed as pores. Size and
arrangement of vessels are used for hardwood identification, in addition to axial
parenchyma distribution and the size of rays. Two basic vessel arrangements have to
38 Fundamentals of wood drying

Figure 2.17. Hardwood cell types (drawn from Grosser 1977, Fig. 16)
Wood anatomy - an introduction 39

be distinguished, i.e., ring-porous and diffuse-porous hardwoods. During the


development of the vessel element diameter the radially oriented ray cells are
meandering in contrast to the straight softwood ray alignment. As vessels are
specialised on water conduction, the end walls of the single elements are typically
perforated by unrestricted holes during the cell maturation process. The pattern of
these perforation plates is either simple, scalariform or foraminate. The lateral
connection between adjoining vessels is provided by numerous closely packed
bordered pit pairs (see above Figure 2.17). The pattern type of the perforation plates
as well as the pitting arrangement are not changing within a given species; both are
very useful for microscopic wood identification. Connections to fibre tracheids,
longitudinal and ray parenchyma and other cell types also form typically pits.

Collapsed sieve element


Active sieve element
Cambium
Vessel element
Ray
Fiber

Ray Par.

Vessel element with


scalariform perfo-
Ray ration plate (L)

acc. to Mägdefrau, 1951

Figure 2.18. Three-dimensional representation of xylem and phloem of birch (acc.


to Mägdefrau 1951, drawn from Grosser 1977, Fig. 15 ). Within the xylem region
take note on the following features on the different surfaces:
Transverse Growth ring end and succeeding growth ring, earlywood and
latewood fibers (distinguished by cell wall thickness), vessel-to-vessel
pitting.
Radial parenchyma cells in homocellular ray and axial parenchyma,
indicated by horizontal sieve plates, vessel elements with scalariform
perforation.
Tangential fiber-to-fiber connections, vessel element with scalariform
40 Fundamentals of wood drying

Unlike parenchyma cells, which die later during heartwood formation, the large
vessel elements die quite soon after formation, i.e., they are soon loosing their
nucleus and cytoplasma. During the transition from sapwood to heartwood and also
caused by injury, fungi infection, or drought, so called tyloses can develop into the
lumen of vessels of some hardwood species. These structures are ”balloon-like
extensions of a parenchyma cell that protrude into the lumen of a neighbouring
xylem vessel or tracheid through a pit in the cell wall. They may eventually block the
vessels and thus help prevent the spread of fungi, other pathogens as well as oxygen
within the plant. Tyloses may become filled with tannins, gums, pigments, etc., and
their walls can remain thin or become lignified” (Anon., Dictionary of Biology
2002).

Figure 2.19.
Parenchyma cell
Formation of tyloses in
Parenchyma cell hardwood vessels
(Anon./University of
Tylosis Wisconsin 2002)

Tylosis

Vessel wall Vessel wall

Vessel lumen

In wood morphology, fibres and fibre tracheids are another specific kind of cells. As
long, tapered and usually thick-walled cells they provide mechanical support to the
xylem structure. The conduction function is of minor importance for these cells.
Compared to softwood fibres, i.e., tracheids, they are significantly shorter with an
average length of about 1 mm. They also tend to be more rounded than the nearly
rectangular shape of softwood tracheids. At the end of the annual growth they
become radially flattened. Pitting between fibres and adjacent cells is either
bordered or half-bordered pits depending on the type of the neighbouring cell.
Pitting between fibres and vessels is rarely observed. An additional fibre type
termed libriform fibre is mostly marked by simple pits.

On a transverse section longitudinal parenchyma as storage elements are recognized


by thin cell walls. These cells are either long and tapered or brick-shaped epithelium
around gum canals. Often crosswalls divide these longitudinal cells into smaller
sections. The occurrence of axial parenchyma in hardwoods is dependent on species
and once more an additional aid for wood identification.

As already mentioned hardwood rays are much larger than softwood rays; the width
in tangential direction covers up to about 30 cells. The rays consist only of
parenchyma cells (homocellular rays), but they are different in shape and
configuration. Most of these brick-like parenchyma cells are radially oriented with
their length axis, and they appear to be lying down. In some species ray parenchyma
on the upper and lower margins of the ray appears to stand upright on end with their
Wood anatomy - an introduction 41

long axis parallel to grain. These two types of ray configuration are termed
homogenous and heterogenous rays, respectively.

On tangential surfaces rays appear either arranged diffuse or into definite, tangential
oriented rows, i.e., storied rays. In the latter case they show roughly the same size in
height and width and they all begin at about the same level along the grain. Also this
arrangement of rays is useful for wood identification.

Figure 2.20.
Characteristic
vessel distribution
on the transverse
surface of
hardwoods

Ring-porous Diffuse-porous Semi-ring-porous


hardwood hardwood hardwood

Homogenous rays

Homogenous uniseriate Homogenous uniseriate Homogenous multiseriate


and biseriate
Heterogenous rays

Heterogenous uniseriate Heterogenous (type I) Heterogenous (type II)

Figure 2.21. Ray configuration in hardwoods (drawn from Grosser 1977, Fig. 21)
42 Fundamentals of wood drying

Apotracheal parenchyma Paratracheal parenchyma

Diffuse Diffuse-aggregate Banded Scanty partracheal Vasicentric Aliform

Net (ladder-like Marginal Aliformconfluent Banded confluent Unilateral (para-


tracheal and aliform)
and marginal)

Figure 2.22. Parenchyma configuration in hardwoods on the transverse surface


(drawn from Grosser 1977, Fig. 20)

Figure 2.23. SEM 3-D picture of birch


wood, showing diffuse-porous vessels,
embedded in a fiber tissues, and
radially oriented multiseriate rays
(SWST 2002)

Figure 2.24. SEM 3-D picture of ring-


porous white oak, large diameter vessels
in earlywood and small vessels in
latewood, embedded in a fibre and
longitudinal parenchyma tissue, and
meandering multiseriate rays. Some of the
large earlywood vessels are blocked with
tyloses (SWST 2002)
Wood anatomy - an introduction 43

2.5.3 Reaction wood

Reaction wood2 is abnormal wood formed in a leaning tree. Its properties differ
considerably from normal wood. The particular distribution of reaction wood in the
plant induces an asymetrical distribution of growth stresses capable of modifying the
orientation of different axes in space. In softwood trees, the reaction wood forms on
the lower side of the lean and is called compression wood. Compression wood is
often very dense, hard, and brittle. In hardwood trees, reaction wood forms on the
upper side of the lean and is called tension wood. Woolly surfaces and excessive
longitudinal shrinkage are typical characteristics of tension wood (Larson 2001).

Compression wood. Compared to normal wood this tissue is characterised by


shorter tracheids (about 30% shorter), higher lignin and hemicellulose content
(about 8...9% higher) and lower cellulose content (about 10% less). This leads to
some undesired disadvantages in using wood containing compression wood zones.
For example local density is increased, some strength properties are significantly
reduced (compared to normal wood of similar density), longitudinal shrinkage is
increased to 1...2% (compared to 0.1...0.2% for normal wood) and may be as great
as to 6...7%. This reaction wood is easily identified on smooth surfaces, in particular
in a transverse view. On the compression side of a leaning tree the growth rings
appear darker, reddish brown and exceptionally wide, and much narrower on the
opposite side. Therefore the cross section of the stem tends to be oval with an
eccentric pith in the centre. The proportion of latewood in the compression wood
zone is increased.

On a microscopic and submicroscopic level differences to normal wood become


more evident. The tips of compression wood tracheids are bent and folded. In cross
section the cells are more rounded than rectangular, showing large intercellular
spaces. The cell wall consists only of ML, P, S1 and S2 layers, the microfibril angle is
about 45° from the axial direction. Additionally the cell wall shows deep, helically
arranged checks from the lumen (see Figure 2.25) (Haygreen and Bowyer 1982).
(More details see Timell 1986).

2
For use of timber reaction wood should be avoided for a number of reasons. The dense hard
wood is less likely to accept an even stain when compared to other parts of a project. The
reaction wood is also more prone to failure under load and will crack and split more easily
when nailed or screwed. Carving and shaping can also be difficult and dimensional changes
with changing moisture levels are likely. The primary problem comes in trying to identify
reaction wood. Even an experienced woodworker can have trouble picking out reaction
wood. There are some hints that a board may contain reaction wood. Crookedness or a
sweep in the log is a sign of reaction wood. Wood fibres that are unusually dense and hard
for the species is another sign. Very small fuzzy fibres on surfaced hardwood can be a sign
of reaction wood as well as cracks and splits that pull wildly away from the board.
Fortunately reaction wood is more of an exception than a rule.
44 Fundamentals of wood drying

Figure 2.25. Compression wood. Left: Spruce, compression wood characterised as


crescent, reddish annual rings. Right: Helically checked S2 layer, inclined fibrils
(NZFRI/Anon. 2002)

Tension wood. This type of reaction wood differs very much from normal wood, e.
g., increased cellulose content and increased density. Sawn surfaces appear fuzzy
and rough. Strength properties are reduced (compared to normal wood of similar
density). Longitudinal shrinkage is up to 1%. According to Tsoumis (1968) both
tension and compression wood should be viewed with concern in bearing structures,
in particular if strength properties are of primary importance (Haygreen and Bowyer
1982). No significant coloration is marking tension wood zones. In primary
hardwood processing, stems containing tension wood also show an elliptical cross-
section. The arrangement of annual rings is similar eccentric as in softwoods
containing compression wood, but in the opposite way, i.e., the wider rings of the
tension wood zone are on the stretched side of the bent stem, branch or root.
Additionally on machined surfaces a lustrous sheen appears on some species,
whereas some tropical hardwoods show a slightly darker colour of tension wood.
Nevertheless all these distinguishing features are not very reliable.

On a microscopic level tension wood is much easier to identify. Vessels are much
smaller and fewer rays occur in this tissue. Fibre cell walls are much thicker than
normally enclosing very small lumens. Secondary walls are loosely attached to the
primary wall and thus, they are responsible for some of the differing mechanical
properties. Also, the fuzzy surface of machined wood is caused by these soft fibre
wall structures. The thick secondary wall of tension wood fibres is significantly
lesser lignified, it consists of almost pure cellulose. Due to this consistency, this
layer is termed gelatinous or G layer. The microfibrils are almost parallel to the
grain at an angle of only 5°. A more comprehensive analysis of the cell wall shows
different sequences of layering, e.g., missing S2 and S3 portions of the cell wall
where the G layer lies direct on the P and S1 layer (see Figure 2.26). The reason for
tension wood shrinkage is the loosely contact between the G layer and the remaining
Wood anatomy - an introduction 45

cell wall, which does not restrain the outer cell region (i.e., P and S1 layer) from
contraction during drying3 (Haygreen and Bowyer 1982).

Figure 2.26. Tension wood. Thick


gelatinous layer detached from the
underlying S2 wall. Note the microfibril
orientation in the S2 layer (inclined)
and the G layer (axial) (drawn from
Haygreen and Bowyer 1982, Fig 6.11).

2.5.4 Juvenile wood

During their growth all tress produce juvenile wood, i.e., the inner core of xylem
surrounding the pith. The time during which juvenile wood is formed is termed the
juvenile period. This period varies among individuals, with species, and with
environmental conditions. The transition from juvenile to mature wood is more
gradually, with a declining proportion of cells exhibiting juvenile wood
characteristics with increasing distance from the pith.

Juvenile wood Adult wood Juvenile wood Adult wood Figure 2.27.
Juvenile to mature
wood transition -
S2 fibril angle
Longitudinal shrinkage changes of wood
Moisture content
Spiral grain properties (acc. to
Density
cell length
Bendtsen 1978, in
strength Haygreen and
cell wall thickness
Transverse shrinkage Bowyer 1982, Fig.
Percentage latewood
6.2 and 6.4)

Pith 5 ...20 Bark Pith 5 ...20 Bark


annual rings annual rings

3
The microfibril angle in the S1 layer is nearly perpendicular to grain, so this layer tends to
shrink longitudinally when water is removed. A firmly attached S2 layer to the adjoining S1
layer - as in normal wood - will provide shrinkage restraint and overall longitudinal
shrinkage will be reduced significantly.
46 Fundamentals of wood drying

In juvenile wood the cells are smaller and less structurally developed than those of
the outer, mature xylem. Particular differences exist in the length of the cells as well
as in the structure of the layered cell wall. The proportion of latewood in juvenile
wood is very small, leading to a relative low density and thus low strength
properties. Another disadvantage of juvenile wood is the spiral grain, i.e., the fiber
orientation deviates significantly from the stem axis. This is the reason for extensive
twisting of dried timber containing juvenile wood. On the ultrastructural level the
microfibril angle in the S2 layer is greater than in cells of the mature tissue.
Analogous to compression wood this causes a higher degree of longitudinal
shrinkage as well as a reduced tensile strength. For practical application it is very
difficult to identify because there is no clear demarcation between juvenile and adult
wood. For further details see the literature, e.g., Haygreen and Bowyer (1982),
Larson (2001) and Zobel and Sprague (1998).

2.6 Macroscopic structure of wood

On the transverse surface of a cut stem three areas are clearly visible: pith, xylem
(wood) and phloem (bark). The central pith is usually barely visible and retains the
size all over the life. A cylinder of wood termed xylem varies in diameter with age
and rate of growth. Additionally within the xylem two wider concentric rings
defined as heartwood (inner zone) and sapwood (outer zone) could be distinguished
more or less clearly due to different colours. Finally, the bark sheath can be
subdivided into inner bark or bast (which conducts sugars) and outer bark (that
serves as a protective layer). New wood and new inner bark are added each year by
the activity of a layer of dividing cells termed cambium, sandwiched between inner
bark and sapwood. This closed cambium ring (cylinder) is not visible without any
magnification aid. New bark production is relatively small compared to new wood
production, and bark is continually being shed to the outside of the stem, thus in
older trees the greatest volume of the stem is wood. Since new wood is added to the
outside of existing wood the oldest wood is close to the pith, and the most recent is
close to the bark (see also Figure 2.28) (SWST 2002).
Figure 2.28.
Pith
Cambium
Xylem
Scheme of a
Cambium trunk cross
Phloem section (Anon.
2002)

Cork
Bark
Cambium
Wood anatomy - an introduction 47

In each growing period a new concentric annual ring or growth ring is formed, also
divided into two separate zones: earlywood and latewood, showing either a
continuous or abrupt transition between both. Cells produced at the beginning of the
growing season are commonly larger with thinner cell walls, and so this earlywood
appears less dense than the latewood produced towards the end of the season.
Although all trees produce concentric layers of wood, not all trees produce visible
growth rings, neither are all growth rings necessarily annual. In some trees seasonal
changes in wood structure may be so slight that growth rings are not evident. Under
conditions of severe drought an annual growth ring may not be produced. On the
other hand under continuously favourable conditions, such as the tropics, several
growth rings may be produced in a year (SWST 2002).

Figure 2.29. Schematic cuts trough a growing tree (Muth 2002)

An important function of wood is to conduct water from the roots to the leaves.
However, at some stages the wood cells may become blocked by air bubbles (from
injuries), tyloses, closing pits or deposition of other substances. Sapwood marks the
water conducting zone of wood. Heartwood is not longer in service for water
48 Fundamentals of wood drying

conduction. With continuous growing the outer boundary of the heartwood core is
continually moving outwards. In general an approximate balance is maintained
between new wood formation and conversion of sapwood to heartwood so that there
is always adequate conducting tissue. The conversion of sapwood to heartwood is
commonly associated with a colour change caused by the deposition of chemical
compounds or extractives. The colour change varies among species according to the
composition of the wood. The extractives also impart the durability to the wood
against fungal decay and insect attack. The degree of durability varies widely among
different species (SWST 2002).

Figure 2.30. Main anatomical directions in wood. Flat sawn and quarter sawn
board (T tangential, R radial, L longitudinal direction) (Wengert and Meyer 2002)

The appearance of sawn wood varies greatly according to the direction in which it is
cut. This is due to growth rings in the xylem and the cylindrical nature of the stem.
In wood three orthogonal planes (i.e., mutually perpendicular) are recognized,
although the stem can of course be cut in any number of intermediate planes. A
horizontal, or transverse cut through the stem shows the growth rings as concentric
circles. In structural lumber, partial growth rings are evident at the ends of timber
and the surface is known as end grain. The other two orthogonal planes of wood are
longitudinal. A longitudinal cut in a plane through the pith exposes a radial
longitudinal surface. In species with distinct growth rings, this surface will appear to
have a series of more or less parallel lines. A board cut to expose a radial
longitudinal surface is known as a quarter sawn board. A longitudinal cut in a plane
at a tangent to the surface of the stem exposes a tangential longitudinal surface.
Growth rings here will appear as a series of wavy lines or cones stacked one above
another. A board cut to expose a tangential longitudinal surface is known as a plain
(or flat) sawn board. It should be realized that only peeling, as in veneer production,
can produce a truly tangential surface, and only the one cut that is in a plane through
the pith can produce a truly radial surface. Intermediate planes of cut are commonly
referred to radial or tangential according to which they more closely approximate
(see Figure 2.30) (SWST 2002).
Wood anatomy - an introduction 49

Figure 2.31. Macroscopic view of cross sections of some wood species. Top, left
and right side: Pine, showing clearly earlywood-latewood transition and resin
canals, the tiny rays on the cross section are only visible at higher magnification
(18X). Bottom, left side: White oak, ring-porous, single earlywood and latewood
vessels, earlywood vessels partially blocked by tyloses, longitudinal parenchyma
and wide rays (18X), and right side: Birch, diffuse-porous, small meandering rays
(SWST 2002)
50 Fundamentals of wood drying

Additionally an extensive number of textbooks and scientific publications on both


general and specific aspects of wood anatomy is available. The cited literature here
is only a very small selection. Also the internet provides many interesting sites, in
particular specific lectures and courses on wood anatomy, plant physiology and
related fields. If only short explanations for specific terms are wanted, also internet
based dictionaries and hypertextbooks are available, e.g., Bergfeld et al. 2002 and
Anon., A Dictionary of Biology 2002.
Last but not least many thanks also to Dr. G. Koch, BFH Hamburg, and Prof. Dr. S.
Avramidis, UBC, for reading the paper and many specific useful hints.
3 Psychrometrics
Ryszard Guzenda, Wiesław Olek
Faculty of Wood Technology
Agricultural University of Pozna
ul. Wojska Polskiego 38/42
60-627 Pozna
Poland
olek@au.poznan.pl

ABSTRACT. The basic properties of moist air are discussed. The relations for relative and
absolute humidity, enthalpy, density etc. are derived. The empirical models for water vapour
saturation pressure and dew point temperature are given. The basic processes of moist air
are discussed and illustrated in the psychrometric chart.

KEYWORDS: Moist air, parameters determination, basic processes, psychrometric chart

Fundamentals of wood drying, Patrick Perré editor, pages 51 to 65.


52 Fundamentals of wood drying

3.1 Introduction

During the drying process energy must be supplied to evaporate moisture from the
dried material. The energy may be transferred to the material by means of the
different methods, i.e. convection, conduction, radiation or generated within the
material (the so-called high-frequency heating). However, the convection using
moist air as the heat transfer medium is still the most popular method applied in
timber kiln drying. Moreover, the combination of moist air parameters determines
the rate of timber drying as well as influences drying quality. Therefore, the
knowledge of the moist air properties is crucial for the proper understanding of
timber drying.

3.2 Moist air

Moist air is a mixture of dry air (consisting of nitrogen, oxygen, inert gasses, carbon
dioxide etc.) and water vapour. It is assumed that each component of the mixture as
well as moist air itself satisfies the perfect gas laws4. Therefore, the basic equations
and relationships for moist air are built with the use of the general thermodynamic
principles derived for gas mixtures and the principle for deriving formulas for gas
mixtures, i.e. the given quantity is the sum of products of the quantity for individual
components and corresponding, appropriate fractions.

The following relationships result from the mass balance:


• Mass of moist air (mma) is the sum of dry air mass (ma) and water vapour
mass (mv)
mma = ma + mv ; kg (3.1)
• According to Dalton’s law the total pressure of moist air (p) or atmospheric
pressure is the sum of partial pressures of dry air (pa) and water vapour (pv)
contained in the air
p = pa + pv ; Pa (3.2)
Moreover, it is assumed that each component of the gas mixture has the same
temperature which is equal to the temperature of moist air. Therefore, we get
T = Ta = Tv ; K (3.3)
It results from Dalton’s law that each component occupies the total volume V of the
mixture. Thus, the following equations result from ideal gas law:
• dry air
pa·V = na·R·T (3.4)

4
For practical applications the deviation between moist air and the ideal gas can be neglected. The
deviations of dry air are very small; the deviations of water vapour are distinctively larger. But, if
partial pressure of water vapour in moist air is only a small fraction of total pressure, i.e. less than 20
kPa, water vapour could be treated as ideal gas (Häussler 1973).
Psychrometrics 53

• water vapour
pv·V = nv·R·T (3.5)
• moist air
p·V = nma·R·T (3.6)

In order to describe conditions of moist air the following temperatures given in ºC


have to be defined:
• Dry-bulb temperature (Cdb) - measured by a thermometer which is freely
exposed to the air flow but shielded from radiation and moisture.
• Wet-bulb temperature (Cwb) - measured by a thermometer whose bulb is
covered by a wick which is kept wet by distilled and clean water, freely
exposed to the air flow and shielded from radiation. The temperature is
measured on water evaporating surface during its isobaric and adiabatic
evaporation ensuring the air to obtain the saturation during the process.
• Dew-point temperature (Cdp) - the temperature to which air must be cooled
down (at constant pressure and constant absolute humidity) to achieve
saturation. The partial pressure of water vapour in moist air (pv) at the
dew-point temperature (Cdp) is equal to the saturated vapour pressure (ps) at
the same temperature.

3.2.1 Humidity

a) Relative humidity (H)

Relative humidity is the most common parameter describing humidity of moist air.
It is mainly due to easy application of the measurement procedure, i.e. the
psychrometric method. Relative humidity is defined as a ratio of water vapour mass
(mv) contained in volume (V) to the maximum mass of water vapour (ms) which
could be contained by the air at the same temperature

mv
H= (3.7)
ms
Equation (3.7) is usually presented in the transformed form after taking into account
that mv = v·V and ms = s·V

ρv
H= (3.8)
ρs
Taking into account ideal gas law and density definition, the following equations
can be written

ρv ⋅ R ⋅ T
pv = (3.9)
Mv
54 Fundamentals of wood drying

ρs ⋅ R ⋅ T
ps = (3.10)
Mv
It results from (3.9) and (3.10) that pv /ps = v/ s and thus, the relative humidity, can
be written as

pv
H= (3.11)
ps
Formula (3.11) is valid for air parameter values which allow the assumption of ideal
gas law for water vapour.

b) Absolute humidity (Y)

The absolute humidity (Y) is defined as the ratio of the vapour water mass (mv)
contained in the volume V of moist air to the dry air mass (ma) contained in the same
volume:

mv
Y= ; kg/kgdry air (3.12)
ma
Taking into account Equations (3.4), (3.5), mv = nv·Mv and ma = na·Ma the absolute
humidity is rewritten as

M v pv
Y= ⋅ ; kg/kgdry air (3.13)
M a pa
Combining equations (3.2) and (3.11) the absolute humidity is determined in its
most common form:

Mv H ⋅ ps H ⋅ ps
Y= ⋅ = 0.622 ⋅ ; kg/kgdry air (3.14)
M a p − H ⋅ ps p − H ⋅ ps
It is assumed here that the ratio of the molar mass of water vapour (Mv) and dry air
(Ma) is constant and equal to 0.622.

c) Moisture concentration (c)

The moisture concentration in moist air is often called volumetric absolute humidity
or density of water vapour in moist air and defined as:

mv
c= ; kg/m3 (3.15)
V
Taking into account Equations (3.5), (3.11) and the relation mv = nv·Mv the moisture
concentration is given by

pv ⋅ M v H ⋅ ps ⋅ M v
c= = ; kg/m3 (3.16)
R ⋅T R ⋅T
Psychrometrics 55

d) Classification of moist air states

According to the definition of the relative humidity the following states of the air
can be listed5:
• Dry air, i.e. H = 0 and pv = 0
• Moist air, i.e. 0 < H < 1 and pv < ps
• Saturated air, i.e. H = 1 and pv = ps
• Fogged air, i.e. H = 1, pv = ps and Y > Ys – the air partially contains
moisture in the gaseous form (i.e. saturated water vapour of the amount
of Ys) and partially as liquid water (i.e. droplets of the amount of Y - Ys at
the air temperature)

3.2.2 Enthalpy (h)

Enthalpy is a measure of heat which should be delivered during the gas isobaric
heating from the datum temperature of 0ºC to the given temperature (C). Therefore,
the enthalpy is defined as
C
h = cp ⋅ C ; kJ/kg (3.17)
0

C
where c p = c p ; kJ/(kg·K) is the mean specific heat of the gas in a given
0
temperature range.

Enthalpy of moist air is the sum of products of enthalpies (h) and mass fractions (g)
of dry air and water vapour, respectively

h = g a ⋅ ha + gv ⋅ hv (3.18)

5
To distinguish moist air and vapour, respectively, from steam additional terms have to be defined
(Merriam-Webster… 2003), (Online Dictionary… 2003):
• Vapour - a substance in the gaseous state as distinguished from the liquid or solid state.
• Steam - the invisible vapour into which water is converted when heated to the boiling point.
• Saturated steam - water vapour in equilibrium with liquid water at or above the normal boiling
point.
• Dry steam - steam containing no free water particles.
• Wet steam - steam composed of water vapour mixed with droplets of liquid water.
• Superheated steam - steam heated to a temperature higher than the boiling point corresponding
to its pressure. It can not exist in contact with water, nor contain water, and resembles a
perfect gas.
56 Fundamentals of wood drying

The mass fractions expressed in terms of absolute humidity are as follows:


• dry air

1
ga = (3.19)
1+ Y
• water vapour

Y
gv = (3.20)
1+ Y
Substituting Equations (3.19) and (3.20) into (3.18) the enthalpy is given as

ha + Y ⋅ hv
h= (3.21)
1+ Y
The enthalpy given by Equation (3.21) is expressed in kJ per kg of moist air.
However, the more convenient form is to relate the enthalpy to the mass dry air

h = ha + Y ⋅ hv ; kJ/kgdry air (3.22)

Equation (3.22) can be expanded by taking into account enthalpy definitions for
individual components of moist air and the latent heat of water vaporisation (lv):

h = ca ⋅ C + Y ⋅ ( lv + cv ⋅ C ) ; kJ/kgdry air (3.23)

The specific heats at constant pressure depend on temperature. Thus, the values of
the specific heats vary with temperature ranges. For practical applications the mean
values of the specific heats in the temperature range of 0-150ºC are applied getting
the practical form of Equation (3.23):

h = 1.01 ⋅ C + Y ⋅ (2501 + 1.87 ⋅ C ) ; kJ/kgdry air (3.24)

For the fogged air a portion of moisture (Y-Ys) exists in the liquid form. Equation
(3.24) has to be modified into the form:

h = 1.01⋅ C + Ys ⋅ (2501 + 1.87 ⋅ C ) + (Y − Ys ) ⋅ 4.187 ⋅ C ; kJ/kgdry air (3.25)

where 1.01·C - the enthalpy of dry air, Ys·(2501+1.87·C) – enthalpy of the vapour
fraction in moist air, (Y-Ys)·4.187·C - enthalpy of the liquid water fraction.

3.2.3 Moist air density ( )

The molecular mass of moist air (kg/mole) is obtained using the principle for
deriving formulas for gas mixtures as well as mass fractions of dry air (3.19) and
water vapour (3.20).

1/ Ma + Y / Mv
1 / M ma = g a / M a + gv / M v = (3.26)
1+ Y
Psychrometrics 57

The density of moist air ( ) results from Equation (3.6) after taking into account
(3.26):

m p M ma p 1+ Y
ρ= = ⋅ = ⋅ (3.27)
V T R R ⋅T 1 / M a + Y / M v
After transformations and introducing the value of the ratio Mv/Ma = 0.622 the
following form is obtained

p ⋅ Mv 1+ Y p ⋅ Mv 1+ Y
ρ= ⋅ = ⋅ ; kg/m3 (3.28)
R ⋅T (Mv / M a ) + Y R ⋅ T 0.622 + Y

where p; Pa, Mv; kg/mol, T; K.

3.2.4 Empirical formulas for calculations of moist air properties

The water vapour saturation pressure (ps; kPa) used for the psychrometric
calculations can be given as (Wilhelm 1976)

6238.64
ln( p s ) = 24.2779 − − 0.344438 ⋅ ln(T ) (3.29a)
T
for 233.16 ≤ T < 273.16 K
and

−7511.52
ln( ps ) = + 89.63121 + 0.02399897 ⋅ T +
T
− 1.1654551 ⋅ 10 − 5 ⋅ T 2 − 1.2810336 ⋅ 10 −8 ⋅ T 3 + (3.29b)
−11 4
+ 2.0998405 ⋅ 10 ⋅ T − 12.150799 ⋅ ln(T )

for 273.16 ≤ T ≤ 393.16 K.

The dew point temperature (Cdp; ºC) is similarly given by a polynomial form
(Wilhelm 1976)

C dp = 5.994 + 12.41⋅ a + 0.4273 ⋅ a 2 for −50 ≤ C ≤ 0 ºC (3.30a)

C dp = 6.983 + 14.38 ⋅ a + 1.079 ⋅ a 2 for 0 < C ≤ 50 ºC (3.30b)

C dp = 13.80 + 9.478 ⋅ a + 1.991⋅ a 2 for 50 < C ≤ 110 ºC (3.30c)

where a = ln( p s ) .
58 Fundamentals of wood drying

3.3 Psychrometric chart (Mollier h-Y diagram for moist air)

Any thermodynamic state of the moist air at the pressure p is clearly determined by
two parameters from among five, i.e. temperature, relative humidity, enthalpy,
absolute humidity or density. The graphical representation of the moist air
parameters was previously made in the temperature – absolute humidity co-ordinate
system. However, the diagram constructed in the enthalpy – absolute humidity
co-ordinate system (the so-called Mollier diagram) obtained the common approval
in Europe. The present-day Mollier diagrams are completed by curves representing
different levels of relative humidity (H) and water vapour partial pressure (pv),
which is a function of absolute humidity (Y) at a given pressure of moist air (p). The
characteristic feature of the Mollier diagram is its slanting co-ordinate system.
Isenthalpes, i.e. lines of the constant enthalpy, form an acute angle with the absolute
humidity axis. The change of the tilt angle of isenthalpes in the chart allows
covering different ranges of the absolute humidity (Y) and therefore to improve the
accuracy of parameter readings from the chart. The accuracy of readings from the
chart depends also on the chart scale, the atmospheric pressure of air (i.e. position
above see level and current weather) as well as air temperature according to the
enthalpy definition (Eq. 3.17).

Figure 3.1 presents the psychrometric chart covering air parameters range typical for
air heating, ventilation, refrigeration and air conditioning, i.e. absolute humidity
varying from 0 to 0.05 kg/kg. The chart, which allows interpreting air processes in
kilns for timber drying, is presented in Figure 3.2. Its characteristic feature is the
expanded range of absolute humidity, i.e. from 0 to 0.2 kg/kg.

The curve of the saturated conditions of the air (H = 100%) divides the chart into
two areas. The moist air area is located above the H = 100% curve, whereas the
fogged air area is below the curve. When using the psychrometric chart the value of
the air pressure and the range of temperature for which the chart was constructed
have to be checked. If the current pressure of air (p) deviates significantly from the
pressure (pch) for which the chart was constructed then the true value of the relative
humidity (H) is calculated as follows:

p
H = H ch (3.29)
pch
where Hch is the relative humidity reading from the chart constructed for the
pressure pch. It follows form Equation (3.29) that air compression (p > pch) causes
the increase of the relative humidity while the air expansion (p < pch) results in the
decrease of the relative humidity.
Psychrometrics 59
h

Figure 3.1. Psychrometric chart. Absolute humidity range 0 – 0.05 kg/kg (generated
by computer program Mollier v. 1.1® by Piotr Narowski©)

p
–4
0

Figure 3.2. Psychrometric chart. Absolute humidity range 0 – 0.2 kg/kg (generated
by computer program Mollier v. 1.1® by Piotr Narowski©)
60 Fundamentals of wood drying

3.3.1 Basic processes of moist air and their interpretation in the psychrometric
chart

At a given pressure (p) moist air may be heated, cooled, humidified or dehumidified,
mixed with moist air of the other parameters etc. When at least two parameters of
the moist air are known the missing parameters can be read from the chart.
Moreover, it is possible to trace a process from the known starting point and
determine the final parameters of the air. Figure 3.3 presents some basic process of
moist air of the initial parameters determined by the “0” point.

a) Heating – process “0” “1”

The moist air heating is characterised by the constant value of absolute humidity (Y),
increase of enthalpy (h) and temperature (C) with simultaneous decrease of relative
humidity (H).

b) Cooling – process “0” “2”

The process is supposed to change the air parameters without dehumidification. The
final temperature of the air (C2) can not fall below the dew point temperature (Cdp),
i.e. C2 Cdp. The process is characterised by the constant value of absolute
humidity (Y), decrease of enthalpy (h) and temperature (C) with simultaneous
increase of relative humidity (H).

c) Cooling with dehumidification – process “0” “3”

The final temperature (C3) of the process is supposed to be below the dew point
temperature (Cdp), i.e. C3 < Cdp. The cooled down air obtains the saturation
(H = 100%) at the dew point temperature (Cdp). The further decrease of temperature
takes place at the constant value of relative humidity (H = 1) and causes the
condensation of the portion of vapour (Y2 – Y3) as well as enthalpy decrease.

d) Humidifying with steam – process “0” “4”

Mixing moist air with steam is characterised by the increase of absolute humidity
(Y), enthalpy (h) and temperature (C). Relative humidity (H) may increase or
decrease. The relative humidity decrease is achieved when the superheated steam is
used for humidifying.

e) Humidifying with water – process “0” “5”

Mixing moist air with sprayed water is characterised by increasing absolute


humidity (Y) and relative humidity (H) with the slight increase of enthalpy (h) and
decrease of temperature (C). If the water temperature (Cw) is equal to the wet-bulb
temperature of air (Cwb) then the humidification process is equivalent to the process
of absorbing evaporated free water from dried timber.
Psychrometrics 61

f) Drying – process “0” “6”

The process is performed at the constant value of the enthalpy (h). It is used as the
simplified description of air parameter changes during the air flow through timber in
the convective drying. The process is characterised by the increase of absolute
humidity (Y) and relative humidity (H) with the simultaneous decrease of
temperature (C). The possible minimum value of moist air temperature after its
passage through the timber is the wet-bulb temperature (Cwb).

g) Dehumidification – process “0” “7”

The process is characterised by the constant value of enthalpy (h) and the decrease
of absolute humidity (Y) as well as relative humidity (H) with the simultaneous
increase of the air temperature (C). The dehumidification is caused by moisture
adsorption on a surface or absorption in the whole volume of a hygroscopic material.

h) Mixing moist air with moist air

When mixing moist air of mass (ma A) and the parameters determined by the “A”
point with the other moist air of mass (ma B) characterised by the parameters
determined by the “B” point then the parameters of the resulting moist air are
determined on the straight line linking points “A” and “B” in the point “M” (Fig.
3.4). Finding the point “M” position in the psychrometric chart requires satisfying
the following condition:

ma A ⋅ AM = ma B ⋅ MB (3.30)

The parameters of the moist air after mixing are determined from the balances of
heat and moisture written as:

( )
ma A ⋅ hA + ma B ⋅ hB = ma A + ma B ⋅ hM (3.31)

( )
ma A ⋅ YA + ma B ⋅ YB = ma A + ma B ⋅ YM (3.32)
62 Fundamentals of wood drying

h = const

Figure 3.3. Graphical representation of the basic process of moist air

Figure 3.4. Graphical representation of moist air mixing with moist air
Psychrometrics 63

3.3.2 Measurements of moist air relative humidity

The determination of two parameters from among five, which form


the psychrometric chart unequivocally characterises the thermodynamic state of
the moist air. The measurements of temperature are trivial. Relative humidity is
the other most often measured parameter of moist air. The parameter is determined
by the use of different methods as well as instruments which differ in their accuracy
of the measurements.

a) Hygrometric methods

The methods are based on the hygroscopic phenomenon, i.e. on the changes of
the physical or the chemical properties of the different materials with relative
humidity changes. Mechanical hygrometers use the phenomenon of shrinking
and swelling of the hygroscopic materials. The influence of hysteresis, creep
and slow response of the materials causes that the method is less recommended for
measurements of fast and small changes in relative humidity. Non-mechanical
hygrometers primary use the phenomenon of the change of the electrical resistance
or capacity of hygroscopic substances (e.g. LiCl) or other materials usually applied
as coatings. The resistance may change several times with the relatively small
change of relative humidity. These hygrometers are susceptible to atmospheric
pollutions (dust, evaporating resin components and acids from drying timber etc.)
and moisture condensation. Moreover, they require periodical calibration also due to
an ageing effect.

b) Dew point method

The method is based on the principle of measuring two temperatures, i.e.


the dry-bulb and the dew point. The dew point temperature measurement requires
for instance a mirror surface made of a non-hygroscopic material. The mirror
surface is cooled to the temperature at which the vapour saturation pressure is
obtained and the vapour condensation begins. In practice cooling is performed to
the temperature slightly lower than the dew point temperature. However, the method
can be used for the hygrometers calibration.

c) Psychrometric method

The psychrometric method of relative humidity determination consists in


measurements of the dry-bulb and wet-bulb temperatures during an adiabatic
process of water evaporation. The difference between the two temperatures is called
the wet-bulb depression and increases with increasing evaporation rate in proportion
to the relative humidity decrease. The process performed in the psychrometer can be
traced in a psychrometric chart as a process of mixing of the air of the sought
relative humidity with liquid water at the wet-bulb temperature. The method requires
the air flow of the minimum velocity of 2 – 2.5 m/s, distilled water supply to get
the wet-bulb readings, shielding from radiation and calibration of both
64 Fundamentals of wood drying

thermometers. The relative humidity can be determined from the psychrometric


formula:

ps wb − A ⋅ ( Cdb − Cwb ) ⋅ p
H= (3.33)
ps db

where

 6.75  −5
A =  65 +  ⋅10 (3.34)
 v 
with v being air velocity in m/s. The water vapour saturation pressures at the
dry-bulb temperature (ps db) and at the wet-bulb temperature (ps wb) can be given in
arbitrary units, however, consistently in the same ones.

Acknowledgments

The authors wish to thank Professor Jörg B. Ressel from University of Hamburg,
Germany for his review of the chapter and the valuable remarks.
Psychrometrics 65

Table 3.1. Parameters of dry (H = 0) and saturated (H = 1) air at p = 1000 hPa

Density; kg/m3 Enthalpy; Saturated air


Saturation absolute
Temperature; pressure; kJ/kg
ºC humidity;
hPa dry air saturated dry air saturated kg/kg
air air
-20 1.03 1.396 1.376 -20.00 -18.62 0.00064
-15 1.65 1.368 1.349 -15.02 -12.61 0.00103
-10 2.60 1.342 1.323 -10.02 -6.08 0.00162
-5 4.01 1.317 1.297 -5.02 1.19 0.00251
0 6.11 1.293 1.272 0 9.56 0.00382
5 8.72 1.270 1.248 5.02 18.78 0.00547
10 12.27 1.248 1.225 10.06 29.57 0.00773
15 17.04 1.226 1.201 15.10 42.42 0.01078
20 23.37 1.205 1.178 20.16 57.98 0.01488
25 31.67 1.185 1.154 25.22 77.07 0.02034
30 42.41 1.165 1.131 30.30 100.75 0.02755
35 56.22 1.146 1.107 35.38 130.45 0.03705
40 73.75 1.128 1.082 40.47 167.96 0.04952
45 95.82 1.110 1.055 45.57 215.86 0.06592
50 123.4 1.093 1.028 50.69 277.56 0.08752
55 157.4 1.076 0.998 55.81 358.09 0.11619
60 199.2 1.060 0.967 60.94 464.85 0.15470
65 250.1 1.044 0.933 66.08 609.60 0.20741
70 311.7 1.029 0.896 71.23 811.89 0.28162
75 385.5 1.014 0.855 76.39 1106.4 0.39020
80 473.6 1.000 0.810 81.56 1563.9 0.55953
85 578.0 0.986 0.760 86.74 2352.0 0.85195
90 701.1 0.973 0.705 91.93 3984.9 1.4588
95 845.2 0.956 0.644 97.13 9195.3 3.3970
99 977.8 0.949 0.590 101.29 73761 27.423
4 Sorption isotherms of wood
Marián Babiak
Faculty of Wood Science and Technology
Technical University in Zvolen
T.G.Masaryka 24
960 53 Zvolen
Slovak republic
babiak@vsld.tuzvo.sk

ABSTRACT. This paper presents the physical mechanisms involved in sorption of water by
wood. Equilibrium moisture content and the factors that influence it are presented. Different
sorption models and different sorption isotherms are reviewed. The procedures how to obtain
sorption equations is given. The integral and differential heats t of sorption are discussed.

KEYWORDS: Wood, equilibrium moisture content, sorption, bound water

Fundamentals of wood drying, Patrick Perré editor, pages 67 to 86.


68 Fundamentals of wood drying

4.1 Introduction

Wood is composed mainly of cellulose, hemicelluloses and lignin. Free OH groups


that are mostly in cellulose and hemicelluloses can bind water molecules from the
surrounding air or from liquid water by hydrogen-bonds (H-bonds). Therefore, wood
belongs to hygroscopic materials, which equilibrate their moisture content (X)
according to the relative humidity (H) and the temperature of the surrounding air. If
a piece of wood is placed into air of certain H and temperature level, the wood X
either decreases or increases until equilibrium moisture content (Xe or EMC) is
reached. Xe of wood under different conditions can be given either in tables, graphs,
diagrams or equations.

The knowledge of Xe is important for the processes of hydrothermal treatment of


wood including drying and seasoning as well as for its final use. More than hundred
models of sorption isotherms can be found in literature. Most of them describe the
wood Xe as a function of H almost equally well so this cannot be the criterion of the
validity of a particular sorption isotherm model. One should test how well the model
describes other physical quantities related to sorption i.e. heat of sorption or
mechanical properties of wood.

4.2 Definitions

The amount of water vapour in air can be characterised by partial vapour pressure.
As gas pressure is generated by the change of momentum of gas molecules on the
vessel wall we can attribute a partial pressure to each gas in a mixture of gases. The
total pressure of the mixture of gases is the sum of the partial pressures of each
component. The conditions over liquid water surface are given by the surface layer
where surface tension can be observed. This layer works like a “doorman”, one that
lets quick molecules escape from liquid water and captures the slow molecules of
water vapour moving closely to the liquid surface. If liquid water is placed into a
closed vessel the fast liquid molecules with high energy evaporate. On the other
hand the slow vapour molecules are captured and condense. The equilibrium state, at
which the number of molecules leaving the liquid state per unit time is the same as
the number of molecules captured by surface layer, is called saturation state. It is
characterised by the saturated vapour pressure. The surface layer forms a potential
barrier that leaves out only molecules possessing higher energy than that one given
by the barrier. As the surface tension decreases with increasing temperature, the
potential barrier is lowered and more molecules have the chance to escape from the
liquid phase. Therefore, the saturated vapour pressure increases with increasing
temperature. This relationship is described in table 4.1. Based on these data the
following equation can be computed

p s = 1.1978 ⋅108 ⋅ exp(5205.46 / T ) [kPa ] (4.1)


where T means Kelvin temperature (Požgaj et al. 1993). The equation is valid in the
temperature interval 0-100ºC, when average error is 1.4% and maximum error 3.5%.
The ratio of water vapour pressure to the saturated vapour pressure at the same tem-
Sorption isotherms of wood 69

perature expressed in % is called relative humidity (H). This quantity is measured


usually by a psychrometer, the device working with two thermometers, namely a
dry-bulb and a wet-bulb one. The wet-bulb is wrapped in wet muslin. The air flow is
ensured by mechanic fan driven by a clockwork. The mechanism of the psychrome-
ter is the following. The wet-bulb temperature is lower than the dry-bulb tempera-
ture because water evaporates faster into the air with lower H. The heat of evapora-
tion cools the wet bulb the more if the evaporation rate is faster. The difference dry-
bulb temperature – wet-bulb temperature is called the psychrometric difference.

The evaporation of water from the free surface is characterised by the heat of evapo-
ration of water. The values of this quantity are given in Table 4.1 and the average
value can be calculated according to Clausius - Clapeyron equation:

0 d (ln p s )
hvap = −R (4.2)
d (1 / T )
0
which gives hvap = 2404.5 kJ·kg-1.

Table 4.1. Saturated vapor pressure and heat of evaporation as a function of


temperature (according to Schmidt 1969)

Temperature Saturated Heat of Temperature Saturated Heat of


[ºC] vapor pressure evaporation [ºC] vapor evaporation
[kPa] [kJ·kg-1] pressure [kJ·kg-1]
[kPa]
0 0.61 2501.6 55 15.741 2370.7
5 0.871 2489.7 60 19.92 2358.6
10 1.227 2477.9 65 25.01 2346.3
15 1.703 2466.1 70 31.16 2334
20 2.337 2454.3 75 38.55 2321.5
25 3.166 2442.5 80 47.36 2308.8
30 4.241 2430.7 85 57.8 2296.5
35 5.622 2418.8 90 70.11 2283.2
40 7.375 2406.9 95 84.53 2270.2
45 9.582 2394.9 100 101.33 2256.9
50 12.335 2382.9

If a piece of metal, wood or other material is placed into the air of a given H and
temperature and the temperature of it are lower, we can observe condensation of
water vapour on this surface. Dew-point is the temperature at which the air at the
given H becomes saturated. Sorption in general sense means the process of
equilibration of wood X with its surroundings. In a more specific sense it means
process of taking up water from the moist air and binding it to wood. Desorption is
the process of releasing water molecules. Generally we can speak about sorption of
any gas (adsorbate) by solid sample (adsorbent). In this publication, however, the
70 Fundamentals of wood drying

term sorption relates only to the sorption of water by wood and wood based
materials. The physical picture of sorption equilibrium is very similar to the
equilibrium between liquid water and saturated vapour. In this case, however, the
controlling agent is not the thin surface layer with its surface tension. The molecules
of sorbed water are bound to wood more tightly and we speak about bound water.
The heat of evaporation of bound water is higher than that of free water and depends
on wood X. The difference of the two quantities is called differential heat of
sorption. It should be stressed here, however, that sorption equilibrium is not the
thermodynamic equilibrium in its real sense. The fact that desorption leads to higher
values of Xe than sorption, is known as sorption hysteresis. The decisive parameters
for the sorption equilibrium are the H and the temperature of the surrounding air.
The relationship that describes equilibrium moisture content as a function of H at a
constant temperature is usually called sorption isotherm and it is expressed by the
respective sorption equation. The relationships among Xe, H and temperature can be
also described by tables, graphs or diagrams. In literature more than one hundred
models of sorption can be found. Some of them describe the physical picture related
to the respective model others are only mathematical representations of the Xe data.
The validity of the models and criteria for their evaluation will be discussed later.

4.3 How to measure Xe?

One of the main problems of Xe measurements is how to keep a constant value of the
relative humidity of the air. This can be done in a conditioning chamber
psychrometrically by controlling the wet bulb and dry bulb temperature. Relative
humidity as a function of these two temperatures is plotted in Fig. 4.1.

The other way of relative humidity control are saturated salt solutions (Table 4.2)
and sulfuric acid solutions (Table 4.3).
Table 4.2. Relative humidity over saturated salt solutions at 20°C (Siau 1995)

Chemical Formula H [%]


Copper sulfate CuSO4·5H2O 98.0
Sodium chromate Na2SO4·10H2O 93.0
Potassium chromate K2CrO4 88.0
Ammonium sulfate (NH4)2SO4 81.0
Ammonium chloride NH4Cl 79.5
Sodium acetate NaC2H3O2·3H2O 76.0
Sodium nitrite NaNO2 66.0
Sodium bromide NaBr·2H2O 58.0
Sodium dichromate Na2Cr2O7·2H2O 52.0
Potassium nitrite KNO2 45.0
Chromium trioxide CrO3 35.0
Calcium chloride CaCl2·6H2O 32.5
Potassium acetate KC2H3O2 20.0
Lithium chloride LiCl·H2O 15.0
Sorption isotherms of wood 71

Table 4.3. Relative humidity over sulfuric acid solutions (according to Siau 1995)

Concentration of Concentration of acid


acid solution H [%] solution H [%]
[10-3·kg·m-3] [10-3·kg·m-3]

1.05 97.5 1.30 58.3


1.10 93.9 1.35 47.2
1.15 88.8 1.40 37.1
1.20 80.5 1.50 18.8
1.25 70.4 1.60 8.5

40

35 0

10
WETT-BULB DEPRESSION (°Cˇ)

30

20
25

30
20 RH (%)
40
Pa

15
,3 k

50
10 1

10 60
70
5 80
90
0
10 20 30 40 50 60 70 80 90 100 110 120 130
DRY-BULB TEMPERATURE (°C)

Figure 4.1. H as a function of dry-bulb and wet bulb temperature (according to


Siau 1995)

Sorption measurements are performed in conditioning chambers or dessicators. Very


precise measurements are done in vacuum sorption apparatus that is kept in water
basin of constant temperature. The measurements itself consist in the determination
of the mass at equilibrium. For dessicator measurement the scales are used, while in
72 Fundamentals of wood drying

vacuum sorption apparatus the mass change is measured indirectly through the
length change of the helix made usually of silicon. The time to achieve equilibrium
moisture content depends on several factors. The main factor is the characteristic
dimension of the sample. According to the Fourier number the time increases with
the square of this dimension and therefore it is reasonable to choose the smallest
samples. Sometimes even sawdust is used for this purpose. In case of small samples
it is reasonable to choose the sample geometry in such a way that the largest surface
is perpendicular to the longitudinal direction. The influence of different factors on Xe
is described below.

4.4 Factors affecting wood equilibrium moisture content

The wood equilibrium moisture content is influenced by the relative humidity of the
air, temperature, species and “specimen history”. The H is the decisive factor and its
increase causes the increase of Xe. With increasing temperature this quantity
generally decreases. One should distinguish, however, between the influence of
momentary temperature that is described in sorption models together with the
influence of relative humidity and long-term influence of temperature that is usually
included into the specimen history. The term “specimen history” used by Wengert
(1979) covers up a lot of factors like hysteresis, stress, temperature and various
treatments i.e. mechanical, chemical and radiation treatment.

4.4.1 Influence of species

Species are considered as a factor that might influence wood equilibrium moisture
content. The reason for it is in different chemical composition of hardwood and
softwood species as well as different content of extractives in wood. The content of
the basic chemical wood constituents for softwood and hardwood is given in
Table 4.4.

Table 4.4. Content of basic wood components in hardwood and softwood

Component Softwood Hardwood


Cellulose 48-56 46-48
Hemicelluloses 23-25 26-35
Lignin 26-35 15-28

Christensen and Kelsey (1959b) came to the conclusion that cellulose contributes
roughly 46% to the total amount of sorption, the hemicelluloses 37% and lignin 16%.
There is also opinion, however, that hemicelluloses contribute to the total sorption more
than cellulose (Stamm and Loughborough 1942). It should be stressed here,
however, that these constituents can show completely different behaviour in wood
structure than that one chemically separated. There are not many sorption
experiments in literature that would completely characterise samples used for
Sorption isotherms of wood 73

sorption from the viewpoint of the chemical composition. Wang and Cho (1993)
measured Xe of six taiwanian species:
• Taiwan red cypress ( Chamaecyparis formosensis Matsum.)
• Japanese cedar (Cryptomeria japonica D.Don.)
• Taiwania (Taiwania cryptomerioides Hay.)
• Griffith’s ash (Fraxinus gryffithii C.B.Clarke)
• Taiwan zelkova (Zelkova formosana Hay)
• Southern red oak (Quercus falcata Michx. var falcata.)

The first three species are softwoods, the second three are hardwoods. The average
content of hollocellulose was 67% for the former and 74% for the latter, the content
of lignin being 34% and 24%, respectively. Xe for these species was measured at
three temperature levels of: 20, 30 and 40°C. According to chemical composition
one should expect higher values for hardwoods than for softwoods. The average
values of Xe for the species and all the three temperatures plotted against relative
humidity for hardwoods and softwoods are in Figure 4.2 for desorption.

32

28 o
s 20 C
24 o
s 30 C
20 o
s 40 C
EMCF[%]
Xe [%]

o
16 h 20 C
o
12 h 30 C
o
8 h 40 C

0
0 20 40 60 80 100
H [%]
RH[%]

Figure 4.2. Equilibrium moisture content (Xe) of hardwoods (h) and softwoods (s)
for desorption (according to Wang and Cho 1993)

Figure 4.2 shows the equilibrium moisture content obtained for these species during
desorption. It is seen in these figures that the influence of species seems to be even
higher than that of temperature especially in the range of H from 30 to 80%.
However, the final value is influenced by many factors and mutual interactions can
yield different results. Wood extractives are believed to decrease sorption by
blocking sorption sites. According to Wengert (1979) the proportion of
hemicellulose, holocellulose and lignin may slightly influence the sorption between
species but much of the variations are caused by extractives. As reported by Spalt
(1997) the removal of extractives leads to higher values of Xe. On the other hand
74 Fundamentals of wood drying

Cooper (1974) observed a decrease of this value for black walnut sapwood and
heartwood after partial removal of extractives. The problem of these comparisons is
that sorption measurements are time and cost demanding and there are not many
results in literature with all aspects described. Usually the measurements are very
precise from the viewpoint of one factor and in that case the influence of it on Xe can
be confirmed. In less precisely designed experiments and in practice this influence
can be suppressed by other factors.

30

24 o
s 20 C
o
s 30 C
18 o
s 40 C
EMC[%]
Xe [%]

o
h 20 C
12 o
h 30 C
o
h 40 C
6

0
0 20 40 60 80 100

H [%]
RH[%]

Figure 4.3. Equilibrium moisture content (Xe) of hardwoods (h) and softwoods (s)
for sorption (according to Wang and Cho 1993)

4.4.2 Influence of sorption hysteresis

As it was stated earlier, the term “specimen history” includes a lot of factors namely
hysteresis, stress, temperature and various treatments i.e. mechanical, chemical and
radiation. Sorption hysteresis means not only that the value of Xe obtained in
desorption is higher than that obtained in sorption. The two curves – sorption and
desorption – meet in the point where relative humidity equals zero. The sorption –
desorption curves show that depending on the sorption history any point in the area
of sorption – desorption loop can be reached according to the sorption history. The
question is whether the sorption and desorption curves meet in the upper part, i.e. for
H = 100%, where the state of saturation is supposed. The situation is even more
complicated with the definition of the term “fibre saturation point” (FSP). The
common definition according to which FSP is the state in which cell wall is
saturated with water and there is no free water in cell lumina is nice but says nothing
about the way, how the value of FSP can be measured. Another aspect of sorption
history is the “hygroelastic effect”. If we have three identical wood samples – one
Sorption isotherms of wood 75

loaded with mechanical compression stress, one stress-free and the last one under
tension stress, the respective equilibrium moisture content increases in this order. It
is quite clear, because sorption is accompanied by swelling, desorption by shrinkage.
So mechanical stresses can cause different equilibrium state from stress-free
conditions acting either in accordance with swelling or partially prevent it.

28

24

20 desorption
sorption
16
e [%][%]

II. desorption
EMC

12
X

0
0,0 0,2 0,4 0,6 0,8 1,0
relative humidity

Figure 4.4. Sorption hysteresis

Hygroelastic effect is in details described by Skaar (1988). If we admit that sorption


equilibrium is influenced by stresses it means also that the sorption or desorption
rate and stresses generated during the sorption process can influence the value of Xe.
From this point of view sorption equilibrium as described by sorption isotherm is
not thermodynamic equilibrium in the full sense and the term “sorption history”
really reflects the actual situation. Some other factors influencing sorption are
discussed below with the comparison of different sorption models.

4.4.3 Influence of temperature


As it was mentioned earlier it is necessary to distinguish between the long term and
a short term influence of temperature on wood Xe. Under the same conditions except
temperature the higher values of it lead to lower values of Xe and vice versa. This
influence is quantitatively described in sorption models. Exposure of wood to higher
temperature causes the decrease of its hygroscopicity. Salamon et al. (1975), Trnka
and Ladomerský (1985) confirmed this by the comparison of equilibrium moisture
content of high temperature, conventional and air - dried wood. Steaming also leads
to the decrease of wood hygroscopicity (Babiak and Németh 1998). The influence of
high temperature on the wood hygroscopicity can be explained by the hydrolysis
76 Fundamentals of wood drying

reaction in the degradation of hemicelluloses, which lowers the number of sorption


sites (Wengert 1979). Another explanation is a possible formation of hydrogen
bonds between pairs of free OH groups that decreases the number of sorption sites
(free OH groups). This effect is more pronounced in the combination of high
temperature with mechanical pressing e.g. plywood and particleboards. Irradiation
of wood can also bring about its lower Xe.

4.5 Sorption models

One of the first sorption models was Langmuir sorption isotherm. It is based on the
assumption that sorption takes place on a surface with isolated sorption sites, each of
them can bind a single water molecule. Dynamic sorption equilibrium is
characterised as the state when the number of molecules sorbed per unit of time is
the same as the number of molecules leaving the sorption surface in the same time.
The respective equation is

b⋅H
X = (4.3)
1+ b ⋅ H
where X is the portion of sorption sites that are occupied at the relative humidity of
the air H not expressed in %. The constant b is related to the free energy change F
of sorbed molecules by the equation:

 ∆F 
b = exp − 
 (4.4)
 R ⋅T 
Although the Langmuir model is too simple to describe the sorption of water by
wood properly it can be found as a part of more sophisticated models.

According to Skaar (1988) we find minimum 77 of them but even larger number can
be expected. Sorption theories can be divided into four main groups.
• The first group includes the “layer theories” that suppose that molecules of
water (or adsorbate) form layers on the large internal surface of adsorbent.
• The second group is based on the assumption that the system polymer-
water forms a solution in which two or more phases can exist in
equilibrium.
• The third group treats sorption as capillary condensation of wood as the
result of the drop of saturated vapour pressure over the curved surface of
water in capillaries. Sometimes the capillary condensation is considered as
a supplementing mechanism for other theories.
• The fourth group includes the theories based on various physical properties
as it is in case of polarisation theory.
Some theories combine various mechanisms.

One of the most known sorption isotherms is that described by Dent (1987). Its
equation is of the form:
Sorption isotherms of wood 77

X b0 ⋅ H
= (4.5)
Xm (1 − b ⋅ H ) ⋅ (1 + b0 ⋅ H − b ⋅ H )
where X denotes wood moisture content based on dry mass, Xm is the same quantity
in hypothetically completed monolayer and b0, b are constants related to the free
energy change of water molecules (Eq. 4.4) sorbed in monolayer and higher levels
respectively. The physical picture of this model is similar to Langmuir one. The
water molecules are sorbed in layers and dynamic equilibrium is assumed between
each two layers. The last equation can be also written in the form:

X
=
(b0 − b ) ⋅ H + b ⋅ H (4.6)
Xm 1 + (b0 − b ) ⋅ H 1 − b ⋅ H

The first term on the right site represents Langmuir isotherm or hypothetical
monolayer, the second term describes sorption in polymolecular layers. This
isotherm is a general model. It can be seen that if one puts b = 0, the Langmuir
isotherm is obtained, if b = 1, Equation (4.6) describes the well known BET
isotherm. The assumption b = 0 means that in the BET model the molecules in
polymolecular layer are bound to sorption sites with the same energy as it is in
liquid water. This energy is lower than that for molecules bound in monolayer.
Dent’s model, on the other hand, assumes that molecules in polylayers are bound
with the same energy for each layer but this energy is higher than that for liquid
water.

The Guggenheim-Anderson-deBoer (GAB) sorption isotherm is the same as Dent’s


one with different notations. It’s equation is of the form (Hartley 2000):

X c⋅k ⋅H
= (4.7)
Xm (1 − k ⋅ H )⋅ (1 + c ⋅ k ⋅ H − k ⋅ H )
It is clear that the two models are identical with the following relationships among
constants: b0 = k·c and b = k.

The “constants” b0, b, Xm in the Dent’s model can be computed via a linearisation of
equation (4.6) or by non-linear regression. According to the former, the expression
can be rewritten into the form:

H
= A+ B⋅H +C⋅H 2 (4.8)
X
The unknowns A, B, C can be computed as the solution of the system:
78 Fundamentals of wood drying

N N N
A⋅ N + B ⋅ ∑
i =1
Hi + C ⋅ ∑i =1
H i2 = ∑X
i =1
i

N N N N
A⋅ ∑H
i =1
i + B⋅ ∑H
i =1
i
2
+C⋅ ∑H = ∑X
i =1
3
i
i =1
i ⋅ Hi (4.9)
N N N N
A⋅ ∑H
i =1
i
2
+ B⋅ ∑H
i =1
3
i +C⋅ ∑H = ∑ X
i =1
i
4

i =1
i ⋅ H i2

and then the solution can be transformed to the constants b0, b, Xm using the
following equations:

B − B + B2 + 4 ⋅ A⋅C 1
b0 = 2 ⋅ b + b= Xm = (4.10)
A 2⋅ A A ⋅ b0

The second procedure uses a non-linear regression that can be computed in various
software packages. As MS Excel is a widely spread one, we used it for this purpose.
The computation procedure is illustrated in Table 4.5.

Table 4.5. Computation of Dent constants through nonlinear regression (The letters
A, B and numerals mean co-ordinates in MS Excel spreadsheet)

A B C D E F
1 H Experiment Theory
2 0.1 2.39 2.53 -0.14 b0 = 3.280
3 0.2 4.58 4.52 0.06 b= 0.680
4 0.3 6.52 6.28 0.24 Xm = 9.046
5 0.4 7.82 7.99 -0.17
6 0.5 9.65 9.78 -0.13
7 0.6 12.26 11.75 0.51
8 0.7 13.43 14.06 -0.63
9 0.8 16.86 16.92 -0.06
10 0.9 21.07 20.63 0.44
11 1 25.64 25.79 -0.15
12 SSD= 1.01

The values of relative humidity H are placed in A2-A11. The experimental values of
Xe are in B2 - B11. In the column C the theoretical values of X calculated with the
estimated values of b0, b, Xm can be found. The column D contains the differences
between theoretical and experimental values. The sum of squares of these
differences is in D12. The non-linear regression is computed using the tool “Solver”
that changes the estimated values F2-F4 in order to minimise D12. If the estimation
is not far from reality, the iteration process converges very fast. The program in MS
Excel is very simple and easy to use.
Sorption isotherms of wood 79

Based on the data of the USDA Forest Service (Annon 1987) we computed the
“constants” b0, b, Xm for each temperature level. These values were processed to
compute b0, b, Xm as functions of Celsius temperature t. We obtained the following
equations:

b0 = 3.3822 + 0.03492 ⋅ t − 0.00033 ⋅ t 2


b = 0.6689 + 0.001486 ⋅ t (4.11)
X m = 9.03 − 0.0523 ⋅ t

The shape of the Dent’s sorption isotherm is shown in Fig. 4.5. In this figure, the
typical monolayer form (MONO) of sorption isotherm can be seen. It approaches the
value Xm. The polymolecular term increases exponentially with increasing relative
humidity. The superposition of both provides the typical s-shaped sorption curve.

The term describing capillary condensation can be added to the original Dent’s
Equation (4.5). Simpson (1973) derived for this term the equation:

π ⋅L 2 ⋅ M v ⋅σ
X = (4.12)
m0 1 
R ⋅ T ⋅ ln 

H 
where L is the length of capillaries that are taken as cylindrical, m0 is the mass of
oven dry wood and is the surface tension of water and Mv is the molecular mass of
water.

24
MONO
POLY
TOTAL
moisture content X[%]

18

12

0
0,0 0,2 0,4 0,6 0,8 1,0
relative humidity

Figure 4.5. Typical s-shaped sorption isotherm


80 Fundamentals of wood drying

It should be noticed that Dent’s isotherm was derived under the assumption that the
adsorbent surface is homogenious and the lateral interactions of molecules are
negligible (Brunauer et al. 1967). This is clearly not the case of wood which swells
during sorption. The Dent model gives the possibility to calculate the specific
internal surface of wood. It is given by:
2/3
 6⋅M  N ⋅ Xm
S=  ⋅ (4.13)
 π ⋅ ρ ⋅ N  Mv

where S [m2·g-1] is the specific internal surface, Mv [kg·mol-1] is the molar mass of
water, N = 6.022·1023 mol-1 is Avogadro’s number, Xm is the hypothetical moisture
content for monolayer (Požgaj et al. 1993). The specific internal surface of wood is
high. Stamm (1964) measured this quantity for Sitka spruce at the temperature of
20ºC when the X for completed monolayer was 6.5% and got the value 254 m2·g-1.
The question of “sensitivity” of sorption isotherm constants arises when we would
like to compare different experimental results.

Another very popular sorption model in wood science is the Hailwood–Horrobin


(HH) sorption isotherm (Skaar 1988). This model assumes that a part of sorbed
water forms a hydrate with wood. The thermodynamic system then consists of three
parts – dry wood, hydrated wood and dissolved water – forming the ideal solution.
The equation of the isotherm is of the form:

X K1 K 2 H K2H
= + (4.14)
Xm 1 + K1 K 2 H 1− K2H

This equation can be linearised in the same manner as (4.6). It can be shown that
from the mathematical point of view Dent and HH models are identical and the
constants are bound by the relations

b0 = (K1 + 1) ⋅ K 2
(4.15)
b = K2

Based on the data of USDA Forest Service (Annon 1987) we computed the
“constants” as functions of Celsius temperature t. Using the same procedure as for
the Dent’s isotherm we obtained equations:

K1 = 4.6184 + 0.05302 t - 0.000544 t 2


K 2 = 0.6947 + 0.00215 t - 0.00000 t 2 (4.16)
2
X m = 9.11 - 0.059t + 0.000001 t
The constants K1, K2 are related to the free energy change of sorbed molecules in the
same manner as in (4.4) for the hydrated and dissolved water.
Sorption isotherms of wood 81

HH model has another point of interest. It introduces “mole of wood” as the mass of
wood that is in the hydrated state with one mole of water. This is to some extent an
analogy with monolayer in Dent’s model. Even the numerical values of both
constants Xm are identical. Because of identical mathematical model of Dent and HH
the calculation procedures as well as graphical representations are the same. There is
therefore no sense to expect that the mathematical fitting could bring different
results for these two isotherms. The question arises whether, based on sorption
isotherm results, one can decide what model is more reasonable. In case of these two
models the choice is the question of personal preference only.

Hunter (1996) proposed sorption isotherm based only on the concept of capillary
condensation. The crucial assumption of the model is the assumption that the
volume of wood cell wall during X decrease is reduced by the volume of released
water. This approach is very similar to the concept of nonconstant capillaries
introduced by Cudinov (1984). The equation of sorption isotherm is of the form:

X 1
= (4.17)
X FSP  ρ ⋅X   2 − H α / 2 ρ ⋅ X
1 + s FSP  ⋅   − s FSP
 ρ 
  H  ρ
where H is relative humidity, index FSP denotes fiber saturation point, s density of
wood substance and density of water in wood. The value of the coefficient is:

ρ ⋅ R ⋅T
α= (4.18)
G FSP

where GFSP is the rigidity of wood substance at FSP. The value of can be adjusted,
however according to the experimental values. The term “fibre saturation point” is
not exactly defined but from the context it is clear that the author means the
maximum moisture content of the cell wall. Through the relationship between G and
moisture content the author avoids common problems of capillary condensation
theories – the validity of surface tension concept for low moisture content. It is
interesting that this model has practically no degree of freedom because the only
constant can be calculated on the base of physical quantities. In spite of this the
agreement of theoretical and experimental data is surprisingly good.

A large group of sorption isotherms is based on polarisation theory. Bradley (as


cited by Skaar 1988) based on the dipole theory proposed the equation:

ln (1 / H ) = K 2 ⋅ K1X + K 3 (4.19)

where K1, K2, K3 are constants, the values of which depend on the intensity of
electric field properties of sorbed water end temperature. The last equation is not
convenient because of problems with derivatives. From this point of view better is
the equation:
82 Fundamentals of wood drying

 − 18 ⋅ A ⋅ B 
H = exp  ⋅ exp( A ⋅ X ) (4.20)
 R ⋅T 
proposed by Anderson and McCarthy (1963), where A and B are constants. One form
of sorption isotherm equation but different equations for constants were chosen by
several authors. Avramidis (1989) tested several sorption isotherms that can be
described by the equation:

(
1 − H = exp A ⋅ X B ) (4.21)

where the constants A, B are given in the Table 4.6.

Table 4.6. Values of constants according to Avramidis (1989) in Equation (4.21)

Model A B
Henderson -0.0001·T 1.46
Day-Nelson 0.34·10-16·T5.98 0.30·103·T-0.93
Zuritz et al. -0.13/T·(1-T/647.1)-6.46 0.11·103·T-0.75

It seems that the main motivation of the authors cited in Table 4.6 was to prescribe
directly the type of the relationship of A, B and T and by nonlinear regression to
avoid mistakes that occur in common procedure when the values of the coefficients
are computed and then their relationship with temperature is determined.

DeBoer and Zwicker (as cited by Požgaj et al. 1993) published the equation of
sorption isotherm in the form:

1 
ln  = A ⋅ exp(− B ⋅ X )
 (4.22)
H 
This isotherm was probably derived from the polarisation theory similarly to
Eq. (4.19). The constants A and B are functions of temperature. Based on the data of
USDA Forest Service (Annon 1987) we computed them as functions of Kelvin
temperature T:

A = 7.7317 − 0.014348 ⋅ T
(4.23)
B = 0.0087 + 0.000567 ⋅ T
Equation (4.23) can be also linearised by taking the logarithm of both sides and
calculating X as a function of H. In Dent’s sorption isotherm we could choose the
values of relative humidity either in % or in 100 times lower number. In this case
percentage expression should not be used because in the practical range of H ln(1/H)
would yield negative numbers and the linearisation procedure could not be used.

Kollmann (1963) proposed model that should have described three parts of sorption
in the form:
Sorption isotherms of wood 83

 (b H − 1)2   (b2 H − 1)2 


X = X 1 + X 2 + X 3 = a ⋅ H n + c1 exp − 1  + c 2 exp −  (4.24)
 2   2 
The first term should reflect adsorption, the second and the third describe
submicroscopic and microscopic capillary condensation.

As it was stated earlier, there are many sorption models in literature. It is beyond the
scope of this chapter to provide an exhausting review of all these models. In practice
the choice of a sorption isotherm is a question of personal preference.

4.6 Thermodynamics of sorption

The heat of evaporation of free water is given by Equation (4.2). Let us consider
wood that is in equilibrium with the surrounding atmosphere a given temperature
and H = p/ps. If we compute the heat of evaporation of bound water in analogy with
Equation (4.2) applied to the pressure p we obtain:

d (ln p s ) d (ln H )
hw = hw0 + ∆hw = R ⋅ T 2 + R ⋅T 2 (4.25)
d (T ) d (T )
It is seen here that the heat of evaporation of bound water is higher than that of free
water. The difference is called the differential heat of sorption. It is the heat needed
to evaporate 1 mol (or 1 kg) of bound water at given moisture content. This quantity
can be measured either thermodynamically from sorption isotherms or using a
calorimeter. In the latter case we measure the so-called heat of wetting given by the
equation:
X FSP

hz = ∫ ∆h
X
w ( X ) dX (4.26)

The samples conditioned to a certain X are soaked into liquid water in a calorimeter.
The heat evolved up to the full saturation is then measured. The relation between the
heat of wetting and the heat of sorption is:

dhz
∆hw = − (4.27)
dX
Coper and Ashpole (1959) described the heat of wetting as a function of moisture
content by the equation

hz = a ⋅ exp(− b ⋅ X ) (4.28)

This leads to the result that the heat of sorption is also exponentially decreasing with
moisture content

∆hw = A ⋅ exp(− B ⋅ X ) (4.29)


84 Fundamentals of wood drying

Skaar (1988) gathered the data on the experimentally obtained values of the
differential heat of sorption evaluated according to this equation. The values of A
varied from 16.090 kJ·mol-1 to 21.316 kJ·mol-1 and B from 0.1240 to 0.1886 [%-1].
Based on all these coefficients we computed the values of the differential heat of
sorption ∆hw for all the constants A and B in the interval 0-30% of X with the step
0.5%. From all these data we computed average value, maximum and minimum for
each value of X. In MS Excel we also computed the coefficients A, B for the average
values of ∆hw . By this procedure we obtained the equation:

∆hw = 18.461 ⋅ exp(− 0.1393 ⋅ X ) (4.30)

Differential heat of sorption is a good quantity for the testing of sorption isotherms.
According to Equation (4.25) ∆hw can be determined from sorption isotherms and
by calorimetry. The agreement of the two results can be a good criterion for the
comparison of several sorption models. The differential heat of sorption for deBoer–
Zwicker model (4.22) is given by the equation:

 dB dA  − BX
∆hw = R ⋅ T 2 ⋅  A ⋅ X − ⋅e (4.31)
 dT dT 

4.7 Testing of sorption theories

Simpson (1979) tested six sorption models. He came to the conclusion that the fit of
experimental data with the model was excellent for all the models except BET. The
average deviation was less than 0.5% of wood equilibrium moisture content. Much
worse results were obtained with the test of the heat of sorption. Values predicted by
the models were in error by 50% or more. Avramidis (1989) tested three sorption
models. He tested only the fit of experimental and theoretical data and all the results
were quite well. If we calculate the differential heat of sorption, the results are not so
good. In Figure 4.6 the results of ∆hw as functions of Xe are plotted. It is seen that
the highest values were obtained for calorimetric procedure. The closest to them are
the results obtained by Day-Nelson then equation (4.29) based on de Boer-Zwicker,
then Zuritz. Finally the Henderson’s model prediction of ∆hw is very poor.

The physical picture of different sorption models is almost the same. Part of the
water is intimately associated with OH groups, (monolayer, hydrated water) part is
less tightly bound (polylayers, dissolved, liquid). The good fit of experimental and
theoretical data does not in any case mean that the physical picture associated with
the sorption theory is correct.

If we are to compare different models we should take into consideration all the
advantages and drawbacks of each model. It can be generally stated that the fit of
experimental data with any model is good. The advantage of layer models (e.g.
Dent’s) is that the model works at the boundary values of relative humidity H = 0
Sorption isotherms of wood 85

and H = 1. Such models give the possibility to define and measure the
hygroscopicity limit HL as an extrapolated value of Xe content to full saturation.
This enables us to avoid great technical problems occurring at H close to saturation
where condensation occurs. The definition of the FSP as an equilibrium moisture
content at full saturation is actually misleading. At full saturation water vapour is in
thermodynamic equilibrium with liquid water. If wood is in thermodynamic
equilibrium with water vapour it must be inevitably in the saturation state in
thermodynamic equilibrium with liquid water. This is not new. Siau (1995)
discusses the problem of the FSP definitions in details. It seems that FSP is not a
sharp boundary between bound and free water. In building physics the retention
curve (what is actually sorption isotherm) shows a rapid increase similar to deBoer-
Zwicker isotherm. From the viewpoint of wood drying, however, these marginal
points are not so important. What is very important is the moisture content at which
wood starts to change its dimensions, because this generates the drying stresses.

14

Eq.(6.5)
12 DAY_NELSON
ZURITZ
HENDERSON
10
Eq(6.4)
QL [kJ.mole ]
-1

0
0 4 8 12 16 20 24 28
EMC [%]

Figure 4.6. Differential heat of sorption according to different models

4.8 Conclusions

Sorption of water by wood is a very important process from the viewpoint of wood
drying. The relationship among equilibrium moisture content, relative humidity and
temperature is a base for the control of the drying process. Equilibrium moisture
content is mainly influenced by relative humidity, temperature then by species and
sorption history. In literature we can find contradicting results of the influence of
different factors. A database that would collect and sort all the results is required. Up
86 Fundamentals of wood drying

till now the most exhaustive data are given by USDA Forest Service in spite of the
fact that they were determined for Sitka spruce.

Equilibrium moisture content as a function of relative humidity is usually described


by the system of sorption isotherms. Sorption models are usually based on some
physical picture of sorbed molecules or in some cases the relationship is only
mathematical (polynomial expansion). The physical picture of sorption can be
hardly confirmed, because the good agreement of theoretical and experimental
results is not an exact proof of the validity of the model. Almost all sorption models
describe the equilibrium moisture content in the practical range equally well. Bigger
differences can be seen in boundary points. The choice of sorption model is still a
matter of personal preferences. One should carefully choose what the respective
model should reliably predict and according to it the choice of sorption model
should be done. The state of standardisation in sorption measurements is very low
and the exact standards are strongly needed.
5 Shrinkage, swelling and warp caused
by moisture changes
Knut M. Sandland* and Peder Gjerdrum**
*Norwegian Institute of Wood Technology
Forskningsveien 3B
N-0371 Oslo
knut.sandland@treteknisk.no

**Norwegian Forest Research Institute


Høgskoleveien 12
N-1432 Ås
peder.gjerdrum@skogforsk.no

Nicholas Clarke revised the English text

ABSTRACT

When the moisture content in wood changes in the hygroscopic range, shrinkage and swelling
will occur during desorption and adsorption, respectively. It is very important to take the
dimensional movement of wood in relation to the moisture content into consideration when
producing various wood products. The moisture content in the products has to be at a level
that minimizes the dimensional movement when the wood is exposed to variable climate.
The shrinkage/swelling potential is different in the three main directions of the wood, with
higher values in the tangential direction compared to the radial direction. Several theories
have been proposed to explain this difference. In the longitudinal direction the potential is
very low compared to the transverse directions.
Because of these differences, and because of different shrinkage potential for various wood
properties (e.g. juvenile wood, compression wood, low/high density wood.), deformations will
occur during drying. In the cross-section traditional cup will occur as a result of the
difference between radial and tangential shrinkage. In the longitudinal direction twist will
occur if the grain angle deviates from the longitudinal direction of the timber, while spring
and bow will occur if the shrinkage potential is variable within the wood pieces. When the
timber is loaded during drying, the deformations in the longitudinal direction, especially the
twist, are considerably reduced.
Several approaches to reduce the dimensional movement in wood have been tried. By various
chemical treatments it is possible to obtain wood where the dimensional movement is
considerably reduced or almost eliminated. Another method is heat treatment, which reduces
the hygroscopicity and decreases the dimensional movement in wood.

KEYWORDS: anisotropy, bow, cup, deformations, dimensional movement, shrinkage, spring,


swelling, twist

Fundamentals of wood drying, Patrick Perré editor, pages 87 to 104.


88 Fundamentals of wood drying

5.1 Introduction

The shrinkage and swelling properties of wood are some of the most important
properties that have to be taken into consideration when using wood for various
purposes. Nowadays the competition from other building materials than wood gets
stronger all the time. These producers have no inhibitions in marketing their own
materials as superior compared to wood, not by focusing on the possible advantages
of their own products, but by focusing on the possible disadvantages of wood. In
particular, shrinkage/dimensional movements and different kinds of deformations in
wood are used as an argument. Therefore, it is important to produce building
materials of wood where these disadvantages are reduced to a minimum to maintain,
or improve, the market share for wood based products.

When the moisture content in wood changes in the hygroscopic range, shrinkage and
swelling will occur during desorption and adsorption, respectively. It is very
important to take into consideration the dimensional movement of wood in relation
to the moisture content when producing various wood products. The moisture
content in the products has to be at a level that minimizes the dimensional
movement when the wood is exposed to the variable climate existing in the
surroundings.

Shrinkage occurs in wood when the water molecules leave the molecular structure in
the cell wall. The wood has an anisotropic character, which results in different
shrinkage/swelling potential in the three main directions of the wood. Various wood
properties can in addition contribute to various deformations in wood when it is
dried below the fibre saturation point.

In this chapter, the most important aspects of shrinkage and swelling of wood are
described. As an introduction, the definitions of the properties are given together
with explanations on how to measure them. Further, explanations are given
concerning the differences in shrinkage/swelling in the three main directions of
wood, and how the shrinkage and swelling vary for various wood properties.
Deformations that might occur during wood drying are described, and a short review
concerning dimensional stabilisation of wood is given.

5.2 Definitions

Shrinkage and swelling are important physical properties of wood, and are a
reduction and increase in the wood dimensions, respectively, when the moisture
content is changed below the fibre saturation point. Changes in moisture content
above the fibre saturation point give no changes in the wood dimensions. However,
when the wood is dried from green condition, various stresses in the wood may be
so high that the cell walls cannot resist them (e.g. capillary tension forces). Then the
result will be collapse in the cell wall structure, with dimensional changes in the
wood as a result, even when the moisture content is above the fibre saturation point.
Shrinkage, swelling and warp caused by moisture changes 89

However, these dimensional changes are not traditional shrinkage, and will not be
discussed further in this chapter.

Shrinkage and swelling in wood have an anisotropic character, meaning that the
magnitude is different in the three main directions of wood. The values are higher in
the tangential direction than in the radial, while they are considerably lower in the
longitudinal direction.

The calculation of shrinkage is based on green dimensions (moisture content above


the fibre saturation point), while the calculation of swelling is based on totally dry
dimensions. The values are expressed as changes in dimensions according to the
original dimension (wet or dry), either in absolute values or in percentage. The
percentage is usually used when shrinkage and swelling values are given. The
shrinkage and swelling can be calculated from Equations (5.1) and (5.2).

l1 − l 2
shrinkage [%] = ⋅100 (5.1)
l1

l1 − l 2
swelling [%] = ⋅100 (5.2)
l2

l1 = dimension at moisture content above the fibre saturation point


l2 = dimension at absolute dry wood

As indicated by Equations (5.1) and (5.2), the shrinkage and the swelling have
different values, even though they are based on the same values for wood
dimensions. One of them can, however, be calculated if the other is known, and the
relationship between shrinkage and swelling is shown in Equation (5.3).

l1 − l 2 = shrinkage ⋅ l1 = swelling ⋅ l 2 (5.3)

During sorption the geometric shape of any wooden plank takes on certain changes.
According to the European norm (EN 844) the term 'warp' is used for any distortion
of a piece of timber. Warp is specified as bow, spring, cup or twist (see Fig. 5.6 and
Appendix).

5.3 How to measure shrinkage/swelling and warp?

Rules for measuring shrinkage and swelling are given in international standards such
as ISO 4858 (1982a): Wood – determination of volumetric shrinkage, ISO 4469
(1981): Wood – determination of radial and tangential shrinkage, ISO 4860 (1982b):
Wood – determination of volumetric swelling and ISO 4859 (1982c): Wood –
determination of radial and tangential swelling.

To determine shrinkage and swelling, a micrometer or a slide calliper with high


accuracy can be used to measure the dimensional changes in the different directions
90 Fundamentals of wood drying

of the wood. To measure the changes in volume, the measurements in the three
different directions can be combined. An alternative would be to use the immersion
method to measure the volume of the wood piece. This method is useful if the wood
piece has a shape that makes dimensional measurements difficult. It is, however,
important to be aware that this method can be slightly inaccurate, due to the fact that
some liquid might be absorbed in the wood during the immersion period. This is a
minor problem for green wood, but can be a problem for dry wood.

Rules for measuring warp are described in European norms (EN 1310). Bow, spring
and twist are measured as the maximum deviation over a 2 m length of the plank.
Cup is measured as the maximum deviation within the plank's width. The
requirements for twist and cup are dependent on the plank width. Some examples of
requirements for allowed warp are given in the Appendix.

5.4 Shrinkage and swelling

5.4.1 Wood – water relationships

In green condition the wood cell walls are usually saturated with water. When the
wood dries below the fibre saturation point, the wood starts to shrink, and will
continue to shrink until the moisture content in the piece is zero. Shrinkage and
swelling are completely reversible processes in small pieces of stress-free wood
(Haygreen and Bowyer 1996), except for the first desorption process. However, in
large pieces of solid wood, swelling or shrinkage may not be completely reversible
because of internal drying stresses.

Cellulose is one of the main components of the wood cell walls (43-52% of the dry
wood substance). The cellulose in the cell wall structure is highly structured, and
exists usually as elementary fibrils that are like threads. In the elementary fibrils the
cellulose molecules are organized in alternating crystalline and amorphous regions.
The crystalline areas are hydrophobic, which means that the molecular forces
between the cellulose molecules are very strong. Therefore, water adsorption only
takes place in the amorphous regions, not in the crystalline areas.

Several elementary fibrils (20-60) make a microfibril, about 10-30 m in diameter.


Several microfibrils (20-50) make a macrofibril. The macrofibrils make the lamellae
that build up the cell wall structure. In the interfibrillar spaces between the
elementary, micro and macrofibrils there is a meshy system that is partially filled
with amorphous materials such as lignin, hemicellulose and pectins. Various
extractives in the heartwood can also be present here. In these amorphous regions,
and at the surfaces of the crystalline regions, the water molecules are in the form of
bound water in the cell walls. During adsorption or desorption of water in these
regions, the size of the wood material will change, and the wood will then swell or
shrink, respectively.
Shrinkage, swelling and warp caused by moisture changes 91

The amount of shrinkage that occurs when water is removed from unconstrained
wood below the fibre saturation point is almost proportional to the amount of water
removed from the cell wall, and it has been assumed that the volumetric decrease is
linearly related with moisture changes in the hygroscopic range as illustrated in Fig.
5.1.
12
11
10 Volume
9
8
Shrinkage [%]

7
Tangential
6
5
4
Radial
3
2
1 Longitudinal
0
0 FSP
Moisture content [%]

Figure 5.1. The relationship between moisture content and shrinkage is almost
linear (typical shrinkage values for wood of Norway spruce (Picea abies)). The
analogous situation exists for the relationship between swelling and moisture
content

Skaar (1988) indicates that this is a convenient approximation, but is not usually
observed when careful measurements are made. It is referred to by Keylwerth
(1964), who made such measurements on European birch from the dry condition
through re-adsorption to cell-wall saturation. In these investigations it was found
that the curves are slightly sigmoid in all three main directions in wood. Such curves
are linear only over a portion of the hygroscopic moisture range, as shown to the
right in Fig. 5.2. The diagram to the left in the same figure illustrates the derivative
of the hygroexpansion in relation to the moisture content, which gives the sigmoid
shape of the swelling curves.

Figure 5.2. Rate of swelling at different moisture contents (to the left), and curves
showing the relationships between dimensions and moisture content (to the right)
92 Fundamentals of wood drying

Fig. 5.2 shows that at low moisture contents (4-5%) the rate of swelling is relatively
small. Skaar (1988) explains this phenomenon by the fact that the first water
molecules sorbed find their way into interstices in the cell wall matrix and do not
contribute their full volumes to the cell wall. In the moisture content region from
about 5% to about 25%, the volume increases almost linearly with increasing
moisture content. From about 25% moisture content to the fibre saturation point,
there is a decrease in the rate of swelling, which may be caused by capillary
condensation in small voids in the wood structure.

When measuring shrinkage and swelling, another phenomenon will also result in
lack of linearity between changes in dimension and moisture content. In a piece of
wood, moisture gradients will occur when the moisture content is changing, which
itself results in non-linearity for the relationship when the piece is considered as a
whole. Another result of such moisture gradients is that they induce internal stress
situations within the pieces, which will partially prevent the free, unconstrained
dimensional changes due to changes in moisture content.

However, for most practical calculations it is sufficient to assume that the


relationship between dimensional change and moisture content is linear over the
whole range. This makes it easy to estimate shrinkage between any two moisture
contents. If the green to oven dry shrinkage values for the wood are known together
with the level of the fibre saturation point (in most situations an assumption of the
fibre saturation point at 30% moisture content is reasonable), dimensional change
for a change in moisture content can be calculated using Equation 5.4 (for moisture
contents below the fibre saturation point):

β ⋅ Xd
βd = (5.4)
X FSP

βd - shrinkage that corresponds to the change Xd in moisture content


β - total shrinkage potential
Xd - change in moisture content
XFSP - moisture content at fibre saturation point
(both shrinkage and moisture content in %)

For swelling the calculations will be analogous.

5.4.2 The anisotropic character of shrinkage and swelling in wood

Shrinkage and swelling differ considerably between the three main directions of
wood. The low longitudinal shrinkage compared to the cross sectional shrinkage has
to be attributed to the cell wall composition. The secondary cell wall layer consists
of the three layers S1, S2 and S3. The S1 and S3 layers are relatively thin compared to
the S2 layer, and have consequently less influence on the shrinkage/swelling
compared with the S2 layer.
Shrinkage, swelling and warp caused by moisture changes 93

The microfibrils are more uniformly oriented in the S2 layer compared to the two
other layers. In the S2 layer the longitudinal direction of the fibrils is oriented almost
parallel to the longitudinal direction of the cells (a deviation of about 10-15° in
normal mature wood). The shrinkage/swelling is caused by a decrease or increase in
the amount of water molecules inside and between the fibrils. This will result mainly
in dimensional changes in the cross sectional direction of the cell wall. Because of
the above-mentioned deviation between the fibril orientation in the S2 layer and the
longitudinal direction of the cells, some longitudinal shrinkage/swelling will occur
in this cell wall layer too. In special wood types such as reaction wood and juvenile
wood, the fibril angle in the S2 layer can be considerable, resulting in higher
longitudinal shrinkage/swelling. In fact, in some cases values above 5% for
longitudinal shrinkage in such wood have been reported.

Various theories have been developed to explain the difference between radial and
tangential shrinkage. One theory is that the rays reduce the dimensional movement
in the radial direction, because the longitudinal direction of the ray cells is oriented
in the radial direction of wood. However, in various investigations different results
have been found concerning the influence of rays. Some investigations of
hardwoods have found that the shrinkage/swelling in the radial direction has
increased when the rays are excluded from the wood (e.g. McIntosh 1955). Skaar
(1988) says that such data clearly indicate that the ray restraint mechanism is at least
one of the factors responsible for transverse shrinkage anisotropy for these
hardwoods.

Another theory is that the higher amounts of bordered pits at the radial cell walls
compared to the tangential cell walls result in a higher degree of irregularity in the
fibril structure, especially in the S2 layer. This will influence the shrinkage, because
in the radial cell walls the fibril angle is on average higher than in the tangential cell
walls. Kifetew (1997) indicates, with reference to investigations performed by
Pentoney (1953), that the difference in microfibril angle between radial and
tangential walls may contribute to the differential shrinkage, but it is not an
important factor.

A third theory is based on differences in the swelling characteristics and


arrangements of the various cell wall layers. Among other factors this is related to
properties of the middle lamellae. In the investigations of Frey-Wyssling (1940a, b)
it is proposed that the difference in shrinkage in the middle lamellae between the
tangential and radial directions contributes to the shrinkage anisotropy. Skaar (1988)
says that this theory has not been confirmed and refers to other authors (e.g. Crews
1965) who have concluded that little shrinkage occurs in the middle lamellae and
that most shrinkage is confined to the secondary walls.

Skaar (1988) refers to various investigations (e.g. Bosshard 1956) where it is found
that there is a higher lignin content in radial than in tangential walls of longitudinal
wood cells. The lignin will reduce the hygroexpansion of wood, and therefore this
94 Fundamentals of wood drying

factor will contribute to the difference between shrinkage in the two transverse
directions of the wood. It is, however, difficult to say to what extent this property
contributes to explaining the difference in shrinkage between the radial and
tangential directions.

The last theory that will be mentioned here is the earlywood-latewood interaction
theory. According to Skaar (1988) the tangential expansion is greater than the radial
because of the alternation of earlywood and latewood zones in the radial direction in
many woods grown in the temperate zone, particularly softwoods. Hygroexpansion
of latewood is greater than that of earlywood (due to the density in the wood), and
the latewood is also stronger than the earlywood. In the tangential direction, the
bands of latewood will influence the shrinkage in the earlywood bands, while in the
radial direction, the shrinkage can almost be considered as a sum of the earlywood
and latewood hygroexpansion.

It should be noted that all models for anisotropy imply a slight deformation of each
cell and a rearrangement of the cell localisation in the matrix of cells building any
piece of wood. Skaar (1988) concludes that it seems apparent that several
mechanisms may operate to cause transverse anisotropy in wood shrinkage and that
these are not necessarily identical for all woods. He further says that considerable
controversy exists concerning the relative importance of the various mechanisms
that have been proposed.

5.4.3 Shrinkage/swelling for various wood species

Some intrinsic distinctions exist in shrinkage properties between species. In general,


tangential shrinkage is approximately twice the magnitude of the radial shrinkage.
Also, volumetric shrinkage tends to increase with the density, but there is a wide
variation in this general pattern. Table 5.1 shows shrinkage coefficients for a set of
species. The intervals given in the table must be considered as an indication of the
level of the mean values, which means that simple values for the coefficients can be
outside the indicated levels. This means that the variation in values for shrinkage
and swelling within the same species is considerable.
Shrinkage, swelling and warp caused by moisture changes 95

Table 5.1. Shrinkage coefficients for a set of important species (Wagenführ 1996)

Dry dens. Shrinkage coefficient [%] Anisotropy


Species [kg/m3] tang. t rad. r long. l t/ r
Gymnosperms
Fir (Abies alba) 410 7.2...7.6 2.9...3.8 0.1 2.0
Larch (Larix decidua) 550 7.8...10.4 3.3...4.3 0.3 2.4
Pine (Pinus sylvestris) 490 7.5...8.7 3.3...4.5 0.2...0.4 2.1
Spruce (Picea abies) 430 7.8...8.0 3.5...3.7 0.3 2.2

Angiosperms
Alder (Alnus glutinosa) 510 7.7...9.3 4.4...4.8 0.5 1.8
Ash (Fraxinus excelsior) 650 8.0...8.4 4.6...5.0 0.2 1.7
Aspen (Populus tremula) 450 6.7...8.5 ~3.5 - 2.2
Beech (Fagus sylvatica) 680 ~11.8 ~5.8 0.3 2.0
Birch (Betula verrucosa) 610 ~7.8 ~5.3 0.6 1.5
Elm (Ulmus glabra) 640 6.9...8.3 4.6...4.8 0.3 1.6
Lime (Tilia cordata) 490 9.1...10.7 5.5...6.6 0.2...0.3 1.6
Locust (Robinia
740 5.4...7.2 3.2...4.6 0.1 1.6
pseudoacacia)
Maple (Acer
590 ~8.0 ~3.0 0.5 2.7
pseudoplatanus)
Oak (Quercus robur) 650 7.8...10.0 4.0...4.6 0.4 2.1
Willow (Salix alba) 330 5.4...8.7 1.9...3.9 0.5...1.1 2.4

5.4.4 The influence of various wood properties

The behaviour of wood during drying is directly linked to the structure of the cell
walls and the matrix of fibres. Most important for the mechanical properties are the
S2 layer of the cell wall and the wood density (Persson 1997).

The thickness of the S2 layer varies between 30 fibrils in earlywood and 150 fibrils
in latewood (Wagenführ 1999). However, even if the two show different shrinkage
behaviour in pure samples, the wood will behave like a homogenous matrix of cells
interacting with each other (and with the rays) in full size samples (Persson 1997).

Density, grain angle and the fibril angle differ in juvenile and mature wood and,
hence, so do shrinkage and distortion properties. The higher fibril angle in juvenile
wood induces a higher value for the longitudinal shrinkage coefficient, and the
predominant left-handed grain angle induces one-directional twist.

Compression wood in softwoods differs considerably from regular wood in having


higher density, shorter cells and a higher fibril angle. Compression wood shows
higher volumetric shrinkage and considerably higher longitudinal shrinkage as
compared to normal wood. Planks partly containing compression wood might be
prone to severe distortions during drying.
96 Fundamentals of wood drying

With respect to the influence of density, the shrinkage/swelling will increase with
increasing density in the radial and tangential directions, while the opposite situation
exists in the longitudinal direction (within the same species). The increase in the
share of the S2 layer in the cell walls when the density increases can explain these
relationships. The fibrils in the S2 layer have a very uniform orientation, and the
deviation between the longitudinal direction of the fibrils and the longitudinal
direction of the cells is small. Therefore, when the density increases, the share of
wood substance with high transverse shrinkage and low longitudinal shrinkage will
increase.

5.4.5 Shrinkage and deformations during drying of full size timber

It is well known that deformations occur in sawn timber when dried to moisture
contents below the fibre saturation point. To understand the mechanisms behind the
different kinds of deformation during and after wood drying, many aspects have to
be considered.

Firstly, the shrinkage potential varies according to the three main directions in the
wood. Likewise, the shrinkage potential is different for the various wood types, e.g.
low-density wood, high-density wood, compression wood and juvenile wood.

Secondly, changes in moisture content in the hygroscopic range result in internal


stresses in wood. The wood is able to undergo creep deformations when it is stressed
and, especially if the wood is cleaved or split after this process, the permanent creep
in the wood results in different kinds of deformations.

The deformations in the first and second instances have different characteristics, and
in most cases it is possible to detect the main reason for them. However, the
development of deformations is an interaction between different kinds of
mechanisms. In the prediction of deformations in wood it is therefore important to
have a complete understanding of how these mechanisms co-operate.

Geometric deformation and shrinkage have great impact on the yield of wood during
splitting and planing, thus substantially affecting the economic efficiency of timber.
Also, distortion in commercial timber, especially twist, is one of the main reasons
for degrading or even rejecting planks for the intended use.

Distortion might partly be predicted in analytical or regression models, applying


moisture content and wood properties as predictors (Ormarsson 1995). However,
approximately half the variance remains unexplained, and thus the description of the
stochastic nature of these traits is equally important.
Shrinkage, swelling and warp caused by moisture changes 97

In full size planks the overall shrinkage and deformation are influenced by several
factors, e.g.:

• Moisture content
• Fibre structure and anisotropy
• Knots, juvenile wood and reaction wood
• Dynamic moisture fluctuation and gradients
• Drying stresses in wood as a partly elastic, partly plastic tissue

This complicates the analyses and modelling. Consequently, most reports are found
concerning work on small, clear samples.

a) Shrinkage in full size timber

Figure 5.3 is based on kiln drying of commercial, fresh cut centre yield battens of
Norway spruce.

10
9 Thickness (edge)
8 Width (outside face)
Estimated SHR
7
Shrinkage, %

6
5
4
3
2
1
0
60 54 48 42 36 30 24 18 12 6 0
Average moisture content, %

Figure 5.3. Empirical shrinkage in full sized commercially kiln dried Norway
spruce timber (Gjerdrum 2000a)

The shrinkage starts once the outer layer reaches the fibre saturation point (FSP),
which in this example corresponds to an average moisture content (MC) of 36%,
however, with substantial variation in MC. The shrinkage proceeds in an
accelerating manner until absolutely dry, distinct from the almost linear MC-
shrinkage relationship for clear wood without moisture gradients within the pieces
given in Figs. 5.1 and 5.2. Despite a most pronounced difference between radial and
tangential shrinkage in small, clear specimens, only small divergences might be
found between the outside face (mostly tangential ring pattern) and the edge (close
98 Fundamentals of wood drying

to radial pattern) of the plank. While the tangential shrinkage is approximately 8%


and the radial 4%, the empirical shrinkage for kiln dried, full size timber takes an
intermediate magnitude of 6.3% when dried to absolutely dry wood. The model is
given in Equations (5.5) and (5.6), for the range of average MC < 36%.
2
 average MC 
estimated shrinkage = 6.3 ⋅ 1 −  (5.5)
 36 

std.dev. (shrinkage) = 0.19 ⋅ estimated shrinkage (5.6)

b) Twist, bow, spring and cupping

The amount of cup in the planks depends on the distance from the pith, plank
thickness and width, moisture content and the degree of shrinkage anisotropy.
Fig. 5.4 shows results from an investigation (Tronstad 1971) where the cup was
measured in 50 mm thick planks of Norway spruce with variable thickness and
moisture content (the planks were taken near the pith, 2 x log, unrestrained during
drying).

7
U = 0%
6
50 mm thickness
2 x log U = 5,5%
5
Cup [mm]

4
U = 15,5%
3

2
U =26,5%
1

0
50 100 150 200 250
Width [mm]

Figure 5.4. The relationships between cup, plank thickness and moisture content for
planks of Norway spruce of 50 mm thickness (2 x log, unrestrained during drying)
(Tronstad 1971)

The extent of cup is influenced by the orientation of the annual rings in the cross
section (Fig. 5.5). However, the most common sawing pattern is to take out the main
yield from the log, and then cleave it into planks and boards with an annual ring
orientation similar to the plank shown in Fig. 5.4, where the cup decreases with
Shrinkage, swelling and warp caused by moisture changes 99

increasing distance to the pith. In this case, cup will occur, more or less, in almost all
the wood pieces.

An alternative is to orientate the annual rings in the vertical direction in the sawn
timber. Then the cupping will be located at the edge side, and will then be
insignificant. Sandberg (1998) describes different properties of sawn timber with
vertical annual rings. Several other investigations have been done to describe the
amount of cupping, both according to wood properties and sawing pattern.
Ormarsson et al. (1996) have, for example, modelled the influence of the annual ring
orientation and the distance from the pith on the cupping in sawn timber.

Figure 5.5. Deformations in the transverse direction attributed to the anisotropy of


tangential and radial shrinkage

Cupping of centre planks has been analysed by Morén and Sehlstedt-Persson (1992).
Cupping during stress free drying was well predicted by an analytically derived
model based on ring curvature and radial shrinkage coefficient. No influence of
wood density, ring width, late wood content or heartwood content could be verified.
Morén (1993) concludes that drying induced stresses affect cupping (reduction),
however only to a minor extent. Further, it is concluded that a conditioning process
at the end of the drying increases the cup because of the creep reversal in the outer
layer.

A reduction in cupping will be attained if the sawn timber is restrained from cupping
during drying. For the system to be efficient, the whole length of the planks has to
be in contact with the load, similar to in press drying. However, in the sawmills the
sawn timber is stacked with sticks between each layer, often 7-13 stickers in the
longitudinal direction. It is assumed that a top load of the stack has minor influence
on the cupping between the stickers. Some experiments done by Sandland and
Tronstad (2001) show that the cupping is to some extent also reduced between the
stickers, but not more than 20-30 cm from them, when the stacks are loaded during
drying.
100 Fundamentals of wood drying

In the longitudinal direction different deformations such as bow, spring and twist
appear when drying wood (Fig. 5.6). Spiral grain is the main reason for twist
deformations, and the juvenile wood is therefore highly exposed to these. Crook and
bow deformations are results of different longitudinal shrinkage potential in the
sawn timber. Compression wood and juvenile wood have for example considerably
higher longitudinal shrinkage potential than normal mature wood. When such wood
is present in a zone of a plank, it will bow and crook during drying.

Figure 5.6. Deformations in the longitudinal direction

The deformations in the longitudinal direction seem to be more or less independent


of the drying schedule, as long as the sawn timber is not restrained or loaded during
drying, as shown in Fig. 5.7. However, when the sawn timber is loaded during
drying, the deformations in the longitudinal direction, especially the twist, are
considerably reduced, as shown in Fig. 5.8. Drying at high temperatures improves
this effect.
Shrinkage, swelling and warp caused by moisture changes 101

Figure 5.7. Twist is particularly pronounced in the top layer planks, where no loads
are applied: small centreboards containing a dominant amount of juvenile wood

45
Non-loaded
40 Loaded, 1350 kg/m²
35

30
Frequency [%]

25

20

15

10

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32
Twist [mm/2 m length, plank width=150 mm]

Figure 5.8. An example showing that the amount of twist is considerably reduced
when the timber is loaded during drying (timber of Norway spruce with dimensions
of 50 mm x 150 mm (Sandland and Tronstad 2001))

Modelling of deformations such as twist, bow, spring and cupping has taken several
different approaches. Due to the complexity, statistical evaluation has been most
widely used. Ormarsson (1995) was able to analyse twist applying the finite element
method. This model has shown its ability to predict twist as a function of spiral grain
and distance from the pith, combined with parameters for wood properties such as
shrinkage coefficients, mechanosorptive creep and elastic properties (Dahlblom et
al. 2001). Twist in the range of -15° to +10° has been fairly well predicted.

A physical explanation of the moisture related changes in twist is given by


Johansson et al. (2001). Growth ring curvature and grain angle were shown to be the
most important factors for predicting twist. Severe twist was located in planks sawn
close to the pith, correlating to the amount of juvenile wood. No influence of
102 Fundamentals of wood drying

density, ring width or shrinkage could be found. The distortion was reversible
throughout several moisture changes.

For industrially kiln dried softwood timber, Gjerdrum (1997) provides some
information. 467 planks of Scandinavian spruce and pine were randomly sampled at
six different sawmills through a period of eight years. The planks represented 40 lots
of different dimensions and varying kiln drying practice. Observations of warp were
made after drying to a moisture content of 12-18% and moisture equalising
(Table 5.2).
Table 5.2. Warp in full size industrial spruce and pine timber (Gjerdrum 1997)

Trait Unit Average Std. dev. Distribution


value
mm/100 mm plank width ·
Twist 2.6 2.5 Skew/normal
2 m plank length
Bow mm/2 m plank length 2.6 1.9 Skew/normal
Spring mm/2 m plank length 2.0 1.4 Normal

No substantial correlation could be verified between the various elements of warp,


or between warp and certain wood parameters such as ring width or species.
Nonetheless, both the magnitude and the variation in warp might be influenced by
different sawmill and kiln drying practices. For example, spring might be introduced
both in the conversion process and during drying, and twist might be counteracted
by using loads during drying. However, the magnitude and the close to normal
distribution of the stochastic elements of the warp should be indicative for
industrially produced Nordic softwood timber.

c) Economic aspects of shrinkage and warp

As regards the question of economics, timber drying is a trade-off between increased


revenue for better drying quality (usually equivalent to lower moisture content) and
the additional cost for this improved drying. The cost may be separated into two
parts: I) the cost of operating the drying plant and II) the cost related to loss of
dimension (shrinkage), and increased distortion and degrading. In this chapter only
the latter cost will be briefly dealt with.

Shrinkage, cupping, twisting, degrading etc. all increase with drying to lower
moisture content. Thus, the timber might be less suitable for some planned uses and
the yield when resawing and planing will be reduced. In the trade, timber might be
offered in dimensions linked to some nominal moisture content, say 18%, or
dimensions have to be met at actual moisture. Also, further drying introduces more
variation in the shrinkage (Equation 5.6). Cupping, warp and degrading cause
additional costs, inevitably increasing for obtaining drier timber. According to
Gjerdrum (2000b), the 'dimension and quality cost' add up to a magnitude twice to
thrice the cost for operating the drying process. It is important, therefore, to use
precise and relevant estimates for all these cost elements.
Shrinkage, swelling and warp caused by moisture changes 103

5.4.6 Dimensional stabilisation by modification

The ability to bind to water molecules is an inherent property of all wood. Shrinkage
and swelling are directly linked to this property. For some applications it is highly
desirable to reduce the volumetric changes. This might be obtained by reducing the
hygroscopicity. In addition, reduced moisture content might add to the resistance
against biodegradation. Several methods have been proposed, and a few shall be
briefly mentioned.

One method is to add some combined hydrophilic-hydrophobic chemical, such as


acetyl (Rowell et al. 1993). While the polar part of the molecule will bind to the
wood, the hydrophobic part will block the connection between wood and water.
Acetyl is one example of such a dual property molecule. In Europe and Japan, some
early industrial plants for acetylated wood have been reported.

Another method is based on partial modification of the wood at elevated


temperatures. Several industrial applications have been reported in Europe. The
wood is heated to approximately 180-210°C under inert conditions. After treatment,
the equilibrium moisture content, and thus shrinkage and swelling, is substantially
reduced. Unfortunately, the timber strength is also reduced. Also, the wood takes on
a darker brown colour compared to natural wood.

Tjeerdsma et al. (1998) have a summary of the literature concerning heat treatment
of wood, and Hillis (1984) has reviewed the literature about hygroscopicity and
stability. Reductions of equilibrium moisture content (EMC) and shrinkage and
swelling seem to depend on treatment duration as well as treatment temperature. The
moisture content during the treatment also influences the effect. Kiln drying at high
humidity reduces hygroscopicity more than drying at low humidity, and higher
temperatures are needed to achieve similar results on dry wood.

The effect of heat treatment has been explained in different ways. One of the first
hypotheses trying to explain the stabilisation of wood by heat treatment was a cross-
linking reaction in which water is eliminated between hydroxyl groups on two
adjacent cellulose chains, with the formation of ether linkages (Stamm and Hansen
1937, references in Seborg et al. 1953).

Mitchell et al. (1953) proposed that the improved stabilisation obtained by heating is
due to the decomposition of hemicellulose and other carbohydrates. Inoue et al.
(1993) have a short literature review and conclude that the stabilisation of wood by
decreasing its hygroscopicity is caused by degrading the hemicellulose. Sehlstedt-
Persson (1995) has found a lower content of monosaccharides in high-temperature
dried (HT-dried) wood compared to low-temperature dried (LT-dried) wood.
Monosaccharides build up the hemicellulose fraction in wood. The influence of
hemicellulose is confirmed by Burmester and Wille (1975), who examined the
content of hydrolysable hemicellulose in teakwood (Tectona gandis L.F.). This
104 Fundamentals of wood drying

species is known for low swelling and shrinkage, and it was found that the
hemicellulose content is very low in teak wood.

Even if acetylated wood and thermally treated wood are not yet extensively used, the
application of such methods to wood seems promising.

Appendix
Definitions according to EN 844 Round and sawn timber - Terminology; numbers
are reference. The terms are equivalently defined in national standards and
summarised for some important languages in CTBA (1995).
Warp 3.26 Distortion of a piece of timber in the process of conversion,
and/or drying and/or storage
Bow 3.26.1 Lengthwise curvature of a piece of timber normal to the face
Spring 3.26.2 Lengthwise curvature of a piece of timber normal to the edge
Cup 3.26.3 Curvature of a piece of timber across the width of the face
Twist 3.26.4 Lengthwise spiral distortion of a piece of timber
Shrinkage 5.12 Decrease in dimension of a piece of timber due to reduction
of moisture content
Swelling 5.13 Increase in dimension of a piece of timber due to increase of
moisture content
Examples of requirements for maximum allowed warp for some products.

EN 518 Structural timber - visual grading, and EN 519 Structural timber - machine
grading
C18 and below Above C18
Bow max. 20 mm max. 10 mm
Spring max. 12 mm max. 8 mm
Twist 2 mm/25 mm width 1 mm/25 mm width
Cup no restrictions no restrictions
6 Bound water migration in wood
Stavros Avramidis
Department of Wood Science
The University of British Columbia
#4028-2424 Main Mall
Vancouver, BC, V6T 1Z4
Canada
stavros@interchange.ubc.ca

ABSTRACT. This paper presents the physical mechanisms involved in the migration of bound
water (diffusion) in wood. It explains the phenomena of isothermal and non-isothermal
moisture diffusion under steady-state and unsteady-state conditions and briefly examines the
effect of various wood and ambient parameters that affect the value of the diffusion coefficient

KEYWORDS: Wood, Fick’s law, concentration gradient, moisture content gradient, diffusion
coefficient, steady-state diffusion, unsteady-state diffusion, non-isothermal diffusion

Fundamentals of wood drying, Patrick Perré editor, pages 105 to 124.


106 Fundamentals of wood drying

6.1 Introduction

The study of the physics of water flow (bulk and molecular) through wood has
become basic for many wood processing applications and uses, quite apart from the
interest it holds for its purely scientific aspects. Such applications include timber
drying, pressure treatments, moisture migration through building wall systems and
mechano-sorption, amongst others.

Diffusion of water refers to the molecular movement of bound water through the
gross wood system by a combination of transfer through cell wall mass and lumen
void volume from a high water concentration area to a low one. This process is a
result of random molecular motions. The motion of a single molecule can be
described in terms of the random walk representation, and it is possible to calculate
the mean-square distance travelled in a given interval of time, but it is not possible
to say in what direction a given molecule will move in that time.

Moisture diffusion is the most important drying rate controlling factor in timber
drying. That is because the whole process is both affected by the wood structure and
water molecular concentration gradients within wood on one hand, and the time
dependent moisture profile evolution that can introduce internal stresses within the
wood matrix and thus affect the sorption site availability and local energy
dispersion.

The understanding of the diffusion process can greatly improve our comprehension
of drying and thus, allow researchers to develop drying schedules and decision
support systems with the help of new robust models that will improve the drying
quality of timbers.

6.2 Observation/Formulation

In 1855 Fick (in Fox et al. 1965) was the first to recognize that transfer of heat by
conduction was also due to random molecular motions, in analogy with the diffusion
process. Because of this, he was the first to place diffusion on a quantitative basis.
For this reason, the process of diffusion carries his name.

The mass flux of a given species is a vector quantity denoting the amount of the
particular species in mass units that passes per given increment of time through a
unit area normal to the vector. Usually, the flux is defined with reference to
coordinates which are fixed in space. The basic relation for molecular diffusion
defines the mass flux relative to the mass-average velocity. An empirical relation for
this mass flux, first postulated by Fick, accordingly, often referred as Fick’s first
law, defines the diffusion of a component in an isothermal, isobaric system. For
diffusion in the x-direction the Fick’s rate equation is,

 ∂c 
J = − D  (6.1)
 ∂x 
Bound water migration in wood 107

where J - water molecule flux, kg·m-2·s-1; D - diffusion coefficient, m2·s-1; ∂c/∂x -


concentration gradient kg·m-3·m-1 (Jost 1960, Crank 1975, Welty et al. 1984, Eckert
and Drake 1981, Bird et al. 1960). This also is referred as diffusion under steady-
state conditions, because the boundary conditions are kept constant with time.

Diffusion into or from a medium (i.e., drying in the latter case) is much more
complicated than the above case, because the boundary conditions continuously
change. This condition can be expressed in differential form by Fick’s second law of
diffusion also called the unsteady-state law,

 ∂c 
∂ D 
= 
∂c ∂x 
(6.2)
∂t ∂x
in this case, one-dimensional. The diffusion coefficient can be either a constant, or
can depend upon concentration, pressure, temperature, and composition of the
medium. In general, the diffusion coefficients of wood have been shown to be a
function of the above mentioned factors, as well as, the method of their
determination (Stamm 1964, Keey et al. 1999).

6.3 How to measure diffusion coefficients

There are two main experimental ways of measuring the water diffusion coefficient,
namely, the steady-state and unsteady-state methods. The former involves the use of
a diffusion cup and a thin wood specimen placed tightly at the top of the cup as
shown in figure (6.1). The cup is partially filled with distilled water (100% relative

Ht, Mt
L
quartz spring
Hb, Mb
specimen

solution

steady-state diffusion cup unsteady-state sorption apparatus

Figure 6.1. Methods of measuring the water diffusion coefficient

humidity) or a salt solution with a magnetic stirrer thus providing a relative humidity
on the bottom part of the specimen which is less that 100%. The cup with the
108 Fundamentals of wood drying

attached specimen is placed in a conditioning chamber where the ambient


temperature and relative humidity are controlled accurately. Depending on the
external/internal cup relative humidity (or vapour pressure) the assembly will either
loose or gain weight. By plotting this weight change as a function of time, after few
days a steady-state is reached where the plot becomes a straight line – linear
relationship between weight change and time. The diffusion coefficient through
gross wood (Dg) is then calculated from the following formula (Siau 1995),
100 m L
Dg = (6.3)
t A G ρ w ∆M

where m - mass of water vapour transferred through the specimen, g; L - specimen


thickness, m; t - time, s; A - surface area of specimen, m2; G - specific gravity of
wood at X ; w - normal density of water, 1000 kg·m-3; and ∆X - moisture difference
between the two parallel surfaces of the specimen calculated from the sorption
isotherm, %.
Xb + Xt
X= (6.4)
2
where Xb - bottom specimen surface moisture content, %; and Xt - top specimen
surface moisture content, %

The unsteady-state method involves the placement of thin wood specimens in a


controlled environment chamber under constant temperature and relative humidity,
and the monitoring of their weight change and thus, of moisture content, as a
function of time until equilibrium. The specimen’s initial moisture content (Xi) is
known and can be greater or lower than the equilibrium moisture content (Xemc) thus
making the specimen either loose (desorption) or gain (adsorption) water. Once the
final moisture content (Xf) at equilibrium is reached, the dimensionless change of
moisture content ( E ) is then plotted against time and from the initial straight part of
the plot (figure 6.2), the unsteady-state diffusion coefficient can be calculated based
on the analytical solution of Fick’s second law for parallel-sided specimens given by

D=
(E ) L2
2
(6.5)
5.1 t

X − Xi
E= (6.6)
X f − Xi

In each sorption step, the average moisture content ( X ) reported for the calculated
diffusion coefficient is given by

(
X = X i + (2 / 3) X f − X i ) (6.7)
Bound water migration in wood 109

The unsteady-state methods can be carried-out with thick specimens in controlled


environment chambers or with very thin specimens, less than 1mm, in diffusion
columns hanging from quartz springs (Kelsey 1956) as shown in figure (6.1). It must
be emphasized at this point that specimen thickness can also affect the calculated D-
values as has been shown by Avramidis and Siau (1987) and Chen et al. (1996).

1.0

_ slope
E

0
t

Figure 6.2. Typical plot of moisture content change against time in unsteady-state
experimentation

Both methods have their advantages and limitations, most important being the fact
that every effort should be made to eliminate potential points of vapour flow other
than the specimen itself in the steady-state method and elimination of exposure of
the specimens to the ambient environment during weight measurements in the
unsteady-state method. Furthermore, low air velocities in the unsteady-state method
can accentuate the effect of the surface resistance coefficient as explained in section
6.4.3. Non-Fickian diffusion phenomena (see section 6.4.4) can also affect the
calculation of D.

Lastly, it must be kept in mind that the diffusion coefficients calculated by the two
methods for the same specimen do not coincide numerically. At lower moisture
contents they are quite similar, but at higher moisture contents the steady-state
D-values tend to be about double of the unsteady-state ones. This is mainly
attributed to stress relaxation phenomena involved in the latter case. In the case of
drying modelling, we are mostly interested in the unsteady-state diffusion
coefficient values (Comstock 1963, Choong and Skaar 1972).
110 Fundamentals of wood drying

6.4 Isothermal water diffusion in wood

6.4.1 Moisture movement in wood

Moisture movement in wood in the hygroscopic range is very important in wood


science in general and wood drying in particular. A great deal of research has been
done, and the descriptions of bound-water diffusion appearing in wood science
literature, the mathematical procedures used, and the conclusions reached are
frequently ambiguous and inconsistent (Stamm 1959, 1960ab, 1967, Stamm and
Nelson 1961, Choong 1963, Skaar 1958, Bramhall 1976, Wadso 1993, Claeson and
Arfvidsson 1996).

The moisture movement in wood below the fibre saturation point is governed by at
least two different mechanisms: one is water-vapour diffusion in the cell lumen and
pit openings (intergas diffusion), and the other is bound-water diffusion in the cell
wall substance (Figure 6.3). The total flow rate in the transverse direction is
controlled by the latter due to the lower diffusion coefficient. In the axial direction,
the resistance of the lumen may be predominant (Skaar 1958, Siau 1995).

cell-wall
lumen
pit
pit chamber
pit pores

cell-wall

Figure 6.3. Water diffusion model

Under isobaric conditions, the water-vapour diffusion in the cell lumen is very
important at low relative humidities. The diffusion coefficient of water vapour in air
is given by a semi-empirical equation as follows:

 5  T 1.75
−5  1.013 ⋅10  
Da = 0.22 ⋅10 (6.8)
  273 
 p 
where p - total pressure, Pa; T - Kelvin temperature; Da - coefficient of inter-
diffusion of water vapour in bulk air, m2/s (Dushman and Lafferty 1962). It is clear
Bound water migration in wood 111

that Da increases with temperature and decreases with pressure. Water-vapour also
moves through pit openings and cell cavities. If the size of the openings is close to
or less than the mean free path of the water (0.1 µm), then the process is not a free,
but a hindered diffusion because the collisions of the water molecules with the walls
will be of the same frequency as collisions with other molecules. It has been found
that the resistance to the passage of vapour molecules through the very small pit
membrane pores is approximately 40 times that of free diffusion in space of the
same total cross-sectional area (Skaar 1988, Stamm 1964, Siau 1995).

The bound-water diffusion mechanism is the most important one because it controls
the total rate of water movement in wood in the transverse direction which is mainly
the way wood dries in conventional kilns. Bound-water molecules are attached to
the sorption sites within the wood by hydrogen bonds. The water molecules, even
though fixed with respect to each other, have considerable vibrational energy, and
there is a random distribution of energy among the molecules and components. The
vibrational velocities, the square root of kinetic energy, of the sorbed molecules
follow a Gaussian distribution. When a water molecule receives enough energy to
break its bond, it becomes "activated" and it is able to migrate to another sorption
site, if it is available in the neighbourhood. The energy required for the molecule to
become “activated” is called activation energy, Eb, measured in kJ per mol and is a
function of the moisture content of wood (Bramhall 1976, 1979, Mrowec 1980,
Skaar 1988, Siau 1995).

6.4.2 Isothermal diffusion equations

The ultimate goal of a diffusion study regarding the wood-water system is to


determine the diffusion coefficient of water through wood. Diffusion coefficients are
generally expressed by Fick’s first or second laws of diffusion. The first law is used
to calculate diffusion coefficients for the steady-state condition, where the two
surfaces of a specimen of wood are held at different but constant moisture contents
by controlling the relative humidities of the atmospheric environment of the two
sides. After some time, a condition of equilibrium is attained and the flux of
moisture becomes constant from the high to the low moisture content side of the
specimen under isothermal conditions. The one-dimensional diffusion coefficient
will be given by solving equation (6.1) (McNamara and Hart 1971, Hart 1964, Skaar
1988, Siau 1995). The steady-state method of measurement is easy to use, involves
simple mathematics, but requires considerable time before steady-state conditions
are attained.

The unsteady-state method is widely used because the time required for completion
is much less compared to the steady-state method. The main disadvantage of it is the
mathematical complexity and the simplifying assumptions that must be made, which
are not always in agreement with the physical situation (Skaar 1988).

The unsteady-state method involves measuring the rate of change of moisture


content of a slab of wood. The specimen in equilibrium with one relative humidity is
112 Fundamentals of wood drying

exposed to a different one and therefore it either loses or gains moisture to or from
the surrounding atmosphere until equilibrium is attained. Fick’s second law of
diffusion is applicable to this situation. For a constant diffusion coefficient equation
(6.2) becomes,

∂c  ∂ 2c 
= D  (6.9)
∂t  ∂ x2 
 
In the case of wood, the diffusion coefficient is strongly dependent on the moisture
content, thus equation (6.2) is transformed to,

 D(c ) ∂c 
∂ 

=  
∂c ∂x
(6.10)
∂t ∂x
and

∂c  ∂c  ∂D(c )  2
=    + D (c ) ∂ c (6.11)

∂t  ∂x  ∂x  ∂x 2
or
2
∂c  ∂c  ∂D(c ) ∂ 2c 2
= ∂  + D(c ) (6.12)
∂t  ∂x  ∂c ∂x 2
For small concentration changes, (∂c/∂x)2 · ∂D(c)/∂c is quite small and can be
neglected so that Equation (6.12) is simplified to equation (6.9). Concentration can
be substituted for moisture content, so that equation (6.9) becomes,

∂X  ∂2 X 
= D  (6.13)
∂t  ∂ x2 
 

∂E  ∂2E 
= D  (6.14)
∂t  ∂ x2 
 
in terms of the dimensionless variable E, defined as
X − Xi
E= (6.15)
X f − Xi

where E - fractional change in moisture content; Xi - initial moisture content, %;


Xf - moisture content in equilibrium with the of external surface or final moisture
content, %; X - moisture content at point x at time t. The solution of Equation (6.9)
gives the fractional change E in wood moisture content at any location x and time t,
Bound water migration in wood 113

when this solution is integrated over the total wood thickness the mean value E of E
is obtained as a function of time t. The value of E may then redefined in terms of
either average wood moisture content X or wood weight W as in equation (6.6) and

E=
(W − Wi ) (6.16)
(W f − Wi )
where W - weight of the specimen at the time E is determined, g; Wi - initial
weight, g; and Wo - final weight, g.

Equations (6.13) or (6.14) were solved by Tuttle (1925), Sherwood (1929), and
Newman (1931a) for the drying of wood or a porous solid. The assumptions made
were that the diffusion occurs normal to the plane of the surface, no dimensional
changes occur, moisture is initially distributed evenly throughout the specimen, and
the surface of the specimen immediately attains equilibrium with the surroundings.
After setting the boundary conditions for these assumptions, and for a specimen of
thickness 2a, the solution of equation (6.14) after integration was given by an
infinite series as:

 
2
 2 2 
 1  exp − π (2n + 1) tD 

8 ∞
E = 1− (6.17)
n = 0    
π2  2n + 1   4 a 2 

The first term of this equation gives a good approximation of the whole series for
times less than that when half of the moisture has been adsorbed, E ≤ 0.5.
Therefore, equation (6.17) can be written (Crank 1956) as,

8  π 2tD 
E =1− exp −  (6.18)
2 
π2  4a 
and

4a 2  8 
D= ln − ln 1 − E  ( ) (6.19)
π 2t  π 2 
The most commonly used equation for calculating the diffusion coefficient has been
derived from Fick’s second law for an infinitely thick slab, and it is known as
Boltzman’s solution (Crank 1956, Newns 1956). This equation is written as,
0.5
 2  Dt 
E =   (6.20)
  2 
 π  a 

 π a 2  E 2 
D =  


 (6.21)
 4  t 
114 Fundamentals of wood drying

and is valid for slabs of finite thickness as long as the moisture change at the center
of the specimen is not large. The assumptions made for the Boltzman’s solution of
Fick’s second law include diffusion occurring normal to the plane of the specimen's
surface, no dimensional change of specimen, even initial distribution of moisture
content inside the specimen, and the surfaces of the specimen immediately attain
equilibrium with the surroundings. Equation (6.21) can be applied to finite thickness
specimen for the interval 0 ≤ E ≤ 0.5, the error being 0.001 or less (Crank 1956) as
explained in section 6.3.

6.4.3 The diffusion coefficient

Quantitative measurements of the rate at which a diffusion process occurs are


usually expressed in terms of a diffusion coefficient. The magnitude of this diffusion
coefficient is a function of many parameters. Before the examination of those
parameters, the distinction between true and apparent diffusion coefficient should be
made.

The transport of water between wood and air is controlled by two resistances,
namely, the external resistance due to the boundary layer adjacent to the wood
surface, and the internal resistance due to wood structure (Choong and Skaar 1969,
1972, Rosen 1978, Avramidis and Siau 1987, Chen et al. 1995, Siau and Avramidis
1996). The true diffusion coefficient, D, reflects the internal resistance, and surface
emission coefficient, S, reflects the external resistance. What is measured during a
diffusion experiment is the apparent diffusion coefficient D' which includes both of
these resistances.

Newman (1931a) separated the external from the internal resistance and derived
solutions for the diffusion equation which included the surface emission factor.
These solutions were put in an empirical first degree polynomial form by relating
Dt/a2 to the Ha values through linear regression by Choong and Skaar (1969, 1972).
The derived equation is,

t0.5 D  1 
= 0.2 + 0.7 
 (6.22)
2
a  Ha 
where D - true diffusion coefficient, m2/s; Ha - ratio of surface emission coefficient
to the true diffusion coefficient, (S/D), m-1; a - half thickness of specimen, m;
t0.5 - time during which half sorption has been completed.

The factor Ha can be considered to be the ratio of the internal-to-external resistance.


The true diffusion coefficient equals the apparent one when the external resistance is
zero, i.e., Ha = ∞, then equation (6.22) becomes,

t0.5 D '
= 0 .2 (6.23)
a2
Bound water migration in wood 115

By substituting Equation (6.24) into (6.23) we attain,


a a 3 .5
= + (6.24)
'
D D S
By plotting a/D' versus a or t0.5/a2 versus 1/a, a straight line is obtained and from its
slope and intercept D and S can be calculated (Choong and Skaar 1972, Avramidis
and Siau 1987, Siau and Avramidis 1993). The surface emission coefficient depends
on the concentration difference between wood surface and air, and the basic
equation is,
J
S= (6.25)
ca − cw

where J - flux of water from or to the wood, kg/m2·s; cw - moisture concentration of


wood surface, kg/m3; ca - moisture concentration of wood in equilibrium with air,
kg/m3. Equation (6.25) can be easily converted if we assume negligible variation of
specific gravity with moisture content to the following one,
 α  dp   α p0  dH 
S = 100   =
 



 (6.26)
 G  dX   G  dX 
where - constant of proportionality relating flux and relative humidity difference,
G - specific gravity; dH/dX - inverse of slope of sorption isotherm; p0 - saturated
vapour pressure, Pa (Choong and Skaar 1972). The surface emission coefficient has
been found to be strongly influenced by the surface flow conditions and by the air
velocity. Increasing the latter, increases S due to the decrease of the thickness of the
boundary layer (Rosen 1978, Siau and Avramidis 1993). Turbulent air flow will
increase S because there will be a physical movement of packets of material (water)
across streamlines transported by eddies present in the turbulent flow (Welty et al.
1984). The surface emission coefficient has also been found to decrease with
relative humidity and partial vapour pressure and increase with temperature
(Avramidis and Siau 1987, Cai and Avramidis 1997).

The way of calculating the surface emission coefficient has been quite controversial
since the early 1990’s. Söderström and Salin (1993) had argued against the above
described method of estimating the surface emission coefficients. Their detailed
arguments are presented in the “External heat and mass transfer” chapter of this
book. More focused research on this particular topic is needed for further
clarification.

In many hydrophilic polymers like wood the diffusion coefficient has been shown to
increase with moisture content (Crank and Park 1968, Fox et al. 1965, Siau 1995,
Chen et al. 1995, Claesson and Arfvidsson 1996, Olek and Weres 2001). Water
swells and plasticizes the polymer structure, causing increased mobility for both
penetrant and polymer segments. Plots of E against square root of time for a
concentration dependent polymer-water system, have the following characteristics:
116 Fundamentals of wood drying

• linear curve from zero up to 50% or more of the total moisture content
change, becoming concave toward the horizontal axis with no inflection
point;
• when the diffusion coefficient increases with X, desorption is always slower
than adsorption, with the reverse true for decreasing coefficient;
• for D increasing with X, the deviation of the adsorption curve from the
curve of constant D is greater than that of the desorption curve and vice
versa;
• plots of E vs. (t/L)0.5, where L is specimens thickness (= 2a) are the same
for different thicknesses specimens;
• the surfaces of the specimens attain equilibrium conditions instantly and
remain constant with time.

Table 6.1. Some values from the literature pertaining to experimentally obtained
diffusion coefficient values by the unsteady-state
X-range T-range Species D Reference
(%) (oC) (m2/s)
8.4 -> 16.1 40 Yellow poplar 0.9 -> 2.38E-12 Comstock (1963)
Green -> 10 45 Redwood 34.2E-7 Chen et al. (1995)
Red oak 8.14E-6
Green -> 10 45 Southern pine 24E-6 Rosen (1974)
Yellow poplar 28E-6
11 -> 27 60 Scots pine (H-R) 10.8 -> 28.6E-10 Claesson and
Scots pine (S-R) 11.1 -> 26.6E-10 Arfvidsson
(1996)
30 -> 18 60 Sweetgum (S-T) 10.4E-12 Choong and
Redwood (S-T) 47E-12 Skaar (1972)
Redwood (H-T) 2.6E-12
0 -> 120 30 Hard maple (R) 1.5E-10 -> 1E-9 Yeo et al. (2001)
Red oak (R) 0.5E-10 -> 6E-10
Southern pine (R) 2E-10 -> 4E-10
8.2 -> 0 40 Beech (S-R) 6.38E-10 Marinescu et al.
8.2 -> 0 100 Beech (S-R) 3.6E-8 (2001)
9.3 -> 0 40 Beech (H-R) 3.8E-10
9.3 -> 0 100 Beech (H-R) 1.72E-8
30 Black walnut (R) 6.9 -> 88E-11 Rosen (1978)
18.8 -> 30 60 Scots pine (H) 6.2 -> 25E-10 Rosenkilde and
Soderstrom
(1996)
Green -> 6 50 Hemlock (S-R) 1.69 -> 0.40E-10 Koumoutsakos
Hemlock (H-R) 1.11 -> 0.38E-10 and Avramidis
Red cedar(S-R) 0.90 -> 0.45E-10 (2002)
Red cedar (H-R) 1.60 -> 0.46E-10
S = sapwood, H = heartwood, R = radial, T = tangential
Bound water migration in wood 117

The above characteristics constitute what is called “Fickian” diffusion, and any
deviation indicates “non-Fickian” or “anomalous” diffusion (Crank and Park 1968,
Fox et al. 1965, Avramidis and Siau 1987, Wadso 1993).

The analytical method described is based on the traditional concept of water


concentration being the driving force for diffusion as shown in equation (6.1).
However, over the last years, various other driving forces have been proposed and
tested. The most promising ones are the partial vapour pressure (Bramhall 1976), the
spreading pressure (Skaar and Babiak 1982), the chemical potential (Siau 1984) and
water potential (Cloutier and Fortin 1993). The last two have extensively used to
model wood drying (Avramidis et al. 1992, Avramidis and Hatzikiriakos 1995,
Fortin et al. 2001).

In the case of wood, dependence of the diffusion coefficient on the moisture content
has been observed (Figure 6.4). Many researchers have found that the apparent
diffusion coefficient calculated by the unsteady-state method increased with
moisture content (Stamm 1956ab, 1959, 1960ab, Skaar 1958, Choong 1962, 1965,
Comstock 1963, Rosen 1976, Avramidis and Siau 1987, Wadso 1993, Siau 1995).
An explanation of this trend was given by Stamm (1956ab), according to whom, the
wood and all cellulosic materials hold water by polymolecular adsorption, up to
about seven molecular layers, by hydrogen bonding. Every successive layer is
attracted to the sorption sites by weaker forces due to the greater distance from the
substrate. As water is moving, one molecule at a time, it is easier to break a weak
water-water bond, than a water-cellulose one. As a result, molecules of the outer
layers have greater mobility and more freedom to move. Additionally, swelling
pressure due to the increased number of molecules, breaks cellulose-cellulose bonds
and increases the area through which diffusion can occur.

The clustering phenomena of water molecules can also explain the increase of the
diffusion coefficient with moisture content (Starkweather 1963, 1975, Jacobs and
Jones 1990, Hartley and Avramidis 1994). Because both cellulose and water are
polar substances, there is a tendency for the water to cluster in the vicinity of the
sorption sites. At high moisture contents the density and size of the clusters will be
sufficiently great such as overlapping of cluster boundaries within capillaries or
voids parallel to the main diffusion axis could take place. As a result, a continuous
liquid-like path through the cell wall structure can be formed. As the amount of
“liquid” phase within the cell wall increases, the observed unidirectional diffusion
coefficient should increase rapidly (Hartley et al. 1992, Hartley and Avramidis
1993).

The overall rate of water transport depends not only on the effective number of
mobile water molecules, but also on the ease of hole formation which in turn is
dependent on the mobility of the polymer segments (segmental mobility). Two types
of processes may occur. One is the breakage of hydrogen bonds by the penetrating
water molecules due to the swelling pressure. The other is the fact that when the
chemical natures of penetrant and polymer are not too different, then the fraction of
118 Fundamentals of wood drying

polymer-penetrant contacts should be proportional to the total sorbed concentration.


If these polymer-penetrant contacts are assumed to be weaker than polymer-polymer
contacts, the energy required to form a hole of certain size decreases linearly with
increasing concentration. Therefore, the density of the holes large enough to permit
a molecular jump increases exponentially with concentration (Kishimoto and Fujita
1958, Rogers 1965).

Some investigators have reported on the other hand, irregular behaviour of the
diffusion coefficient with respect to moisture content. Christensen (1960), and
Christensen and Kelsey (1959a) reported an initially increasing and then decreasing
D' with moisture content over the same thickness, and a constantly decreasing
coefficient with increasing moisture content below that particular thickness, for
Klinki pine in the tangential direction in the absence of air. This behaviour was
attributed to a stress relaxation mechanism controlling the diffusion rate.
Prichananda (1966), also reported a continuous decrease of D', as moisture content
increased in the transverse direction. Relaxation phenomena were also “blamed” for
this irregularity.

5.01E-09 20C
Diffusion Coefficient (m^2/s)

40C
60C
4.01E-09
80C
100C
3.01E-09

2.01E-09

1.01E-09

1.00E-11
5 10 15 20 25 30
Moisture Content (% )

Figure 6.4. Transverse diffusion coefficients for wood with oven dry specific gravity
of 0.5 and various T and X

Temperature is another factor by which diffusion is affected. The bound-water and


free-water-vapour diffusion coefficients have been shown to be affected positively
by the increase of the temperature (Stamm 1956ab, 1959, 1960ab, 1961). The same
thing was also found for the combined diffusion coefficient. When the logarithms of
the diffusion coefficients are plotted against the reciprocal of the absolute
temperature, a straight line results, which indicates that the process is an activated
one (Stamm 1959). The diffusion - temperature relationship of moisture through
wood, is generally expressed by an Arrhenius type equation as,
Bound water migration in wood 119

 E 
D = D0 exp − b  (6.27)
 RT 
where Eb - activation energy, J/mol; D0 - constant calculated from a plot at 1/T = 0.

Stamm (1956ab, 1959), also concluded that the increase of D' with temperature in
cellophane and wood is approximately in the same magnitude as the increase in
vapour pressure of water. The above results strongly support the theory that water in
wood moves in a molecular-jump fashion rather than as a continuous phase.

Lastly, the gross wood diffusion coefficient is inversely proportional to wood


specific gravity due to thicker cell walls, and it is also affected by the flow direction
in relation to the fibre direction, the longitudinal ones always being greater than the
transverse. The ratio of longitudinal to transverse diffusion coefficient is about 2.5:1.
There are mixed results regarding the effect of sapwood and heartwood on the
diffusion coefficient. Diffusion coefficients for sapwood were reported as being
greater compared to heartwood ones by Choong and Skaar (1972) for sweetgum and
redwood, but the opposite trend was reported by Chen et al. (1994) for five
hardwoods.

Empirical models that can be used to calculate the values of D for various cases as a
function of the wood’s anatomy and flow direction can be found in Siau (1995). In
general, the transverse diffusion coefficient values range from 5·10-11 to 8·10-10 m2/s
for 20ºC and 20% moisture content to 3·10-9 to 5·10-8 m2/s for 100ºC and same
moisture content. For the same temperature, i.e., 20ºC, the same coefficient will
range from 3·10-11 at 5% moisture content, to 4·10-10 at 30% moisture content. The
same values will range at 100ºC from 7·10-10 to 5·10-9, respectively. Little difference
exists between radial and tangential diffusion coefficients, the former mostly being
greater by about 20-50%.

6.4.4 Non-Fickian diffusion

In non-Fickian or anomalous diffusion the E versus t0.5 plots are not linear in the
initial stages (E<0.5), meaning that the process does not obey Fick’s law. This
behaviour has been observed many times with wood and polymers below the glass
transition temperature when adsorbing water or other gases that swell the polymer.
There are three different non-Fickian curves resulting from the above plot, namely,
pseudo-Fickian, two-stage, and sigmoid, shown in Figure 6.5 (Kishimoto et al.
1960).

Rogers (1965), trying to explain the causes of non-Fickian behaviour, points out the
following as possible reasons:

• a diffusion coefficient which is a function of concentration, time, and


history of the specimen;
• time dependent boundary conditions;
120 Fundamentals of wood drying

• multiple transport mechanisms

fickian pseudo-fickian

sigmoid two-stage

Figure 6.5. Various types of water diffusion curves

Generally speaking, there is an agreement among researchers on the cause of non-


Fickian diffusion, focused principally on the rate of changes of polymer structure
due to the stresses imposed upon the medium during the sorption-diffusion process.
These effects can be directly related to the changes in specimen structure affecting
diffusivity and solubility, or due in part to internal stresses exerted by one part of the
medium on another as they change with time, distance, and concentration (Barkas
1942, Park 1968).

The rate at which the final equilibrium is attained in a sorption study, depends on the
rate of diffusion and configurational changes, the latter affecting directly the rate of
stress relaxation. Stress relaxation times decrease with an increase in temperature or
solvent or sorbate concentration due to enhanced segmental mobility (Kishimoto and
Fujita 1958, Fujita and Kishimoto 1958). Park (1953), working with diffusion of
methylene chloride in polystyrene and cellulose acetate films, concluded that the
process is Fickian when the initial concentration and imposed concentration
gradients are very small. At higher concentrations, the diffusion changed to non-
Fickian. Similar behaviour was also reported by Barrer and Barrie (1958) and Long
and Thompson (1955), working with diffusion of water in cellulosic polymers, and
by Watt (1960), and Watt and Algie (1961), for the wool-water system.

Christensen (1960), and Christensen and Kelsey (1959a) reported similar results for
the wood-water system. He showed an increase of the time required for the same
amount of moisture to be adsorbed as the moisture content of the specimen
increased, meaning a decrease of the apparent diffusion coefficient as moisture
content increased. Park (1953), reported an increase of the apparent diffusion
Bound water migration in wood 121

coefficient with the thickness of the specimen, a phenomenon which was attributed
to “time effects”. The same phenomenon was shown theoretically by Crank (1951).
Christensen (1960), Christensen and Kelsey (1959a), and McNamara and Hart
(1971), reported the same phenomenon with wood. They attributed this abnormality
in part to a time dependent deformation phenomenon in the wood. Skaar (1954),
additionally suggested that it might result in part from insufficient air velocity, and
consequently an increased surface resistance. This dependence of the apparent
diffusion coefficient or the rate of adsorption on the thickness, was observed during
the linear part of the E vs. t0.5 curve (initial stage), and not in the curved part (second
stage).

Bramhall (1976), trying to explain the abnormalities observed and the driving forces
during diffusion of water in wood, postulates that the process is not a Fickian one,
because only an extremely small, temperature-dependent proportion of molecules is
migrating at any instant, whereas in Fickian diffusion all molecules are migrating at
all times; hence, Fick’s laws do not necessarily hold in the case of bound-water
diffusion and that can explain in part the anomalies observed.

6.5 Nonisothermal water diffusion in wood

6.5.1 General considerations

The movement of moisture through wood in the hygroscopic range, is an important


phenomenon which has been studied extensively. Though there is hardly a natural
situation in which the surface layers of wood are in an isothermal state, nearly all the
experiments and analyses of moisture transfer, have been done under constant
temperature conditions. Today, there is little reference to nonisothermal diffusion of
water vapour in wood in the wood science literature.

Babbit (1940), was the first to measure moisture gradients in the transverse direction
of fiberboards under nonisothermal conditions, and zero flux. These results are not
typical of wood transverse to the grain due to continuous openings in which gaseous
interdiffusion would dominate. Nonisothermal experiments on solid wood were first
performed by Voight et al. (1940), and Choong (1963). Encapsulated wood
specimens were used to prevent loss or gain of moisture, on which a steep
temperature gradient was imposed. After a period of time, when the specimens
reached equilibrium, the moisture profiles inside them were measured by slicing the
specimens. The results showed the existence of a moisture gradient even though the
flux was zero.

Bramhall (1979), trying to explain this phenomenon analyzed the experiments of


Voigts and Choongs, and showed based on their data, which at the termination of the
experiments there was a moisture and an opposite vapour pressure gradient. The
above shows that, if a vapour pressure gradient exists under equilibrium conditions,
it means that it is not the driving force for diffusion. By examining the experiments
more closely, he concluded that in some of them the equilibrium state was probably
122 Fundamentals of wood drying

not reached. In those that equilibrium was reached, after correction of X for
hysteresis due to the fact that one surface was adsorbing and the other was desorbing
until the equilibrium state, he concluded that the vapour pressure gradient was zero
when diffusion was not occurring. His final conclusion was that vapour pressure
gradient under nonisothermal conditions is the force responsible for diffusion.

Siau and Babiak (1983), Siau and Jin (1985), Siau et al. (1986), and Avramidis et al.
(1987), studied nonisothermal movement of moisture through wood in the transverse
direction. The experimental procedure made it possible to determine the magnitude
and direction of the moisture flux and to calculate the moisture profiles inside the
specimen. The results strongly supported the assumption that, under nonisothermal
conditions, the isothermal forms of Fick's law did not apply. Siau and Jin (1985),
showed clearly hat the partial vapour pressure gradient is not the driving force for
diffusion under nonisothermal conditions because the predicted flux direction from
Fick's first law is always from the high vapour pressure to the low one. In the
nonisothermal experiments, the opposite phenomenon was observed, which is the
movement of moisture from low vapour pressure region (cool surface) to high one
(warm surface) in the initial experimental stages when the relative humidity of the
warm surface was low. An additional analysis using the same law, based on the
moisture content gradient as the driving force, could not explain the movement of
moisture from low to high moisture content regions as the relative humidity of the
warm surface increased. Only equations combining moisture and temperature
gradients as the driving forces for diffusion could predict the above mentioned flux
directions.

The phenomenon of the coexistence of a moisture gradient with a zero flux, or that
of flux direction opposite to the moisture gradient or partial vapour pressure gradient
can be explained by the principles of nonequilibrium thermodynamics and in
particular by the thermal diffusion or Soret effect phenomenon.

6.5.2 Theoretical equations

The derivation of the main steady-state nonisothermal moisture diffusion model


validated by Avramidis and Siau (1987) will not be included here, however,
interested readers can look it up in the series of papers listed in the references
section of the above publication.

The model (thermodynamic) was based on chemical potential being the driving
force of diffusion
H  ∂X  dµ 0 E L + E0 + Eb  Hp0   dT dX 
J = −K X    + + R ln  +  (6.28)
 RT   
 ∂H  dT T  7600   dx dx 
where KX - coefficient for bound diffusion based on a gradient of percent moisture
content (= DTG/100 w); 0 = chemical potential of water vapour at 1 atm at
temperature T, Pa; p0 - saturated vapour pressure, Pa; E0 - heat of vaporization of
Bound water migration in wood 123

water, J/mol; EL - differential heat of sorption, J/mol; Eb - activation energy, J/mol.


The model successfully predicted the flux magnitude and direction in various
nonisothermal cases (Avramidis and Siau 1987a, 1987b).

Siau and Avramidis (1993) tested a more simplified version of equation (6.28) with
the same experimental data
 H ∂X Eb dT dX 
J = −K X  +  (6.29)
 RT ∂H T dx dx 
with good prediction results of the direction and magnitude of moisture’s flux.
Peralta and Skaar (1993) also did “zero-flux” type experiments and tested an
activated moisture type nonisothermal diffusion model however, when compared to
equation (6.28), it gave poorer predictions of the moisture fluxes as shown in Figure
6.6 (Siau 1995).

X, %
Figure 6.6. Steady-state nonisothermal water flux changes as a function of the high
moisture content side, X (from Siau and Avramidis 1993)

Avramidis et al. (1992, 1994) took the nonisothermal diffusion model a step further
by transforming it to an unsteady-state form for predicting moisture fluxes during
wood drying. Evaluation of the new moisture model listed as,
2 2
∂X  ∂2 X   ∂D  ∂X   ∂φ ∂D  ∂X  ∂T   ∂φ  ∂T   ∂ 2T 
= D +   + +   +   + φ 
 ∂ x2   ∂X  ∂x          ∂ x2 
∂t       ∂X ∂T  ∂x  ∂x   ∂T  ∂x   

(6.30)
124 Fundamentals of wood drying

where

 Eb  DH  ∂X 
φ =     (6.31)
  
 T  RT  ∂H 
A very good agreement between the measured and predicted average moisture
contents were reported. The results of a simulated situation when moisture flows
under unsteady-state nonisothermal conditions is shown in Figure 6.7 adopted from
Avramidis et al. (1994).

Figure 6.7. Average moisture content predicted by Equation (6.30) in unsteady-


state nonisothermal diffusion (drying)
7 Fluid migration in wood
Patrick Perré

LERMAB
Laboratoire d´Etude et de Recherche sur le Matériau Bois
UMR1093 INRA, ENGREF, UHP,
ENGREF 14, rue Girardet 54 042 Nancy Cedex
France
perre@nancy-engref.inra.fr

ABSTRACT. This chapter presents the physical mechanisms involved in the migration of fluid
(gas and liquid) in wood. The principle of single-phase flow in porous media is presented
first. Then, specific features of wood as a material are summarised: orders of magnitude,
anisotropy, and specific features. Possible deviations from the linear Darcy’s law are also
addressed. Finally, because two fluid phases exist in wood during drying, a gaseous phase
and a liquid phase, the last part of this chapter is devoted to the important concepts of
capillary pressure and relative permeability.

KEYWORDS: Wood, Darcy’s law, pressure gradient, capillary pressure, single phase flow,
multiphase flow, slip flow, percolation model, effect of sample length, relative permeability.

Equation Section 7

Fundamentals of wood drying, Patrick Perré editor, pages 125 to 156.


126 Fundamentals of wood drying

7.1 Introduction

Liquid flow in wood appears at first in the ascent of sap in tree. To induce this
migration, the menisci present in the leaf stomata pull up water. Because most trees
are more than 10 meters high, one can deduce that the absolute liquid pressure in the
sap column is negative. No gaseous phase can exist in such conditions: this is an
excellent example of single-phase flow in porous media. The vascular system
developed in trees has many other surprising features:
- designed to allow the sap flow from roots to leafs (or needles), the wood
material is strongly anisotropic,
- because of negative pressure, the vascular system must be able to support a gas
invasion due to injury or cavitation. This is the role of bordered pits, or vessel to
vessel pits. These anatomical features may inhibit dramatically the fluid
migration in the wood material,
- when the wood produced by the tree is more than some years old, typically
10-year old, the need for sap flow disappears. During heartwood formation, the
permeability is often reduced by one or several orders of magnitude,
- the wood is fully saturated in the sapwood part of logs (an air-free sap column is
required to obtain negative pressures), whereas the heartwood part is generally
only partly saturated.
The understanding of the fluid migration in wood is of utmost importance to
understand the drying process. The differences in drying behaviour between species
and within the log primarily come from the permeability value and from the initial
moisture content.

The principle of single-phase flow in porous media is presented first. Then, specific
features of wood as a material are summarised: orders of magnitude, anisotropy, and
specific features. Possible deviations from the linear Darcy’s law are also addressed.
Finally, because two fluid phases exist in wood during drying, a gaseous phase and a
liquid phase, the last part of this chapter is devoted to the important concepts of
capillary pressure and relative permeability.

7.2 Darcy’s law: single-phase flow

7.2.1 Observation, formulation

Permeability refers to the conductivity of a porous medium with respect to fluid


migration under the influence of a pressure gradient. We refer here to a bulk flow,
rather than to a diffusive flow. This concept originates from Darcy’s work on the
public fountains of the city of Dijon, a French city (Darcy 1856). From numerous
and careful experiments on samples of sand, he concluded ”It thus appears that for
sand of comparable nature, one can conclude that output volume is proportional to
the head and inversely related to the thickness of the layer traversed”. Note that this
conclusion stands for moderate values of the mass flux (see 7.6 for possible
deviations from this linear relationship). This leads to a simple formula, between the
pressure gap through the sample ∆P, the sample length L in the flow direction, the
Fluid migration in wood 127

cross-sectional area of the section perpendicular to flow direction S and the


volumetric flow rate Q (Fig. 7.1):
QL
= constant (7.1)
S ∆P
Using the international standard, Q is in m3·s-1, L in m, S in m2 and ∆P in Pa.

By introducing the dynamic viscosity of the permeating fluid µ (Pa·s), the constant
depends on the porous medium only. This constant, usually denoted as K (m2), is the
specific permeability:
QLµ
K= (7.2)
S ∆P
The specific permeability K, simply denoted permeability in the following part of
this text, represents the resistance of the porous medium to fluid flow. This is an
intrinsic value, which accounts for the complex morphology of the pores in the
specimen. At the microscopic level, one has to imagine an intricate system of
capillary tubes. The fluid viscosity imposes the fluid to be at rest at each solid wall
present in this system, hence this resistance to fluid migration.

S L

P1 P2 < P1

Flow rate Q Flow rate Q

Porous medium

Figure 7.1. Fluid flow in a porous medium: the volumetric flow rate is proportional
to the pressure gap P1 – P2

From equation (7.2), it is easy to derive the well-known local expression of Darcy’s
law:

K
v=− ∇( P) (7.3)
µ
where v is the apparent velocity of the fluid through the specimen (m·s-1), K the
permeability (m2) and µ the dynamic viscosity of the fluid (Pa·s-1).
128 Fundamentals of wood drying

In the case of a gas flow, due to the mass conservation, the volumetric flow rate
varies along the pressure gradient, even in the steady-state. By taking into account
the compressibility of the air, the permeability value, K, is calculated by integrating
the local expression of Darcy’s law (7.3) along the sample length:
Q L µg P
Kg = (7.4)
S ∆P P

with: K - Permeability (m2)


Q - Volumetric flux (m3/s)
µ - Dynamic viscosity of the air (Pa·s)
e - Sample thickness (m)
P - Pressure at which flux Q is measured (Pa)
S - Area of the right section of the sample (m2)
∆P = P2-P1 - pressure gap (Pa)
P = (P1+P2)/2 - average pressure (Pa)

7.2.2 How to measure permeability

Permeability can be measured using liquids or gas. The experimental principle is


rather simple. Although transient methods have been tested, especially to deal with
very low-permeable specimens, the steady-state method is usually well adapted for
permeability determination. Depending on the permeability range and on the fluid,
one of the two following possibilities can be used in practice:
- the pressure gap through the sample is controlled and the resulting flow rate is
determined,
- a certain flow rate is imposed to the specimen, by a volumetric pump for
example, and the resulting pressure gap is measured.
In spite of this very simple principle, several traps have to be avoided, most often
thanks to imaginative tricks (Stamm 1963, Resch and Ecklund 1964, Banks 1968,
Choong and Kimbler 1971, Tesoro et al. 1972, 1974, Bolton and Petty 1978):
- due to the huge range of permeability values encountered on porous media, and
especially wood, the apparatus must be designed to measure accurately wide
ranges of flow rates and/or pressure gaps,
- whatever the flow rate and the pressure gap, bypassing of the specimen by the
fluid must be unquestionably avoided,
- in the case of permeability to liquid, the specimen can be clogged by the
retention of small particles or by the formation of entrapped air bubbles. A
proper pre-treatment of the permeating fluid is required (filtration, deaeration).

Figure 7.2 depicts a schematic diagram of the first device used by the author to
determine gaseous permeability (Perré 1987). It allows the permeability to be
measured over a huge range of values (typically from 10-11 m2 down to less than
10-18 m2). The gas flows from one container to the other through the sample and the
flowmeter. The sample support used the clever idea proposed by Kesoro et al.
(1972). The latter consists of an aluminium pipe covered inside with a rubber tube
Fluid migration in wood 129

whose diameter is less than the specimen diameter. By using a vacuum pump, the air
between the pipe and the rubber tube can be aspirated through a small hole in the
pipe, causing the adhesion of the inner tube to the pipe. This trick ensures an easy
implementation of the samples. When the air is released, due to its small diameter,
the pneumatic vice sticks strongly to the specimen. Because the measured flux can
be very low, absolutely no lateral leakage can be accepted. This is why the chamber
between the rubber tube and the aluminium tube is inflated to 2 to 3 bar during the
experiment to tight the rubber joint even better against the sample. It is important to
notice that this configuration is efficient to avoid bypassing only when the lateral
face of the sample is insulated with a double coat of epoxy resin.

Pressure or
vacuum

Calibrated capillary tube


Porous discs Wood specimen

Figure 7.2. Schematic diagram of the device used to measure gaseous permeability
(after Perré 1987)

The volumetric flow rate is determined by the measurement of the pressure gap
through a calibrated capillary tube (Fig. 7.2). The existence of several calibrated
tubes with a wide range of sections allows the operator to match any value of the
actual flux. The limitation for very low flow values comes from the time constant of
the system, which required very stable conditions: any temperature change causes
measurement disturbances. One of the advantages in using a pressure gap to
measure the flux lies in the fact that the air viscosity vanishes in the equations. So
the viscosity value does need to be known: it is sufficient for the specimen and the
tube to be at the same temperature to obtain accurate values.

Thanks to the presence of two rigid tanks, the entire system is isolated from the
surrounding air: the noise level of the pressure and flow measurements is
130 Fundamentals of wood drying

dramatically reduced and the permeability to air can be determined at various


pressure levels (from 20 mbar to 1 bar).

The new device, used at present in our laboratory (Agoua 2001, Perré and Agoua
2002) is based on the same principle. It just takes advantage of recent equipment,
such as an electromechanical pressure controller and a mass flowmeter. The latter,
which is based on the thermal disturbance produced by the flow rate, is not sensitive
to the pressure level (the slip flow – see 7.6 – can be important in the small
calibrated capillary tube at very low pressure levels). To widen the range of flow
rates, calibrated tubes (6 mm of internal diameter and different lengths) allow one
known part of the mass flux to bypass the flowmeter. So, this device is well adapted
to measurement at different pressure levels, down to about 15 mbar for the average
pressure in the specimen.

The experimental protocol comprises the following important steps:


- the wood samples are selected and prepared along the material directions of
wood: boards must be prepared from split logs,
- discs are cut from the boards using a bell sawn,
- two coats of epoxy resin are applied on the lateral cylindrical face of each
specimen,
- the specimen is placed in the device using the vacuum pump,
- the first measurement point is performed by choosing one pressure gap and one
flow scale that allow both the pressure level and the gas flux to be accurately
determined,
- the previous step is repeated with several pressure gaps. The corresponding
fluxes are collected at steady-state, so that the linear relationship can be
checked,
- the permeability value is calculated by fitting a straight line to the data points
obtained for different level of pressure gap. This procedure reduces the
experimental noise and corrects automatically the offset values of the sensors.

7.3 Specific features of wood regarding permeability

7.3.1 Orders of magnitude: variability and anisotropy of wood

Fluid migration in wood uses the vascular system developed by trees for
physiological requirements. For this reason, wood has several specific features
concerning permeability, among which the two most important are:
- this material has dramatic anisotropy ratios: the longitudinal permeability can
be 1000 times greater than the transverse permeability for softwoods, and more
than 106 for hardwoods,
- heartwood is usually much less permeable than sapwood: pit aspiration, tyloses
development or extractives accumulation are responsible for this difference.

Table 7.1 summarises some values of directional permeability available in the


literature for different species. Depending on the experimental apparatus and the
Fluid migration in wood 131

protocol used by the authors, some data are missing in the papers. Indeed, it is not
always easy to calculate permeability ratios from permeability value, or vice versa.
Choong et al. (1974), for example proposes the permeability values for sapwood and
heartwood, for the longitudinal direction, but not for the transverse directions. Only
the mean anisotropy ratio is available in this paper. Perré (1992) and Perré et al.
(2002) use an experimental procedure able to determine the longitudinal
permeability and the anisotropy ratio on the same sample. In this case, we just
collected the ratios and decided not to calculate the transverse permeability
accordingly.

The variability is impressive, for the permeability values and for the anisotropy
ratios as well. Concerning longitudinal permeability, in spite of the scattering,
general trends are exhibited:
- the most permeable species in the longitudinal direction are among the ring
porous species (genus Quercus) that have very large vessels (up to 500 µm in
diameter),
- the pore diffuse hardwood are quite permeable too: probably the large number
of vessels can offset their smaller diameter (around 50 µm),
- softwood species are generally less permeable: gymnosperms have no specific
elements for sap flow, so the fluid has to path through the small opening, the
pits, at the end of each tracheid along the path, which means each 1 to 2
millimeters.

The transverse permeability and the anisotropy ratios are very variables. Note also
that the heartwood part of logs is usually much less permeable than the sapwood
part. This is due to tyloses development (Fig. 7.3) and extractives deposition
(tannins, gums…) in the case of hardwood and to the aspiration of bordered pits in
the case of softwood.

Fibers
T

Ray cells
L

Tyloses in
vessels

Figure 7.3. Example of tyloses development the heartwood part of Quercus robur
L. (Courtesy of Riad Bakour, LERMAB)
132 Fundamentals of wood drying

Table 7.1. Order of magnitude of permeability for some species along the three
material directions of wood

Authors Species Notes Permeability (m2) Anisotropy ratio


12 16 16
L·10 R·10 T·10 KL/KR KL/KT KR/KT
Populus sp. Sla 0.59 === 0.44 === 13 000 ===
Populus sp. Hla 0.61 0.15 0.18 40 000 33 000 0.835
Alnus rubra Sla 1.9 0.15 0.12 125 000 166 000 1.327
Choong Liquidambar sp. Sla 2.7 0.19 0.69 145 000 39 000 0.273
and Liriodendron sp. Hla 5.4 = 0.80 = 67 000 =
Kimbler Sequoia sp. Sla 4.9 11.2 19.4 4 000 2 000 0.580
1971 Sequoia sp. Hla 0.25 0.112 0.88 23 000 3 000 0.127
Pseudotsuga sp. Hla 0.026 0.000 0.000 = = =
Pseudotsuga sp Sla 0.049 = 0.37 = 1 300 =
Pinus sp. Hla 0.0026 0.17 = 153 =
Acer rubrum Sgo 10.3 = = 8 300 11 400 1.4
Acer rubrum Hgo 7.4 = =
Liriodendron sp Sgo 28.9 = = 1 450 1 600 1.1
Choong. Liriodendron sp Hgo 1.87 = =
et al. Liquidambar sp. S go 13.9 = = 1 300 2 750 2.1
1974 Liquidambar sp. Hgo 15.3 = =
Quercus rubra Sg 62 = = 13 000 18 000 1.4
Quercus rubra Hg 56 = =
Quercus falcata Sg 69 = = 110 000 143 000 1.3
Quercus falcata Hg 13 = =
Liriodendron sp Sl 26 4.5 = 57 000 = =
Liriodendron sp Hl 0.1 = = = = =
Chen Juglans nigra Sl 27 0.08 = 3.106 = =
et al. 1998 Juglans nigra Hl 0.0052 = = = = =
Quercus rubra Sl 61 0.68 = 900 000 = =
Quercus rubra Hl 45 = = = = =
Picea sp. ga 0.02 0.03 = 6 600 = =
Perré Pinus silvestris ga 0.07 0.42 = 1 600 = =
1987 Pinus pinaster ga 0.15 8.6 = 170 = =
Picea sp. Sga 0.2 = = 700 = =
Perré Fagus silvatica Sga 3.8 = = 3 000 65 000 21
1992 and
2002 Fagus silvatica Hga 1.4 = = 3 000 = =
Populus sp. Hga 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
Fluid migration in wood 133

7.3.2 Simple pore models applicable to wood: from fluid viscosity to Darcy’s
law

Due to its strong anisotropy ratio, a simple pore model can be applied to wood in the
longitudinal direction. In this model of porous medium, all pores are supposed to be
placed in parallel (Fig. 7.4). This model is defined by two independent parameters:
- the equivalent radius < r > ,
- the tortuosity factor τ, ratio squared of the microscopic distance ℓ over the
apparent macroscopic distance L.
The macroscopic transfer properties can be deduced from these morphological
parameters by a simple expression (Mason and Malinaukas 1983):

ε  r 
2
K= (7.5)
τ 8 
 

ε is the porosity of the porous medium, τ the tortuosity factor and r the
equivalent pore radius.

<r>

Figure 7.4. A simple model of porous medium usable for wood

In equation (7.5), the equivalent pore radius squared is involved. Indeed, the Hagen-
Poiseuille equation applied to one single tube tells us that the power four of the tube
radius appears in the volumetric flow rate. On the other hand, at constant porosity,
the number of capillary per section area in the porous medium is proportional to the
radius squared, hence the final power two in the permeability expression.

Such a bundle of capillary tubes placed in parallel is well adapted to hardwood


species. In the case of beech wood for example, the value of the longitudinal perme-
ability calculated from the vessel diameters, as determined by image processing, is
higher than the measured permeability, but only by a factor 2 to 4 in sapwood (Perré
and Karimi 2002).
134 Fundamentals of wood drying

In the case of softwood, this model has to be improved. Instead of one constant
diameter, large tubes standing for the cell lumens are placed in series with a certain
number of small tubes, for the pits connecting tracheids together. The same model
applies for longitudinal and transverse flow, simply by changing the tortuosity
factor. The model proposed by Comstock (1970) already contains some of this
modelling approach. In particular, it was able to predict the order of magnitude of
the permeability ratio.

7.3.3 Pit aspiration

In softwoods, the openings between two tracheids are small valves called bordered
pits. Due to a local detachment of the secondary wall from the middle lamella, the
double cell wall takes the shape of the external part of a tore (Fig. 7.5). The external
diameter of this tore is in the range 10 to 20 µm depending on the position in the
annual growth ring (the diameter is smaller in latewood). The torus is in suspension
in the middle of this semi-tore, thanks to the margo: a net of microfibil, radially
oriented with openings up to some micrometers wide. In normal operation, the sap
flows from one tracheid to the next one simply by using these small openings
existing in the margo (Fig. 7.5a).

In case of gas invasion in one tracheid, the gas-liquid interface is blocked in the
margo due to capillary forces (see section 7.5). These forces are able to push and
press the torus against the opposite border of the pit (Fig. 7.5b). Then, hydrogen
bonds keep the torus in this position: the pit is now aspirated and impervious
(Fig. 7.5c).
Middle lamella

Capillary forces
Margo

Torus Position of the


≅ 10 µm air/liquid
meniscus
GAS GAS GAS
LIQUID

LIQUID LIQUID

Secondary wall

CONDUCTIVE BORDERED PIT PIT ASPIRATION ASPIRATED PIT

Figure 7.5. The mechanism of pit aspiration: a clever strategy to limit the damage
caused by any gas invasion due to injury or cavitation of the sap column

This subtle mechanism is vital for trees, but causes a lot of problems to the wood
material. Indeed, pit aspiration occurs as soon as water is removed from wood, for
Fluid migration in wood 135

example during heartwood formation. In particular, it is impossible to avoid pit


aspiration when drying softwoods in normal conditions.

Only freeze-drying, or changing the liquid phase for a solvent with low surface
tension before drying allow air-dried samples without pit aspiration to be obtained.
This kind of experiment permits the dramatic reduction of permeability to be
quantified, but it not appropriate for the industry. Table 7.2 proposes some results
available in the literature on different works able to quantify the effect of pit
aspiration on permeability. The values depend on the species and on the author, but
it can be noticed that air-dried samples with aspirated pits have permeability values
equal to some percent only (typical range 1-10%) of the samples without pit
aspiration. Due to thicker cell walls, smaller pit radii and more rigid structures, the
percentage of aspirated pits is much less in the latewood part (Siau 1984).

Table 7.2. Effect of pit aspiration on the longitudinal permeability of different


softwood species. Most samples come from sapwood
Authors Species Permeability (m2·1012) Ratio (%)
Green Solvent Air dried
wood dried
Bolton and Tsuga heterophylla - 7 0.1 0.1
Petty (1978) " " - 5 (l) 0.06 (l) 0.06
Comstock and Tsuga heterophylla 2.8 - 0.2 – 0.02 7 – 0.7
Cote (1968) " " 5.3 6 0.02 0.4
Pinus resinosa 5.3 6 0.24 4.2
Tesoro and Pinus silvestris 1.2 - 0.16 13
Choong (1976) Taxodium distinctum 0.83 - 1.6 ??
Fumoto et al. Pinus resinosa - 11.4 0.36 3.2
(1984) Pinus banksiana - 7.4 0.11 1.5
Ward (1986) Abies concolor (S) - 13 1.3 10
Abies concolor (H) - 0.8 0.6 75
Meyer (1971) Pseudotsuga sp. - 8 1.5 19
- 2 0.08 4

7.4 Deviation from the linear law

7.4.1 Laminar flow, turbulent flow

The linear relationship between the flux and the pressure gradient (Eqn 7.1) indeed
assumes that the flow is laminar at the pore level. Classically, one knows that the
transition between laminar and turbulent flow depends on the Reynolds number:
D vρ
Re = (7.6)
µ
where D is the capillary diameter, m the dynamic viscosity of the fluid, r its density
and v the average velocity in the tube (for a laminar flow, the average velocity in a
cylindrical pipe is half the maximum velocity).
136 Fundamentals of wood drying

For Reynolds number less than 2300, perturbations are damped out and the flow
remains laminar. Above this critical number, depending on different parameters
including the surface roughness, turbulence usually develops. The velocity profile
inside the tube changes from a parabolic profile for laminar flow to a flat profile for
a turbulent flow. At the macroscopic level, an exponential relationship exists
between the flux and the pressure gradient rather than a linear one in the case of
laminar flow.

To have a better idea whether this transition may appear or not in wood, just take
one example. Assuming beech wood to be represented in longitudinal direction by a
bundle of capillaries in series, each one having 60 µm in diameter. To calculate the
pressure gradient required to obtain Reynolds number equal to 2300, we need the
relationship between the pressure gradient and the volumetric flow rate in laminar
regime in a cylindrical pipe:

π D2 v π D 4 ∆P
Q= = (7.7)
4 128µ L
Combining (7.6) and (7.7), the pressure gradient required to attain the critical
Reynolds number for an air flow is of the order of 108 Pa/m, that is 10 bar per
millimetre! This results would have been ten times greater for liquid water. These
values are huge, by far greater than what is usually encountered during wood drying
or when measuring permeability values. Therefore, it seems that turbulence is not
likely to occur in wood.

However, the gradual failure of the linear relationship between the flux and the
pressure gradient as the flux increases is commonly observed. It appears for
Reynolds number in the range 1-10, about three orders of magnitude smaller than
the transition towards turbulence (Dullien 1992). Although the velocity field
remains stationary in these conditions, this deviation is explained by the distortion in
the streamlines, including re-circulations, when the fluid inertia becomes significant
compared with viscous forces.

Several expressions have been proposed in the literature to account for this
deviation. The Forchheimer equation is certainly the most known of these
expressions (Dullien 1992):
∆P
= α µv + βρ v 2 (7.8)
L
a is a reciprocal permeability and b the co-called inertia parameter.

Equation (7.8) is consistent with Darcy’s law for small values of velocity. Last
remark: during all permeability measurements performed by the author, the failure
of the linear relationship was never observed.
Fluid migration in wood 137

7.4.2 Slip flow

For a gas with mean free path λ, a slip flow exists at the solid wall. Its value depends
on the velocity gradient at the boundary (Fig. 7.6). For a cylindrical tube, the total
flux must be corrected from the expression assuming a pure viscous flow (Poiseuille
flow). In the case of our simple model of porous medium, the apparent permeability
reads:

ε  1 3 
K app = 2
 r + λ r  (7.9)
τ 8 2π 

2µ RT
with λ =
P Ma

Fluid at rest at
the solid wall With slip flow

Figure 7.6. For gas flow, the slip flow at the solid wall can increase the viscous
flow significantly

By using the physical values at 20°C, a numerical expression can be derived:

ε  2 1  −3

K app = 125r + 5.54 r  ⋅ 10 (7.10)
τ P
Relation (7.10) is similar to the approach proposed by Klinkenberg (cited by Siau
1984):

< r > = 0.04


( )
intercept of curve K = f 1 / P
(7.11)
slope
This linear relation between the apparent permeability and the inverse of gas
pressure works well for certain materials, like MDF (Perré and Agoua 2002) but
fails in wood when measurement are available over a wide range of measurement
pressure. In the case a curved shape rather than a linear relationship is observed
(Figures 7.7 and 7.8). In this case, the improved model (see section 7.3.2), with two
different pore sizes, must be used to fit the experimental data.
138 Fundamentals of wood drying

-12
1.5x10

Experiment
Double pore size model
Apparent permeability (m )
2

-12
1.0x10

-12
0.5x10

0
0 10 20 30 40 50 60 70
Patm/P

Figure 7.7. Increase of the apparent permeability, due to slip flow, as the average
pressure in the sample decreases (Scots pine in longitudinal direction). A double
pore size model is sued to fit experimental data

-15
2.0x10

Experiment
Double pore size model
Apparent permability (m )

-15
2

1.5x10

-15
1.0x10

-15
0.5x10

0
0 10 20 30 40 50
Patm/P

Figure 7.8. Increase of the apparent permeability, due to slip flow, as the average
pressure in the sample decreases (Scots pine in radial direction). A double pore size
model is sued to fit experimental data
Fluid migration in wood 139

7.4.3 Effect of sample length on permeability

Numerous studies proved that the permeability value, which is supposed to be an


intrinsic property of the porous medium, indeed depends on the sample length. This
observation is generally explained by a certain probability of connection of the pores
within the porous medium (Bramhall 1971, Meyer 1971, Trénard 1980, Siau 1984,
Kauman et al. 1994). Usually, the permeability variations are nicely fitted when
using an exponential dependence of permeability with length.

Figure 7.9 depicts the gaseous permeability measured at different sample lengths for
sapwood and heartwood. On average, the permeability value increases by a factor 2
for sapwood and by a factor 5 for heartwood when the length is reduced from 40 to
10 centimetres. The experimental data have been fitted using a classical exponential
dependence of permeability with length. The exponential factor is equal to –3.2 m-1
for sapwood and –6.5 m-1 for heartwood. This observation is consistent with other
experimental works: the decrease in the logarithm of the permeability with length is
higher for samples with low permeability values (Siau 1984). Thanks to this
difference in decreasing rate and in spite of the great permeability difference
measured on samples, the intercept with the y-axis, which represents the
extrapolated value for a zero-length sample, K0, is similar for both zones:
12.2·10-12 m2 for sapwood and 9.5·10-12 m2 for heartwood. It is also important to
notice that these values are reasonably close to the values predicted using formula
(7.5) (Perré and Karimi 2002).

12
Sapwood
Heartwood
10
Permeability (m .10 )
12

8
K = 12.2 exp(-3.2*L)
2

K = 9.5 exp(-6.5*L)

0
0 5 10 15 20 25 30 35 40
Sample length (cm)

Figure 7.9. Effect of sample length on gaseous permeability (Fagus silvatica,


sapwood and heartwood, average over 3 samples) (after Perré and Karimi 2002)
140 Fundamentals of wood drying

7.4.4 Percolation models

A more rigorous observation of wood permeability reveals surprising trends (Siau


1984, Perré and Karimi 2002):
- the actual sample permeability is much smaller than the value calculated
assuming that the wood sample is a bundle of capillary tubes (the vessels),
- the sample permeability decreases as the total sample length increases,
- the number of active vessels decreases both when the sample length increases
and when the distance from the injection point increases (Fig. 7.10).

Only statistical models can account for these outcomes. In such models (Adler 1992,
Dullien 1992), the porous medium is assumed to be made up of bonds having a
certain conduction probability (Fig. 7.11a). In order to account for wood anisotropy,
this 2-D model considers two different connection probability values: one along the
longitudinal direction and the other in the transverse direction. When injecting the
fluid in one part of the external surface, only the connected pores act for the liquid
migration (Fig. 7.11b). But, among these connected connections, the flow is
effective only in a few of them. This is the concept of active pores, those for which
the flow rate exceeds a threshold value, which requires the pressure field within the
sample to be solved (Fig. 7.11c).

The theoretical permeability value can be deduced from this dimensionless flux
together with the zero-length permeability value, as calculated from anatomical
cross-sections. In this approach, consistently with wood anatomy, the difference
between sapwood and heartwood only lies in the probability values, not in the zero-
length permeability. In this simulation, the vessels, composed of many vessel cells,
are supposed to have an averaged length of two centimetres (Zimmerman 1983).

Although this concept has been developed for a 2-D network here, it allowed all the
reported steady-state observations to be predicted (Perré and Karimi 2002):
- the effect of sample length on permeability,
- the variation of the active pores fraction within the sample and with the sample
length,
- the differences between gaseous and liquid permeability or between sapwood
and heartwood,
- the apparent deviation of the permeability calculated on cross-sections from the
actual value measured on real samples.

This concept of percolation can explain other important macroscopic observations,


such as the existence of a thin dry shell at the exchange surface of boards during
drying, namely in the case of softwoods (Wiberg et al. 2000, Rosenkilde and Glover
2002, Rémond et al. 2005). Recently, the invasion percolation concept has been used
to analyse this mechanisms during drying (Salin 2005).
Fluid migration in wood 141

a
a

Figure 7.10. Microphotograph of a beech sample penetrated by dyed water: only a


small percentage of vessels were active (a = active vessel) (after Perré and
Karimi 2002)

(a) Conducting bonds (b) Connected bonds (c) Active bonds

Figure 7.11. The random networks to explain the liquid flow in wood. The
probability for a conducting bond is 0.8 in the longitudinal direction and 0.2 in the
transverse one. The thick line is the zone of fluid injection (after Perré and
Karimi 2002)

7.4.5 Multiscale effects

Finally, among deviations from Darcy’s law, multiscale effects are worth a mention.
Because some anatomical wood structures are strongly heterogeneous, the behaviour
of a sample in transient state might be very different that stipulate by the steady-state
142 Fundamentals of wood drying

laws. One example of such a mechanism is presented hereafter. It concerns re-


saturation of beech samples in the longitudinal direction. Experiments are available
in which the moisture content gradient is measured in the sample using x-ray
attenuation (Perré and Thiercelin 2004). During re-saturation, the moisture content
increases quickly very close to the cavity, but requires a very long time for the
remaining part of the sample to absorb moisture (Fig. 7.13a). For this configuration
and this material, the macroscopic approach fails: using macroscopic laws and
reasonable parameters for the permeability and the capillary function, the predicted
imbibition time is smaller than for the experiment (Fig. 7.13b).

In a previous study (Perré and Karimi 2002), the injection of dyed water into beech
samples allowed us to observe that a very small percentage of vessels are active
(Fig. 7.10). This fact can be explained by a certain probability of connection
between different lines of vessel cells. Based on this small percentage of active
vessels, a double scale interpretation of the mechanisms that govern moisture
migration in beech was proposed. The vessel network represents an easy pathway
for liquid water, whereas it is very difficult for the water to wet the fibre zones (Fig.
7.12). The medium is therefore out of local equilibrium during most of the
experiment.

Consequently, a dual-porosity approach is proposed. The computational domain


uses a 2-D axisymmetric configuration for which the axial coordinate represents the
macroscopic longitudinal direction of the sample whereas the radial coordinate
allows the slow migration from each active vessel towards the fibre zone to be
considered (Perré 2004). The latter is a microscopic space variable.

The moisture content field evolution depicts clearly the dual scale mechanisms: a
very fast longitudinal migration in the vessel followed by a slow migration from the
vessel towards the fibre zone. The macroscopic moisture content field resulting from
this dual scale mechanism is in quite good agreement with the experimental data
(Fig. 7.13c).

Slow migration in the fibre zone

Fast liquid migration in the vessel

Figure 7.12. Re-saturation of beech: physical interpretation of the dual-scale


moisture migration mechanisms
Fluid migration in wood 143

1 hour
a) 2
Moisture content increase (kg/m )
6

13 days

3 days
5 hour
2

8 minutes

0
0 10 20 30 40

Distance (mm)

1.5

b) 60 minutes

1.0
Moisture content (%)

1 minute 5 minutes 20 minutes 40 minutes

0.5

0
0 5 10 15 20
Length (cm)

1.5

c) 40 hours

20 hours
Macroscopic moisture content (%)

1.0

10 hours
5 hours

2 hours
0.5

30 minutes

0
0 1 2 3 4
Length (cm)

Figure 7.13. Moisture profiles during re-saturation of a beech sample in


longitudinal direction. a) Profiles measured by x-ray attenuation (Perré and
Thiercelin 2004); b) Profiles predicted using a macroscoscopic formulation
(Darcy’s law and capillary function); c) Profiles predicted using a dual-scale
modelling (Perré 2004)
144 Fundamentals of wood drying

7.5 Capillary pressure

7.5.1 Surface tension: observation, formulation

During drying, two immiscible fluid phases co-exist in the porous medium: a liquid
phase and a gaseous phase, hence the existence of an interface between these phases,
which property is very important for mass transfer in unsaturated porous media.

The molecules in the interior of a liquid experience interactions with other


molecules equally from all sides. So, the cohesive forces between molecules inside a
liquid are shared with all neighbouring atoms. Those on the surface have no
neighbouring atoms above, and exhibit stronger attractive forces upon their nearest
neighbours on the surface. This enhancement of the intermolecular attractive forces
at the surface is called surface tension σ. This forms a surface “film” which behaves
like an elastic skin under tension. Systems with an interface between two fluids tend
to decrease the area of this interface. When the surface tension is the predominant
force acting on the system, the surface area at equilibrium minimises the area. This
is why small water droplets (so the gravity force can be neglected) or Champagne’s
bubbles have a spherical shape.

Surface tension is typically measured in N·m-1, the tensile force in Newton acting on
a unit length of interface. It can also be defined as the energy necessary to increase
the surface by one square metre (J m-2, which is equivalent to N·m-1). The value of
the surface tension depends on both fluids in contact and depends on temperature.
The water/air interface at 20°C has a surface tension of 72.75 10-3 N·m-1 compared
to 22.6 10-3 for methanol/air and 472 10-3 for mercury/air (Atkins 1994).

The effect of temperature is almost linear in the range 0°C to 100°C (Fig. 7.14):

σ = (76.06 − 0.1676 T ) ⋅10 −3 N ⋅ m −1 (T in °C) (7.12)

Due to this layer under tensile stress, a difference of pressure between the two fluids
exists as soon as the interface has a curvature. At any point P of the surface, this
pressure gap can be expressed using Laplace’s law:

1 1 
∆P = σ  +  (7.13)
 r1 r2 

r1 and r2 are the two principal radii of curvature of the surface and σ is the surface
tension. Note that the surface curvature can be calculated using the radii of curvature
R1 and R2 of any pair of orthogonal lines (intersection with the surface of two
orthogonal planes containing the normal to the surface in P). The mean curvature of
the surface at this point 1/rm, can therefore be defined as follows:
Fluid migration in wood 145

1 1   1 1  2
 + = + = (7.14)
 r1 r2   R1 R2  rm

From equation (7.14), it is evident that the mean curvature of a sphere of radius r is
simply 1/r. Finally, one has to consider that curvatures may be positive or negative.
At a saddle point, for example, the mean curvature is equal to zero because two radii
of curvature, along orthogonal directions, have the same magnitude but opposite
signs. The same kind of opposite effect exists in the inner part of a tore, or around
cylindrical stick put in water.
80

Data from (Lide, 1995)


Fitted linear curve
Surface tension (water-air) (mN/m)

75

70

65

σ = (76.06 − 0.1676 T).10 −3 N.m −1 (T in °C)


60

55
0 20 40 60 80 100
Temperature (°C)

Figure 7.14. Surface tension of the water-air interface as a function of temperature:


data from the Handbook of Chemistry and Physics (Lide 1995)

7.5.2 The case of three-phase system

In the porous medium, the solid phase is always present. So, two fluid phases and
one solid phase coexist. Surface tensions can be defined for each interface: gas-
liquid, liquid-solid and solid-gas. At the intersection between these three surfaces,
the mechanical equilibrium can be satisfied only at a certain value of the angle θ
(Fig. 7.13). Angle θ can take any value between 0° and 180°. If θ is less than 90°,
the liquid phase is the wetting phase, if θ is greater than 90°, the liquid phase is the
non-wetting phase (typical situation with mercury). If the angle is equal to zero
( σ ℓg ≤ σ sg − σ sℓ ), liquid will spread spontaneously on the surface of the solid.

Several methods have been developed to measure this angle: tilting plate method,
sessile drop, sessile bubble, contact angle cell (Dullien 1992). However, it is
important to be aware that this angle has not a unique value: the advancing angle
146 Fundamentals of wood drying

(the value measured when the liquid replaces the air) is usually much larger than the
receding angle (the value measured when the air replaces the liquid). Therefore, the
experimental determination of the contact angle is very difficult and subject to
uncertainty. This phenomenon, known as angle hysteresis, can be explained by
surface roughness or by physicochemical mechanisms. It has important implications
on the capillary pressure in porous media.

a) b)

σℓg
GAS GAS
SOLID θ
σsg LIQUID
σs ℓ
θ
SOLID
LIQUID

Figure 7.15. Two configurations that allow the contact angle θ to be observed:
a) meniscus in a capillary tube, b) a drop of liquid placed on a solid surface

In the case of a cylindrical capillary tube (Fig. 7.15a), equation (7.13) leads to:
2σ cosθ
∆P = (7.15)
r
Equation (7.15) supposes that the solid walls are parallel. In the case of a conical
tube the equation should be modified accordingly:

2σ cos(θ + ϕ )
∆P = (7.16)
r
ϕ stands for the semi-angle of the conical tube.

In equation (7.15), r is the radius of the capillary tube. In the case of a perfectly
wetting liquid (θ = 0), equation (7.15) is even simpler:


∆P = (7.17)
r
The pressure is always lower on the convex side of the interface. For example, if the
liquid phase is the wetting phase, the liquid pressure is lower than the gaseous
pressure in a capillary tube. This is why the meniscus inside a tube rises above the
surface (Fig. 7.16). The height of capillary rise is simply tied to the pressure gap via
the classical result from the static of fluids:
Fluid migration in wood 147

∆P 2σ cosθ
∆H = = (7.18)
ρℓ g rρ g

ρℓ is the liquid density (kg·m-3) and g the acceleration of gravity (m·s-2).

H2
H1

Figure 7.16. Effect of the tube radius on the value of the capillary pressure

The configuration schematised in Figure 7.16 is the typical situation encountered


within wood. Inversely, the pressure is higher on the concave side of the interface:
the liquid pressure inside a droplet is higher than the pressure of the surrounding air.

7.5.3 Vapour pressure at the surface of a meniscus

Due the curvature of the interface, a deviation of the saturated vapour pressure
exists. This deviation can be calculated from the definition and properties of the
Gibbs free energy: the method is similar to the one used to derive the Clausius-
Clapeyron equation, with a change of liquid pressure at constant gaseous pressure:

P  1 1 M
ln  v  = −σ  +  v
(7.19)
 Pvs   r r  ρ RT
 1 2 ℓ
Pvs is the saturated vapour pressure (case of a plane interface) and Pv is the
equilibrium vapour pressure at the curved interface, r1 and r2 are the two principal
radii of the surface and Mv is the molar mass of the water vapour.
Table 7.3 proves that this effect remains negligible for capillary tubes whose radius
is larger than 1 µm, whereas the pressure gap is already very important. Below
0.1 µm, the deviation from the saturated vapour pressure becomes significant. The
last part of the sorption isotherm curves measured for wood is usually explained by
this effect (capillary condensation in the micro-pores of the cell walls).
148 Fundamentals of wood drying

Table 7.3. Pressure gap and relative humidity at the surface for different values of
tube radii (values calculated at 20°C, for a perfectly wetting liquid and cylindrical
tubes)
Radius of the ∆P Pv/Pvs
capillary tube
1 mm 146 Pa 0.999999
100 µm 1456 Pa 0.999989
10 µm 14 560 Pa 0.99989
1 µm 1.44 atm 0.9989
0.1 µm 14.4 atm 0.989
0.01 µm 144 atm 0.898

7.5.4 The capillary pressure in porous media

As already stated, unless the porous medium is fully saturated with one of the fluid
phases, numerous menisci exist in the medium at the pore level. At the macroscopic
level, a capillary pressure may be associated to the “effective” pore diameter where
the menisci are located. The “effective” pore radius decreases as the liquid
saturation (wetting phase) decreases. This dependency is usually called the capillary
pressure function. The saturation is defined as the proportion of pore volume filled
with a fluid phase. Therefore, a liquid and a gaseous saturation can be defined. Their
sum is equal to the unit if only two fluid phases exist in the porous medium.
However, unless otherwise specified, the saturation S, is usually defined for the
wetting phase:

Volume of pore occupied by the wetting phase


S= (7.20)
Total pore volume
In the case of wood, this parameter has a physical meaning only in the domain of
free water: bound water should not be included in this definition:

Volume of pore occupied by free water


S ( in wood ) = (7.21)
Total pore volume at the Fiber Saturation Point

Several methods can be used to determine the capillary pressure curves (Dullien
1992):
- The drainage method: the porous medium is initially fully saturated with the
wetting phase and the non-wetting fluid is used to reduce the saturation, by
increasing very slowly its pressure. Mercury porosimetry is a typical and widely
used example of drainage method.
- The drainage column: for porous media with relatively low capillary pressure
level, the gravity can be used to obtain a gradient of equilibrium saturation due
to the shape of the capillary pressure function.
Fluid migration in wood 149

- The centrifuge method: using similar initial conditions, centrifugal forces are
used to reduce the saturation of the wetting phase. The saturation profile,
determined at equilibrium, is connected to the centrifugal force via the
rotational velocity and the distance to the rotation axle squared. This relation
allows the capillary pressure function to be established.
- The imbibition methods: the specimen is initially fully saturated with the non-
wetting phase. Then, the three previous configuration presented above can be
used to determine the capillary function curve during imbibition.
- Assuming that the wetting fluid present in the medium is located in such a way
that the energy level is minimised, the smallest possible pores will be filled.
This rule gives a unique capillary function curve. This curve can be calculated
from the knowledge of the pore distribution of that porous medium. Usually,
this distribution is calculated by image processing on 2-D (cross-sections or
polished faces) or 3-D (microtomography) views of the porous medium.

DRAINAGE IMBIBITION

Figure 7.17. A simple sketch to explain the hyteresis of the capillary pressure
function due to the variation of pore diameter along the pathway

All these procedures should produce the same, ideal, capillary pressure function.
Unfortunately, the reality is not that simple. In real configurations, the allocation of
the wetting phase can differ from the ideal distribution, which corresponds to the
minimum of the wetting phase energy. Initial conditions and boundary conditions
usually produce distributions out of equilibrium (at least over reasonable observation
times, see Dullien 1992). This is why capillary functions determined using
imbibition and drainage experiments are different. The hysteresis of the contact
angle is partly responsible for this observation. However, the classical explanation
for the hysteresis of the capillary function lies in the pore morphology of porous
media. Along the fluid pathway, the effective pore radius varies, sometimes
considerably:
- a large pressure is required for the non-wetting phase to drain the wetting phase:
the meniscus has to path through the smallest radii of the pathway,
150 Fundamentals of wood drying

- a much lower pressure is enough to prevent the wetting phase to invade the
medium during the imbibition process: the meniscus will be blocked in the
largest part of the pathway.
This mechanism is known as the bottleneck effect (Fig. 7.17). It explains, for
example, why it is so difficult to completely re-saturate a sample of wood.

A few papers can be found on the determination of capillary pressure functions in


wood. Mercury porosimetry exhibits a dramatic effect of the sample thickness in the
longitudinal direction (Trénard 1980). For short samples, the cell lumens are directly
accessible to the mercury whereas for longer samples, the liquid has to path through
the small openings, the pits that exist between cells: a clear illustration of the
bottleneck effect. The centrifuge method has been used successfully by Spolek and
Plump (1981) on softwoods and Choong and Tesoro (1989) on various species. To
determine the moisture content-water potential relationship of wood, Cloutier and
Fortin (1991) used a tension plate and a pressure plate. This is a drainage method
and their results could easily be converted into classical capillary pressure function.

Finally, because the morphology of wood pores is governed by anatomical features,


particular methods can be applied to wood:
- The geometrical model of the tracheid shape proposed by Comstock (1970), has
been developed by Spolek and Plump (1981) to compute a capillary pressure
curve. Although it may be simplistic to assume that all tracheids have exactly
the same shape, they obtained a good trend for the capillary pressure function.
- Because the longitudinal direction of wood is a very particular direction, it is
quite simple to path from a cross-section to the three-dimensional structure of
the material. Figures 7.18 and 7.19 depict examples of capillary pressure curves
calculated from microscopic images of cross-sections of wood. In this case, the
pore size distribution has been calculated using image processing (Perré 1997,
Perré and Turner 2001a).
Fluid migration in wood 151

0.4

Late wood
Middle wood
Capillary pressure (Pc/Patm)

0.3 Early wood

0.2
3
641 kg/m
3
458 kg/m

0.1

3
349 kg/m

0
0 0.2 0.4 0.6 0.8 1.0

Saturation
Figure 7.18. Capillary function curve determined in different parts of spruce wood
using image analysis. The cells have thicker walls and smaller radial extension in
latewood part, hence the highest value of the capillary pressure curve
0.8

Late wood
Capillary pressure (Pc/Patm)

0.6 Early wood

656 kg/m3
0.4

543 kg/m3
0.2

0
0 0.2 0.4 0.6 0.8 1.0

Saturation
Figure 7.19. Capillary function curve determined for beech wood using image
analysis. For this pore diffuse hardwood species, earlywood and latewood look
similar. The low pressure obtained for saturation values above 0.2 correspond to the
meniscus radii located in the vessel elements. The dramatic increase for low
saturation is due to the small lumen diameters of the parenchyma and fibre cells
152 Fundamentals of wood drying

7.6 Generalised Darcy’s law: multiphase flow

7.6.1 Observation, formulation

In section 7.2, the permeability has been defined with one single fluid phase present
in the porous medium. When two phases co-existent, an extended formulation, the
generalised Darcy’s law, must be used. The volumetric flow rate of each phase is
considered to be proportional to the pressure gradient of the corresponding phase.
The phenomenological coefficient is the product of the permeability K by a function
of saturation called “relative permeability” to the considered phase.

For the gaseous phase

K krg ( S )
vg = − ∇Pg (7.22)
µg

For the liquid phase

K krℓ ( S )
vℓ = − ∇Pℓ (7.23)
µℓ
The liquid pressure is tied to the gaseous pressure through the capillary pressure
function:

Pℓ = Pg − Pc ( S ) (7.24)

Obviously, equations (7.22) and (7.23) must be consistent with Darcy’s law when
one single fluid phase occupies the porous medium. Consequently, the relative
permeability functions fulfil the following conditions:

Gas only Liquid only


k rℓ (0) = 0 k rℓ (1) = 1 (7.25)
k rg (0) = 1 k rg (1) = 0

As for the capillary pressure curve, the relative permeability is usually supposed to
be function of saturation only. However, the same problems exist for these
parameters. Because initial conditions and boundary conditions lead to different
distributions of the wetting phase, the relative permeability values will change
accordingly and have different values for the same averaged saturation.

However, even with the simple assumption that the relative permeability curves
depend only on saturation, their experimental determination remains very
challenging. The initial conditions resembles those used to determine the capillary
pressure function, but the boundary conditions are such that permanent flow rates of
each phase is attained in steady-state (both co-current or counter-current
Fluid migration in wood 153

configurations are used). Different values of imposed fluxes (or pressure) allow the
relative permeability to be determined at different saturation values.

Very few results are available in the literature for wood. The works published by
Tesoro and co-workers are among the seldom-available results (Tesoro et al. 1972,
1974). Note also that a more recent paper by Tremblay et al. (2000a) proposes an
indirect method to measure this property.

Based on the geometrical model of tracheids proposed by Comstock (1970), Spolek


and Plumb (1980) derived an expression for the relative permeability curves in
softwoods. As for the capillary pressure, this model assumes that all tracheids are
exactly similar, and that the wetting phase distribution is ideal. Their final
expression involves an irreducible saturation, below which no liquid flux is possible.

This concept of irreducible saturation has been widely discussed, especially in the
active domain of petroleum production (Dullien 1992). Depending on the boundary
conditions, one part of the fluid phase that occupied the medium at the beginning of
the experiment remains in the medium, even though its flow rate vanished. This part
seems to be trapped in the medium, in what is called a “pendular” state: just imagine
blobs that are trapped because isolated. The amount of the residual part increases
with the imbibition or drainage velocity. In addition, time seems to be able to reduce
the amount of trapped phase after the experiment: surface spreading in the solid
phase and the edge capillary pressure are the two mechanisms able to explain this
observation (Dullien 1992). Due to the microporosity that exists in the cell walls,
such mechanisms are likely to exist in wood. In addition, the concept of irreducible
saturation leads to unrealistic computed moisture content profiles during drying
(Perré 1987).

For all these reasons, the concept of irreducible saturation should be discarded for
wood. Based on the measurements in the longitudinal direction reported by Tesoro
et al. (1972), on the curve computed from the tracheid model (Spolek and Plumb
1980) and on these considerations regarding the concept of irreducible saturation,
the following functions have been proposed for wood by Perré et al. (1993)
(Fig. 7.20):

In the transverse direction (radial or tangential)


T
k rg = 1 + (2S − 3) ⋅ S 2 and k rTℓ = S 3 (7.26)

In the longitudinal direction


L
k rg = 1 + (4S − 5) ⋅ S 4 and k rLℓ = S 8 (7.27)
154 Fundamentals of wood drying

1
Transverse
Longitudinal
0.8
Relative permeability

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
Saturation

Figure 7.20. Relative permeability curves calculated using expressions (7.26)


and (7.27)

7.6.2 Migration mechanisms during drying

Equations (7.22) to (7.24) forms the macroscopic formulation of the complex


physics of multiphase flow in porous media that takes place during drying.

The first important effect is the well-known capillary migration (Fig. 7.21). If liquid
water is removed at a certain location of the medium, by evaporation for example,
the liquid-gas interface is obliged to recede towards zones with smaller pore radii.
The capillary pressure increases at this point. Because the liquid water is the wetting
phase, the liquid pressure is reduced (Eqn (7.24)). A liquid pressure gradient
appears, which drives liquid from other zones towards this particular point (Eqn
(7.23)).

GAS Evaporation

SOLID

Smaller meniscus
radius
WATER

Capillary migration

Figure 7.21. Principle of capillary migration


Fluid migration in wood 155

Another important effect embedded in these equations is the mechanism of mass


migration induced by a gradient of total gaseous pressure. This effect comes across
in all processes with internal overpressure, such as vacuum pressure, radio
frequency drying, high temperature drying…

The gaseous pressure gradient is the driving force for bulk migration of gas (Eqn
(7.22)). Consequently, provided the relative permeability to gas is not equal to zero,
any gaseous pressure gradient gives rise to a gas flux. If the gaseous phase is made
up mostly of water vapour, such a bulk migration is very efficient to drive moisture
from the material.

Even more efficient is the possibility of liquid migration induced by a gaseous


pressure gradient. Indeed, through equation (7.24), any pressure gradient induces a
liquid gradient, which can drive liquid through equation (7.23). Obviously, the
efficiency of this pressure gradient depends directly on the permeability K, which is
a constant intrinsic parameter, and on the relative permeabilities to each fluid phase,
that depend on saturation (Fig. 7.20):
- for high values of moisture content, the permeability to the liquid phase is high
(Fig. 7.22). The gaseous pressure acts mainly on the liquid interfaces to drive
liquid,
- for low values of moisture content, the liquid saturation S is low, or even equal
to zero (hygroscopic domain). The effect of gaseous pressure is mainly, or
exclusively, on bulk migration of the gas phase.

GAS
SOLID # K """"#
vℓ = − ℓ grad(Pℓ )
WATER
µℓ

""""#
grad(P)
GAS
SOLID

# K g """"#
vg = − grad(Pg )
µg

Figure 7.22. Effect of a gradient of total gaseous pressure on the fluid migration in
a porous medium
156 Fundamentals of wood drying

In this last section, we just started to describe the dynamic of fluid flow occurring
during drying: in fact we shifted from a single phenomenon to coupled phenomena
that will be analysed in detail in Chapter 10.

7.7 Conclusion

This chapter presented the physical mechanisms involved in the migration of fluid
(gas and liquid) in wood. It explains the phenomena of fluid migration in the case of
single fluid phase in the porous medium. The most important deviations from
Darcy’s law are shortly described: slip flow, effect of sample length on permeability
and percolation effects. Finally, the concepts of capillary pressure and multiphase
migration are presented. All these features, general phenomena and deviations from
the general laws, are more or less present during wood drying, namely when the
moisture content is high.
8 Creep deformation in drying wood
Antti Hanhijärvi
Senior Research Scientist
VTT Building and Transport,
P.O.Box 1806, FIN-02044 VTT, Finland
E-mail: Antti.Hanhijarvi@vtt.fi

ABSTRACT

The paper reviews the significance of creep deformation to the drying process of wood and
some characteristics of creep in wood. Creep in drying wood occurs as a consequence of the
internal stresses within the wood caused by shrinkage deformation due to drying. The
prominent significance of creep during drying is that it releases drying stresses and thus
reduces the amount of cracking. The amount of creep is dependent on prevailing conditions,
temperature and moisture content. The fact that creep is accelerated at higher temperatures
enables faster drying of timber without increased amount of cracks, and accounts partly for
the effectiveness of high temperature drying. Two different mechanisms of creep are observed
in wood: the simple viscoelastic creep and the mechanosorptive creep. Simple viscoelastic
creep occurs as delayed, time-dependent, deformation under stress. Mechanosorptive creep
occurs as additional deformation caused by simultaneous action of stress and moisture
content changes.

KEYWORDS: timber, drying, creep, time-dependent deformation, viscoelasticity,


mechanosorptive creep

Fundamentals of wood drying, Patrick Perré editor, pages 157 to 174.


158 Fundamentals of wood drying

8.1 Introduction

Drying wood undergoes various physical phenomena as its moisture content


decreases. Drying below fibre saturation induces shrinkage, i.e. a deformation
caused by the reduction of the amount of water in the cell walls, the so-called bound
water. The shrinkage deformation causes an internal stress field within the wood,
which in turn induces other, stress-driven, deformation phenomena. These include
elastic and creep deformations. Although hygroexpansion, shrinkage or swelling,
caused by moisture changes, is the most important factor for internal stress
generation in drying wood, they can be caused by other phenomena, like thermal
expansion and hygrothermal deformation. From the practical point of view,
shrinkage and shrinkage stresses are important as they are high enough to cause
cracking, which may dramatically diminish the value of dried timber. The
importance of creep for timber drying comes from the fact that creep deformation
releases shrinkage stresses, and if the amount of creep deformation is sufficient, the
stresses are reduced enough so that cracking is avoided. Creep deformation can also
reduce the amount of warping and twisting of boards during drying, if the
deformation of the boards is restrained during drying.

The objective of the present paper is to review the basics of the creep deformation in
wood material during drying and give a short review of literature published. The
content is limited to consider only transverse material directions (radial and
tangential directions) as these are the decisive for drying crack formation.

8.2 Importance of creep for drying crack formation

During the first stages of drying, the surface layers of the board lose water and dry
below fibre saturation, and begin to shrink. At this point an internal stress field is
created inside the board such that tensile stress is created near the surface and
compression stress at the interior (Fig. 8.1). This stage is usually decisive for crack
formation: if the drying of the surface is too fast, the differential shrinkage between
the surface layers and the interior grows large, and consequent stresses get high and
generate cracks on the surface. However, if the drying speed is less, the differential
shrinkage between surface and middle layers does not rise too high, the stresses are
kept lower and cracking limit is not exceeded. What then is a critical value for
drying speed that does not cause cracking? It depends largely on the amount of creep
that occurs in the wood. Viz., depending on conditions, the stresses generated in the
board induce creep deformation, which relaxes the stresses and lowers the risk of
cracking. Creep deformation is dependent on conditions and it is much higher at
high temperature and wet conditions than at low temperature or dry conditions.
Therefore high temperature relaxes stresses more effectively than does the medium
or low temperature range. This is one of the reasons, why drying speed in high
temperature drying can be faster than in low temperature drying and still produce
less cracking.
Creep deformation in drying wood 159

a) b)
Tension Compression

Compression Tension

Figure 8.1. Schematic of the stress state in a board section. a) In the early part of
the drying. b) In the late part of drying

Later on, when drying proceeds, and the moisture content difference between the
surface and middle layers is evened out, the stress distribution inside the board may
turn the opposite: compressive stresses are generated near the surface and tensile
stresses in the middle. This phenomenon is caused by the fact that the creep that
occurred in the surface under high tensile stress has caused an ‘elongation’
compared to the nearly unstressed (in fact slightly compressed) wood in the middle
part, which just undergoes more or less normal shrinkage deformation. The reversed
stress state does not normally induce cracking during drying, however, in fast high
temperature drying these stresses can induce internal cracking.

Creep occurs in all material directions of timber. The natural choice for the
coordinate system for wood is the cylindrical RTL-coordinates system, where R
denotes the radial direction, T – tangential and L – longitudinal, Fig 8.2. For drying
applications, creep in the transverse (RT) plane is of concern, when considering
cracking. For twist and warp, the longitudinal direction is important, but not covered
in this paper.

R
T
L

Figure 8.2. The natural choice of coordinate system for describing material
directions in wood, cylindrical coordinate system with directions R - radial, T -
tangential and L - longitudinal
160 Fundamentals of wood drying

8.3 Definition of creep and creep mechanisms

Creep in general is considered as stress driven deformation rate. This means that
creep is rate-dependent and creep deformation requires time to occur. On the
contrary, rate-independent deformation mechanisms include elastic and plastic
deformations, which are considered to occur instantaneously. In the case of wood, it
is sometimes difficult empirically to distinguish between instantaneous (elastic) and
rate-dependent (viscoelastic creep) deformation (Hanhijärvi 1999), since loading can
not be made instantaneously. However, from both the theoretical and practical point
of view, it is convenient to distinguish between the two concepts and recognise the
instantaneous character of elastic deformation and the time-dependent nature of
creep.

Creep in wood resembles in many ways the creep behaviour observed in synthetic
polymers. This is no surprise, since wood is a complex naturally grown arrangement
of different polymers, the main constituents being cellulose, hemicellulose and
lignin. Wood retains its properties from both its constituents as such but also from
the arrangement of them in the anatomical and ultrastructural sense. This holds true
for all mechanical properties and therefore naturally also for creep. Creep of wood is
of viscoelastic character, which is typical to polymers. This means that creep
deformation recovers, if the load is removed, although the recovery may be slow
(Fig. 8.3). Viscoelastic nature of creep also implies that under constant conditions
and constant load, creep rate decreases. Creep curves are often presented on
logarithmic time scale which leads to curves with increasing slope as shown in
Figure 8.4. Creep of wood is also strongly anisotropic, i.e. its magnitude depends on
the direction of the acting stress, which is a consequence of the arrangement of the
constituents in the wood internal structure in a very oriented way.

Creep is an important term in the deformation of drying wood. The amount of creep
can in many circumstances exceed manifoldly the amount of elastic deformation and
creep plays the practically decisive role in relaxing shrinkage stresses. Calculations
have shown that, if creep was absent and wood behaved only elastically, it would be
impossible to dry wood without causing severe cracking (e.g. Ranta-Maunus and
Kortesmaa 1988).
Creep deformation in drying wood 161

Spruce S, tangential, 95 & 125°C


0.25
125°C
0.2
Compliance [1/MPa]

0.15

0.1

0.05 95°C

0
0 50000 100000 150000 200000
Time [s]

Figure 8.3. Experimental creep curves of wet spruce (Picea abies) specimens
(Hanhijärvi 1997) at high temperature conditions. The particular features of
viscoelasticity are clearly visible: the eventually decreasing creep rate and (partial)
recovery of creep deformation after load removal
162 Fundamentals of wood drying

Spruce S, tangential, 95 & 125°C


0.25

0.2
Compliance [1/MPa]

125°C
0.15

0.1

0.05
95°C
0
1 10 100 1000 10000 100000 1000000
Time [s]

Figure 8.4. The creep curves of Fig. 8.3 presented on logarithmic time scale
(Hanhijärvi 1997)

8.3.1 Different mechanisms of creep

For most creeping polymeric materials, creep rate is a function of environmental


conditions (temperature etc.) and past stress history. This is true for wood also:
creep rate is dependent on the prevailing temperature and moisture content. Wood
exhibits the peculiar behaviour that creep rate is also dependent on the change rate
of one state variable, viz. the moisture content change rate. This effect is denoted as
the mechanosorptive effect. Therefore one can distinguish two types of creep
deformation mechanisms in wood: the viscoelastic and the mechanosorptive. To
make an exact definition of the two types of creep is slightly complicated
(Hanhijärvi and Mackenzie-Helnwein 2003), but in practice the two types can be
characterised as follows: (1) viscoelastic creep is the part of the creep deformation
that occurs in constant conditions as stress driven deformation rate, and (2)
mechanosorptive creep is the extra creep that is generated due to changing moisture
content.
Creep deformation in drying wood 163

8.4 Review of experiments

8.4.1 Experimental techniques for measuring creep in drying wood

Measurements for determining creep in wood in drying conditions are challenging.


The first challenge is that the condition range that wood undergoes during the
complete drying process is very wide: both the temperature levels reached are wide
and moisture contents change strongly. If one wants to obtain a comprehensive
picture of the creep of wood during drying, all these conditions should be covered
by experiments. Second, creep measurements must be long-standing (should last
around as long as the drying itself takes) and it is challenging to create and maintain
controlled climatic conditions that correspond to the drying conditions and maintain
them during the creep measurement. Third, the actual measurement of creep
deformation is a challenging task, since it has to occur in difficult conditions (high
temperature, high humidity, e.g. Hanhijärvi 1998).

A comprehensive picture of the creep in wood in obtained only by performing


experiments both in constant states and changing states. The constant state
experiments measure the viscoelastic creep of wood. In changing state experiments,
however, both viscoelastic creep and mechanosorptive creep occurs. So, if one
wants to distinguish the mechanosorptive creep, the viscoelastic creep deformation
should be in one way or another be subtracted from changing state results. Besides,
shrinkage, too, occurs in changing conditions and it should be subtracted from
changing state results.

The loading of specimens can be either applied as a direct load, which is the more
common method used in creep experiment, but also as restrained deformation. In the
former case, the deformation is measured. In the latter case, the force induced by the
restraining of deformation is measured and these experiments in constant conditions
are called relaxation tests, since creep, as it occurs, relaxes the stresses. A special
way to do experiments for drying wood is to restrain shrinkage deformation and
measure the force that is generated as the test piece dries (Ranta-Maunus 1992,
Svensson 1995, 1996).

a) Measurements in constant conditions

Maybe the easiest constant condition set-up is the measurement of creep of saturated
(wet) wood in water bath. Water bath is easy to arrange and different temperatures
levels can be covered by a simple thermostat to control water heating. Any constant
condition below fibre saturation requires the use of a climatic chamber where both
the temperature and humidity are regulated. This requires in the general case a
system that contains apparatus to heat and cool air as well as moisten and
dehumidify it. In order to really have a constant state in the pieces of wood to be
tested, they should always be allowed to obtain an equilibrium, in particular
equilibrium moisture content, with the conditions that they are placed in before any
164 Fundamentals of wood drying

loading occurs. Examples of constant condition set-ups can be found from the
references to viscoelastic creep results below.

b) Measurements in changing state

Changing state experiments can be basically made in two ways. (1) One can place a
test piece in a certain constant condition and load it before it has attained
equilibrium, (particularly equilibrium moisture content). Then the tested piece
undergoes changing state due to its tendency to obtain the equilibrium. In this way,
deformation under known load during drying can be measured (e.g. Hisada 1986). In
order to determine the creep deformation in the conditions, the amount of shrinkage
has to be measured and subtracted from the result. Further, if the mechanosorptive
creep is to be determined, the viscoelastic creep has to be subtracted also in some
way. (2) Alternatively, one can use a climate chamber to change conditions during
tests. In this way, also non-monotonous states (drying–wetting cycles) can be
covered. Obviously, the need for subtracting shrinkage and viscoelastic creep
deformation applies in this case, too, if mechanosorptive creep is to be
distinguished.

8.4.2 Viscoelastic creep

Results of static viscoelastic creep tests in the conventional drying temperature


range have been published by Svensson (1995, 1996) at two temperatures (60 and
80°C) and two moisture contents for Scots pine (Pinus sylvestris). Also the work of
Joyet (1992) on Maritime pine (Pinus pinaster) contains measurements at many
moisture contents, even if testing temperatures are lower than the normal drying
temperatures. Hanhijärvi (1999) has published results of the creep of Scots pine and
Norway spruce in constant conditions in high temperature drying conditions (95–
125°C), mostly in wet conditions.

Sawabe (1974) described experiments on the creep of dry hinoki (Chamaecyparis


obtusa Sieb.) in the temperature range 20–180°C, although creep in dry conditions is
rather uninteresting in regard to drying. Also some experiments on dynamic
viscoelastic properties touch the drying temperature range: Salmén and Fellers
(1982) and Salmén (1984) report dynamic measurements between 20 and 140°C.

8.4.3 Mechanosorptive creep

The mechanosorptive creep effect is very important for the deformation of wood in
drying conditions. In fact, its magnitude usually exceeds the amount of viscoelastic
creep during drying. Mechanosorptive creep occurs, when wood is stressed and
subject to moisture change below fibre saturation simultaneously, which obviously
is the case in drying. As mechanosorptive creep occurs as function of moisture
changes, it is often convenient to plot creep results in drying (or changing moisture)
conditions against moisture content or shrinkage strain at zero load. Especially,
using zero-load shrinkage strain is straightforward for drying creep test, because it
Creep deformation in drying wood 165

can be measured simultaneously by 'dummy'-specimens without loading (see Fig


8.5). This very useful way of presenting results was first adopted by Svensson
(1996).

0.04 T110K4 110


Dry bulb temp.
100
0.02 Wet bulb temp.
Specimen 4, K5A3, 0.2MPa 90

Temperature [ºC]
Spec. 1, K5A1, 0.05MPa
0 Spec. 3, K5B1, 0.1MPa 80
Strain [–]

Spec. 2, K5A3, 0.1MPa


-0.02 70
Loading
60
-0.04
Dummy 1, K5A2 50
-0.06 Dummy 3, K5B2 40
Dummy 4, K5B4
-0.08 30
0 10 20 30 40 50 60 70
Time [h]

Figure 8.5. An example result of a drying creep test experiment at 110°C showing
the dry and wet bulb temperatures, and the total strain development of 7 specimens,
of which 3 were 'dummy'-specimens with no load to measure pure shrinkage strain.
Loading of creep specimens occurred at ca. 29 h. Even if efforts have been taken to
have specimens with matched properties, the natural scattering of the behaviour
between specimens and corresponding dummies makes the analysis of results
difficult. It is necessary to test several specimen-pairs to improve accuracy
(Hanhijärvi 1997)

Due to its important practical consequences, creep during drying in conventional


drying temperatures (which consists mostly of mechanosorptive creep) has been
investigated by several researches. Hisada (1986) performed on extensive study on
the deformation behaviour of makanba (Betula maximowicziana Reg.) and hinoki
(Chamaecyparis obtusa Sieb.) specimens (tangential) during drying under variable
loads, including zero load. Drying conditions included constant temperatures 20, 30,
50, 70 and 80°C. The experiments included different loading schemes, in which the
166 Fundamentals of wood drying

load was applied in the beginning of drying and removed in the late part of it, or was
applied only at different stages of drying.

Kangas (1990) investigated the creep of drying spruce and pine wood at
temperatures 40, 60 and 80°C under tensile load. Ranta-Maunus (1992) measured
the force which is generated when the shrinkage deformation is prevented of a
circular ring cut from wood along the annual rings. Both investigations have been
summarised by Ranta-Maunus (1993).

Svensson’s (1995, 1996) results contain both drying strain measurement results
(under load or zero-load) and restrained shrinkage force results, giving a
considerably better picture of the deformation behaviour than creep results alone.
Temperatures include 60 and 80°C.

Wu and Milota (1995, 1996) measured the tangential mechanosorptive deformation


of Douglas-fir at 65.5°C under monotonous desorption and monotonous adsorption.

The work of Joyet (1992) is directed towards investigating creep in service


conditions (~ room temperature) but provides useful fundamental information about
the mechanosorptive creep, especially concerning the effect of many repeated
moisture cycles.

Hanhijärvi (2000a) studied the perpendicular-to-grain creep of drying Scots pine and
Norway spruce in the high temperature conditions. The results include also a
comparison to other researchers' results and determination of the dependency on
temperature, Fig. 8.6.
Creep deformation in drying wood 167

0.09
Hisada, 20°C, 1.92MPa
Hisada, 30°C, 0.48MPa
Hisada, 50°C, 0.34MPa
0.08 Hisada, 70°C, 0.21MPa
Hisada, 80°C, 0.15MPa
Svensson, 60°C, 0.8MPa
Svensson, 60°C, 0.8MPa
0.07 Svensson, 80°C, 0.4MPa
Svensson, 80°C, 0.4MPa
Wu&Milota, 65.6°C, 0.83MPa
0.06 Kangas, 60°C
Kangas, 80°C
Compliance [1/MPa]

Hanhijärvi, 95°C

0.05

0.04

0.03

0.02

0.01

0
0.02 0 -0.02 -0.04 -0.06 -0.08
Zero-load shrinkage strain [–]

Figure 8.6. TOTAL compliance during drying creep tests showing results from
several researchers. Hisada’s (1986) results are on hinoki (Chamaecyparis obtusa).
Wu and Milota’s (1995) results are on Douglas fir (Pseudotsuga menziesii).
Svensson’s (1996), Kangas’ (1990) and Hanhijärvi’s (1997) results are on Scots
pine (Pinus sylvestris). Kangas' and Hanhijärvi’s curves are estimations of average
168 Fundamentals of wood drying

8.4.4 Relative magnitude of the creep mechanisms

In low and conventional temperature drying conditions the magnitude of viscoelastic


creep is in the same order of magnitude, but is still usually smaller than the
magnitude of mechanosorptive creep. With higher temperature the magnitude and
relative importance of viscoelastic creep increases compared to mechanosorptive
creep. The viscoelastic creep is particularly important, when the wood is wet or
moist, but practically its importance disappears compared to mechanosorptive creep
for dryer wood.

8.5 Multidimensional effects

The anisotropy ratio between different directions in creep is of the same magnitude
as of the anisotropy of the elastic deformation. The elastic constants have been
rather well studied, the anisotropic elastic constants have e.g. been dealt with in the
textbook by Dinwoodie (1981). Based on elastic properties of wood a tentative
picture of the anisotropic behaviour of creep can be obtained. A good correlation can
be assumed especially for the viscoelastic creep and elastic constants. However, how
the Poisson's constants behave in long term loading is unclear (are they constant or
time-dependent), since measurements are lacking. For mechanosorptive creep the
correlation to elastic deformation holds also, but can be weaker since the mechanism
is basically dependent on hygroexpansion, too. Unfortunately, solid experimental
data is lacking for reliable conclusions.

8.6 Modelling of creep

8.6.1 Viscoelastic creep

Modelling of creep is of importance for different purposes. Lately, the


computational simulation of drying stresses has been the main motivation for
application of creep models. Other deformation processes that occur during drying
(shrinkage, etc.) are relatively easy to model, so that creep forms the most difficult
task in modelling and also constitutes maybe the greatest source of uncertainty of
prediction.

Various models from the general theory of viscoelasticity can be and have been
adapted for wood. The used models include:
– purely mathematical expressions for the creep curve, e.g. the power law,
– mechanical analogy models of springs and dashpots: Kelvin, Maxwell,
Burgers models etc. These models can also be understood as pure mathematical
differential equations for the evolution of creep apart from any mechanical analogy,
– relaxation or retardation spectra or their discrete counterparts, generalised
Maxwell and Kelvin materials.

In principle, all of these can be represented in a hereditary integral form in which the
kernel function depends on the chosen model. In the hereditary integral form all the
Creep deformation in drying wood 169

models are also capable of modelling recovery. However, the disadvantage of the
integral formulation is that it is inconvenient with numerical methods. Therefore the
differential formulations have been more utilised in recent models.

Numerous models have been used for the quantitative description of the viscoelastic
creep of wood. Most of them are of the two first types mentioned above. All
approaches have been able to give satisfactory fits to experimental results – but
mostly only in the studied range of conditions. Namely, a common feature for
different models is that they perform well in a certain experiment length range to
which their parameters have been fitted but give poorer results, when they are used
for an experiment whose length is of a different order of magnitude. Therefore, care
should be taken when applying any model to a wide range of conditions. Usually the
application range can be widened if more material parameters are introduced and a
new fit to more experimental results is made, e.g. by adding more elements to
generalised Kelvin/Maxwell material or adding extra terms to the power law. Both
linear and non-linear modifications of these models have been used, the non-linear
ones having, of course, a wider application range, but with the drawback of the need
to determine values for more parameters

It is not fruitful to try to put the different models in the order of superiority, since
different models are suitable for different purposes. Simple models based on a
mathematical expression for the creep curve in constant conditions (e.g. the power-
law model) are difficult to apply for cases of varying conditions and stress, which is
the case in drying, but they can be used in simple calculations of creep magnitude.
Models based on simple viscoelastic models (Kelvin-Voigt model, Burgers model)
are easier to apply in changing conditions and stress, but the their application is
limited to rather small range of conditions and short times due to inaccuracy if wide
application is needed. Nevertheless they are probably still the most widely used
models. For application in a wide range of conditions and long times like drying
simulation the more complicated viscoelastic models are required (generalised
Maxwell model or generalised Kelvin model), but then the number of parameters to
be experimentally determined increases greatly.

It can be concluded that the quantification of viscoelastic creep can be performed


with different models, but it still remains as a task to find simpler models, with less
parameters and internal variables to fit very wide time and condition ranges, that are
encountered during drying. Of course all model application is limited by the lack of
experimental data for verification.

8.6.2 Mechanosorptive creep

a) General

Modelling of the mechanosorptive creep deserves more attention as it is so specific


to wood and important for drying. Actually, many models of the mechanosorptive
creep are actually counterparts of the above mentioned and commonly used
170 Fundamentals of wood drying

viscoelastic ones. The following is a review of the progress of models and therefore
not a complete list of all introduced models.

The early models of mechanosorptive creep and also most later models in general
have been based on the assumption that it is separable from other strain components.
This is to say that the total strain and its rate are assumed to be divisible into parts:

ε = ε e + ε x + ε ve + ε ms (8.1)

εɺ = εɺe + εɺx + εɺve + εɺms (8.2)

where ε is the strain and εe the elastic, εx the hygroexpansion (swelling and
shrinkage), εve the viscoelastic and εms the mechanosorptive part of it. The strain
terms are treated by associating a separate differential equation (evolution law) to
each one.

b) First generation (Maxwell type) models

The earliest models date to the 70’s (Ranta-Maunus 1973, 1975; Leicester 1971).
The approach by Ranta-Maunus was done in the original works using integral
formulation, but the same idea for mechanosorptive creep behaviour is described by
the following differential form (Ranta-Maunus 1990, the elastic and viscoelastic
parts have been omitted here for simplicity):

εɺ = εɺx + εɺms = (α + aσ )Xɺ (8.3)

where σ is the stress, α is the hygroexpansion coefficient and Xɺ the moisture


content change rate. The mechanosorptive parameter a obtains different values
depending on the how the moisture content X is changing:

a = a+ , if Xɺ > 0
(8.4)
a = a− , if Xɺ < 0

Equations (8.3) and (8.4) can be rearranged in the following manner:

a+ − a−
εɺms = σ Xɺ = mσ Xɺ = mσ sgn Xɺ Xɺ ( ) (8.5)
2

 a+ + a− 
εɺx =  α + σ  Xɺ = (α + βσ )Xɺ (8.6)
 2 
where the difference to (8.1) and (8.2) is that the partitioning of strain between
mechanosorptive and hygroexpansion is made differently. Equation (8.3) can be
characterised as mechanosorptive creep accumulation as a function of moisture
Creep deformation in drying wood 171

content changes in either direction and equation (8.4) as stress dependent


hygroexpansion. Apparently Bazant (1985) was first introduced the absolute value
(or signum) notation as given in Eq. (8.3).

Equation (8.3) can be characterised as a ‘mechanosorptive dashpot’, because it gives


exactly the same expression for the creep rate as is the flow rate of a linear ‘dashpot’
in conventional rheological models just with the time differential dt replaced by |dX|.
In other words, equation (8.3) can be written

ε ms = mσ (8.7)

where the *-notation is used to denote d/|dX| and m can be interpreted as a parameter
which is the inverse of the ‘viscosity’ of moisture change generated creep. Equation
(8.7) together with the elastic response thus corresponds to a ‘mechanosorptive
Maxwell unit’ in analogy to the Maxwell model of viscoelastic creep.

The Maxwell model is probably still the most widely used model for describing the
mechanosorptive creep in drying, even if more advanced models are nowadays
available. Nevertheless, it is not completely unsatisfactory for drying applications,
especially if the drying is such that the moisture change is a monotonously
decreasing or at least does not include several intermediate moistening periods.

c) Dependence of hygroexpansion on strain

An important aspect of modelling deformation in cycling moisture is demonstrated


by the comparison of equation pairs (8.1)–(8.2) and (8.3)–(8.4), viz. the division of
strain between a fluctuating part and an accumulative part. The former can be
characterised as stress dependent hygroexpansion and the latter one as true
mechanosorptive creep. This separation was first pointed out by Hunt and Shelton
(1988). This division is essential for the development of a more advanced category
of models, the Kelvin type models described below.

Equation (8.4) can be interpreted as the dependence of hygroexpansion coefficient


on stress. Hunt and Shelton (1988) proposed similar kind of dependence, but used
strain as the base for it:

εɺx = (α − bε )Xɺ (8.8)

where b is a positive constant. If this dependence is understood as dependence on


total strain (excluding hygroexpansion), then it inevitably includes some dependence
on stress (cf. Equation (8.4)) through the elastic strain.
172 Fundamentals of wood drying

d) Kelvin type models

Hunt (1989) also performed an empirical fit of mechanosorptive creep compliance


with uniformly-sized moisture cycles to functions of one and two exponential terms.
A very close fit could be obtained with a two term exponential equation

( ) ( )
ε ms = J1 1 − e-n/N 1 σ + J 2 1 − e-n/N 2 σ (8.9)

where n is the number of cycles, J1 and J2 are characteristic compliances and N1


and N2 characteristic cycle numbers. This is the mathematical form that expresses
the accumulation of mechanosorptive creep as equivalent to two ‘mechanosorptive
Kelvin units’. In the form given in Eq. (8.9) it is applicable only to uniform cycles,
but Hunt (1989) proposed also the use of the quantity Σ|δu| (δu denoting any
moisture content change) in substitution of n and moisture changes U1 and U2 in
substitution of N1 and N2, which makes the model applicable to non-uniform cycles
and instants within cycles. The differential form of the Kelvin element type model
has been introduced independently by Salin (1992) and Yahiaoui (1991):

εɺms =
(J1σ − ε ms ) Xɺ (8.10)
U1

U1/J1 equals the ‘viscosity’ of the flow element in the Kelvin unit. Salin used the
model in drying stress simulation. Ranta-Maunus et al (1995) used two Kelvin type
units for mechanosorptive creep in series in two-dimensional drying simulations.

Mårtensson and Svensson (1997) introduced a model that is otherwise similar to the
Kelvin type model, but with the difference that the total strain excluding
hygroexpansion, ε–εx, instead of εms is used in the numerator of an equation like
(8.10). Their model was used with success in drying simulation (Svensson and
Mårtensson 2002).

e) Generalized Kelvin and Maxwell type models

Exactly as the application range of the Kelvin and Maxwell model for viscoelastic
creep can be widened by placing several of them in series or parallel, respectively,
the application of their mechanosorptive counterparts can be widened in the same
way. Yahiaoui proposed the use of N Kelvin units in series, i.e. a generalised Kelvin
model for mechanosorptive creep:
N
ε ms = ∑ε
i =1
ms , i (8.11)
Creep deformation in drying wood 173

εɺms =
(J i σ − ε ms,i ) Xɺ (8.12)
U1

Similarly the stress–strain relation of a system of N Maxwell elements in parallel


becomes:

σɺ i
εɺ = + σ i mi Xɺ (8.13)
Ki

N
σ= ∑σ
i =1
i (8.13)

in which σi’s are the stresses of the individual Maxwell units, Ki and ηi are the
spring moduli and viscosities of the elementary elements within the Maxwell units.

Apparently, these generalized models have not been applied in drying stress
simulation, but a modification of the generalized Kelvin model has been introduced
by Hanhijärvi (2000b) and used later in a refined form by Hanhijärvi and
Mackenzie-Helnwein (2003).

f) Models of combined activation

The models of the combined type differ from all the previously introduced models in
the important respect, that the total creep strain is no more (directly) separable into
viscoelastic and mechanosorptive parts as in equations (8.11) and (8.12), but the two
creep mechanisms are dealt with the same equation and have a coupling effect.
These models have been developed for applications in analysis of creep in timber
structures and longitudinal creep, and, accordingly, they have not been applied to
drying simulation.

Bazant (1985) proposed a possibility to model the strain induced by simultaneous


moisture content change and stress in the context of a generalised Maxwell material
model for creep at constant moisture content. In order to take into account the effect
of changing conditions a dependence of the viscosities ηi on the absolute value of
the moisture content rate is added. The stress–strain relation of a system of N
Maxwell elements in parallel can thus be written:

σɺ i
εɺ = +
σi
(1 + a hɺ )
i (8.15)
Ki ηi

N
σ= ∑σ
i =1
i (8.16)
174 Fundamentals of wood drying

in which σi’s are the stresses of the individual Maxwell units, Ki and ηi are the
spring moduli and viscosities of the elementary elements within the Maxwell units;
i=1…N. hɺ is the relative vapour pressure change rate (or, could be also the moisture
content change rate). ai’s are constants. This approach signifies the addition of the
mechanosorptive creep effect within the dashpots, of the parallel Maxwell units.

In the same way, based on the concept of generalised Kelvin material, Yahiaoui
(1991) proposed the use of Kelvin units in which the accumulation of
mechanosorptive creep is taken into account by suitable ‘hygrometric activation’. It
is done by modification of the retardation times τi of the Kelvin elements by making
them dependent on moisture content change rate. The expression for the creep strain
in case of several Kelvin elements in series is
N
ε ve = ∑ε
i =1
ve, i (8.17)

εɺve,i =
(J i σ − ε ve,i ) (1 + β Xɺ ) (8.18)
i
τi
βi’s are material parameters associated with mechanosorptive creep development.

The author is not aware of any applications of the combined type models to drying
simulation, nor any other transverse application. Even if there is some evidence of
coupling between the viscoelastic creep and mechanosorptive creep in the
longitudinal direction (Hanhijärvi and Hunt 1998), no efforts for clarifying such a
phenomenon for the transverse direction have been made. Although such a coupling
may exist for the RT-plane too, there is no indication that it would be large enough
show practical importance for model applications with the present level of accuracy.
9 External heat and mass transfer
Jarl-Gunnar Salin

Swedish Institute for Wood Technology Research


P.O.B. 5609
SE-114 86 Stockholm
Sweden
JarlGunnar.Salin@tratek.se

ABSTRACT. This chapter investigates the interaction between the wood surface and the
surrounding medium (air, steam). Heat and mass transfer coefficients can describe this
interaction. The main part of the chapter concerns the prediction of the heat transfer
coefficient in various situations occurring in a kiln timber stack. The results are presented as
dimensionless correlations. The most important case is the flow between board layers (with
boards in all positions). Entry phenomena and the influence of gaps between boards and
other irregularities are discussed. When the kiln stack is built from random length boards,
then the ends of the stack will have empty board positions, resulting in a quite different flow
pattern. Heat transfer for this case, as well as situations occurring in small scale laboratory
tests, are also investigated.

Mass transfer coefficients may be estimated from the same correlations as the heat transfer
coefficient, after a simple transformation based on the analogy between heat and mass
transfer. Apparent deviations from this analogy have, however, been reported and the reasons
are discussed. It is suggested that the formation of a dry shell in drying above the FSP
explains the deviations observed. Below FSP the importance of the sorption hysteresis is
pointed out as a possible explanation.

Further the surface emission concept and its drawbacks and in addition some other means of
energy transfer to the wood are mentioned. Finally, some pitfalls related to the handling of
external heat and mass transfer calculations are listed.

KEYWORDS: Wood, kiln drying, flow pattern, transfer coefficients, dry shell, hysteresis

Fundamentals of wood drying, Patrick Perré editor, pages 175 to 201.


176 Fundamentals of wood drying

9.1 Introduction

A drying process can be viewed as consisting of two parts. First, an internal part
primarily composed of the moisture migration from the inner parts towards the wood
surface. Second, an external part composed of the interaction between the wood
surface and the drying medium (air, steam). This includes the transfer of heat to the
wood surface and the transfer of evaporated moisture from the surface to the drying
medium. The heat of evaporation consumes most of the heat that is transferred. The
drying resistances that are associated with these internal and the external parts are
both important. In wood science the main interest has been focused on the internal
part, as this is directly related to certain properties of the wood material. The
external part is related only to the properties of the drying medium and its flow, and
has thus often been partly neglected in investigations regarding the wood drying
process.

In this chapter the heat and mass transfer between the drying medium and the wood
surface will be discussed.

9.2 Basic formulation

Heat and mass transfer to and from a solid (or liquid) occurs in a multitude of cases
in many different processes. There is as a consequence a vast literature on this
subject and the reader is referred to standard textbooks, such as (Incropera and
DeWitt 1996, Rohsenow et al. 1998), for descriptions of the theoretical foundations
of these phenomena. The basic equations are

Qɺ / A = hh (T − Ts ) (9.1)

mɺ / A = hm (cs − c ) (9.2)

where
T, Ts are the temperatures of the drying medium, in the bulk and in
contact with the surface, respectively, K
hh is the heat transfer coefficient, W/m2K
Qɺ is the heat flux, W
A is the transfer area, m2
c, cs are the moisture concentrations in the drying medium, in the
bulk and in contact with the wood surface, respectively, kg/m3
hm is the mass transfer coefficient, m/s
mɺ is the moisture flux, kg/s

Due to thermal equilibrium, Ts is also equal to the wood surface temperature.


Eqs. (9.1) and (9.2) are the defining expressions for the heat and mass transfer
coefficients (hh and hm). The moisture concentration (kg/m3) has been used in
Eq. (9.2) as a driving force for the mass flux. However, any other variable
External heat and mass transfer 177

transformable into a concentration could have been used. The general task is now to
be able to predict numerical values for the transfer coefficients as functions of
drying medium properties, flow velocity and flow pattern.

It should be noted that these coefficients are not directly dependent on any wood
properties, only indirectly through the influence of the form of the piece of wood on
the flow pattern, and through the sorption curve that determines the moisture
concentration of the medium in contact with the surface. This means that results
obtained with other materials, or even other drying media can be used (after proper
transformations) for predictions of hh and hm for wood drying, provided that the flow
geometry is equal, or at least similar. As a lot of such results are found in the
literature, especially for the heat transfer coefficient, this source of information
should be utilised as far as possible. This is done below in section 9.4.

Due to the “analogy between heat and mass transfer” a heat transfer coefficient
value can be transformed into a mass transfer coefficient value for the corresponding
situation, and vice versa. It is thus in principle sufficient to be able to predict either
one. For wooden surfaces some complications have however been observed. These
questions are discussed in section 9.5.

The surface emission concept has been used to some extent in wood science to
describe the interaction between the wood surface and the drying medium. This
method, that has some serious drawbacks, is discussed in section 9.6. The chapter
ends with a discussion on the influence of the drying method and with a list of some
common pitfalls regarding external heat and mass transfer.

9.3 How to measure external drying resistances

Direct measurement of heat and mass transfer coefficients in a wood drying process
is extremely difficult. Such direct approaches are however seldom needed. It is
easier to measure pure heat transfer (without simultaneous mass transfer) for a
“board” made of a more suitable material than wood, using an otherwise equal
geometrical configuration. A traditional way is to determine the heat flux from the
electric power fed to the heaters involved and if the temperature difference in
Eq. (9.1) is measured also, then the heat transfer coefficient is obtained. The
difficulty is to measure the surface temperature accurately, but sensors embedded in
the solid or infra-red sensors may be used. Details regarding this and other methods
are found in textbooks on heat and mass transfer.

Another traditional way to determine transfer coefficients is the naphthalene


sublimation method. The solid is covered with a layer of naphthalene and the mass
flux is determined by measuring the change in the thickness of this layer. This
method enables determination of the local transfer coefficient, instead of the average
value over a certain area, as is the case with many other methods. The naphthalene
sublimation method was used by (Kho 1993) for the determination of the local mass
transfer coefficient in a kiln stack configuration.
178 Fundamentals of wood drying

The main problem regarding direct measurements with wood is the determination of
the surface MC (and thus of the corresponding cs-value in Eq. (9.2), from
equilibrium considerations). An interesting method was introduced in (Yeo 2001,
2002a,b, Yeo et al. 2001) where the optical properties of cobalt chloride (CoCl2)
hydrate are utilised. The colour of the hydrate – and a wood surface “painted” with it
– depends on the relative humidity of the surrounding air. A colorimetric
measurement thus makes a determination of the surface MC possible.

Finally a theoretical calculation – as an alternative to direct measurements – has to


be mentioned. The computational fluid dynamics (CFD) modelling of the fluid flow
around the solid can be extended with a heat and mass transfer calculation. Today
the reliability of such calculations is comparable to direct measurement, or
sometimes even better. Some of the results presented in section 9.4 are obtained in
this way. The CFD technique is described in more detail in the “Airflow within
kilns” chapter.

9.4 Heat transfer coefficient

An annotated review of reported data on convective heat transfer in flow


arrangements similar to those occurring in a kiln timber stack is given in the
following. In the traditional way, dimensionless variables are used, which
considerably increases the generality of the formulas presented.

9.4.1 Centre part of normal kiln stack

Figure 9.1 presents the basic situation in a kiln stack where boards are placed edge
to edge in layers and the layers are separated by stickers. The drying air, or steam, is
flowing between these layers, perpendicular to the board direction. This is the
general situation in the centre part of a normal kiln stack.

Figure 9.1. Basic kiln stack configuration


External heat and mass transfer 179

a) Fully developed flow between parallel plates

When heat transfer coefficients are considered in a timber stack, the natural starting
point is to investigate the case of fully developed flow between parallel plates. In a
timber stack where boards are located edge to edge, without a gap between boards,
the board layers create a flow geometry of this kind. The traditional procedure for
turbulent flow in non-circular ducts is to define an equivalent characteristic length,
Dh , called hydraulic diameter and then use correlations for circular tubes. The
hydraulic diameter is defined as four times the cross section area divided by the
wetted perimeter. For a circular tube the hydraulic diameter is thus equal to the
actual diameter. In this way the heat transfer coefficient can be estimated from
correlations for circular tubes that are found in every handbook on the subject. One
well known formula is the Dittus-Boelter equation (Incropera and DeWitt 1996)

Nu = 0.023 Re 0.8 Pr1 / 3 (Re ≥ 10000) (9.3)

where

Nu - hhDh/λ - Nusselt number


Re - vDh/ν - Reynolds number
Pr - Prandtl number
- thermal conductivity of fluid
v - mean fluid velocity
ν - kinematic viscosity of fluid

For parallel plates the hydraulic diameter is twice the distance between the plates,
i.e. twice the sticker thickness in this case. If the Re and Nu numbers are based on
the sticker thickness, Eq. (3) transforms into

Nu s = 0.020 Re0.8
s Pr
1/ 3
(Re s ≥ 5000) (9.4)

where
Nus - Nusselt number, based on sticker thickness s
Res - Reynolds number, based on sticker thickness s

For circular tubes the transition from laminar to turbulent flow occurs for Re about
2300. If this value is adopted for the present case also, it turns out that turbulent flow
normally can be assumed for velocities above 1 m/s. Eq. (9.3) is valid for
Re ≥ 10000 which corresponds to 3-4 m/s in a normal stack. For Re values in the
range 2300…10000 Eq. (9.3) gives slightly too high heat transfer coefficients.

Direct measurements for large aspect rectangular ducts are found in (Haynes et al.
1980), but concern high Pr values (Pr = 10…13). Both theoretical and experimental
values for a fully developed flow between parallel plates are found in (Özisik et al.
1989, Sakakibara et al. 1978, Hatton et al. 1964, Kays et al. 1963) and in references
listed in these. Heat transfer in non-circular ducts is reviewed in (Rohsenow et al.
1998). A correlation based on these values gives for air between parallel plates
180 Fundamentals of wood drying

Nu = 4.3 + 0.0195 Re 0.79 (10 4


≥ Re ≥ 106 ) (9.5)

In the range valid for timber stacks, the Eqs. (9.3) and (9.5) give rather equal values,
which indicates that the hydraulic diameter concept is a good approach.

The correlations above are valid primarily for smooth surfaces and for fully
developed turbulent flow. These limitations are very important and as will be shown
below, the Eqs. (9.3)-(9.5) are actually rather worthless for the prediction of heat
transfer coefficients in real timber stacks.

b) Entry region

The bluff entrance and sudden contraction of the flow into the stack will cause
disturbances and increased turbulence. This results in enhanced heat transfer
coefficients within the entry region. Information on the entry length for flow
between parallel plates is found in (Özisik et al. 1989, Hatton et al. 1963, 1964) and
reviews in (Ashworth 1977, Rohsenow et al. 1998). It seems that these disturbances
penetrate at least 10 hydraulic diameters into the kiln stack and in some cases
20…30 diameters. This corresponds to 0.5…1.5 m for a sticker thickness of 25 mm.
It is thus obvious that the entry region will represent a considerable part of the stack
width and that assuming a fully developed flow throughout, will not give acceptable
results. There are usually several timber stacks in the flow direction in a normal kiln
load. But as the stacks are 100…500 mm apart, the gap between stacks will more or
less create a new entry phenomenon for the next stack.

The flow pattern at the entrance certainly also depends on the thickness of the board
(the board/sticker thickness determines the flow contraction). It is thus natural to
start with a study of the heat transfer for parallel flow over a single flat plate with a
blunt leading edge. This also resembles the rather common case where the drying of
a single piece (or a single layer of pieces) is studied in laboratory conditions.

c) Plate(s) in parallel flow

The flow pattern past a single truncated slab is indicated in Figure 9.2. An eddy is
formed just behind the leading edge. The eddy gives high heat transfer coefficients
close to the point of reattachment, which is located 1…4 board thicknesses from the
edge.

Figure 9.2. The flow pattern past a single truncated slab


External heat and mass transfer 181

The single plate case has been studied by (Sörensen 1969, Ota et al. 1974, 1979,
Kottke et al. 1977a,b, McCormick et al. 1984, Hwang et al. 1996, Danckwerts et al.
1962, Malmquist 1964) among others. A comprehensive review of separated and
reattached flow is found in (Ota 2000). Results for a single plate are valid for a real
timber stack only close to the leading edge. Further into the stack, the board layers
above and below will influence the flow pattern.

Measurements for timber boards in a kiln stack have been reported by (Kho 1993,
Kho et al. 1989, Milota 1994, Langrish et al. 1992, 1993, Langrish 1994, Keey 1994,
Keey et al. 1994, Pang et al. 1995, Pang 1996). Most of the data seems to originate
from the works by Kho and Milota. Figure 9.3 gives an example of such results. It is
seen that there is a maximum in the mass transfer value some distance from the
leading edge, in the same way as for a single plate. There is also an increase at each
gap between adjacent boards, which will be analysed in more detail below.

2,5
3 m/s
Mass transfer coefficient,

2 5 m/s
7 m/s
mol/m2s

1,5

0,5

0
0 100 200 300 400 500
Distance from leading edge, mm

Figure 9.3. The local mass transfer coefficient as a function of the distance from
the leading edge of a stack made up from 25·100 mm2 boards, separated by 25 mm
thick stickers (from Kho 1993)

Sörensen (1969) presented analytical correlations for the heat transfer coefficient for
a single truncated slab, that are useful for the analysis of (laboratory) tests with a
single board layer. The same approach can be used for multiple slabs arranged in the
same way as in a kiln stack. The following analysis is based mainly on the
experimental results found in (Kho 1993) and is an improved version of the
correlation given in (Salin 1996). Sörensen considers separately the regimes before
and after the point where the heat transfer coefficient has its maximum. The position
of this point can be predicted from
182 Fundamentals of wood drying

−0.61
 4.59 1 
Re ∆ = 49.8 +  (9.6)
 
 Re s Re t 
where
Re∆ - Reynolds number, based on ∆
Ret - Reynolds number, based on board thickness t
∆ - distance from the leading edge to the maximum point
t - board thickness

Eq. (9.6) becomes equal to the corresponding Sörensen equation when Res
approaches infinity, i.e. for a very big distance between board layers.

There is not enough data for a detailed analysis of the upstream range z = 0…∆, and
thus a Sörensen correlation for the average local heat transfer coefficient in this
range is recommended

3270 Re t Pr1 / 3
Nu t ,0 − ∆ = (9.7)
202000 + Re1t .3

where Nut - Nusselt number, based on board thickness t.

Far downstream a fully developed turbulent flow occurs and Eq.(9.4) should be
valid. In the intermediate range the following equation seems to fit available data
reasonably well

(
Nu s = Nu 3s0 + Nu 3s1 ) 1/ 3

0.42
 1 − (Re / Re )4.2 
Nu s1 = 0.726  ∆ z  Re 0s .90 Pr1 / 3 (9.8)
 Re − Re 
 z ∆ 
where
Nus0 - Nus according to equation (9.4)
Rez - Reynolds number, based on z
z - distance from the leading edge (z ≥ ∆)

Figure 9.4 presents results calculated according to Eqs. (9.6)-(9.8) for the following
situations; sticker thickness 25 mm, board thickness 25, 50 and 75 mm, air velocity
between board layers 3 and 5 m/s. The upper group of curves corresponds to 5 m/s
and the lower group to 3 m/s.
External heat and mass transfer 183

100
90
80
70
60
Nu s

50
40
30
20
10
0
0 20000 40000 60000 80000 100000
Re z

Figure 9.4. Nus-values for board thickness 25, 50 and 75 mm, sticker thickness
25 mm and air velocity 3 m/ s (lower curves) and 5 m/s (upper curves)

It is seen that the influence of the board thickness can be neglected on the right hand
side of the maximum point. Thus, from a practical point of view, the board thickness
does not influence the local heat transfer coefficient for the horizontal surfaces of
each board, except for the first board on the air entry side. In the examples presented
in Fig. 9.4, the maximum value is located 14…19 mm from the leading edge.

The equations given above are based on rather limited experimental results and are
in this sense not very reliable. Nevertheless they give a possibility to predict the
magnitude of the enhancement of the heat transfer coefficient in the air entry region.
Fig. 9.4 clearly shows that assuming fully developed flow does not give correct
values for the first boards in this region.

d) Influence of gap between boards and other irregularities

However, there are still some complications that influence the result. The
experimental data used for the analytical correlations above have to be considered as
valid for smooth or nearly smooth surfaces only. A sawn (but not planed) surface is
normally rather rough. Thus a higher turbulence level is expected, which results in
higher transfer coefficient values also (based on equal air velocities). Furthermore,
there will in reality be small gaps between adjacent boards in a board layer and these
gaps will widen due to wood shrinkage during the drying process. These gaps will
also increase the turbulence so that a fully developed flow is never attained. Finally,
real boards are not completely aligned in the stack and not perfectly rectangular in
shape (anisotropic shrinkage). Small differences in board thickness and drying
induced warp (cup, twist, bow and crook) result in a non-ideal flow duct between
board layers.
184 Fundamentals of wood drying

The influence of gaps between boards has been studied by (Kho 1993, Kho et al.
1989, Langrish et al. 1992, 1993, 1994, Esfahanian et al.1999, Sun 2001). It should
however be noted that the influence seen in Fig. 9.3 is most probably exaggerated.
The measurements were made using a single naphthalene coated “board” that was
moved from location to location. This means that the naphthalene boundary layer
concentration profile started to develop from the front edge of that single board and
not already from the first board in the board layer as in a real kiln stack. In addition,
only the lower horizontal surface of the flow duct was coated with naphthalene.
There seems thus to be a lack of reliable experimental information regarding the
influence of gaps between boards. According to a theoretical calculation by (Sun
2001) the influence of the gap should in reality be almost neglectable. A result partly
in the same direction was found by (Salin et al. 1998). The theoretical calculations
include some difficulties, which mean that these results are not fully reliable either.

The influence of varying board thickness has been investigated by (Kho 1993, Kho
et al. 1989, Langrish 1994, Esfahanian et al. 1999). It is generally found that non-
alignment of board surfaces will increase the heat transfer coefficient. This and other
irregularities in the air duct geometry occur of course in a stochastic manner in a real
kiln stack. The possibilities to accurately predict local heat transfer coefficients for a
specific board are thus limited. However, all gaps and irregularities will probably
influence the coefficients in such a way that the curves of the right hand side of
Fig. 9.4 will approach a horizontal asymptote more quickly and that this asymptote
is moved upwards. In a few full-scale tests the final MC of individual boards in a
board layer was measured in a project conducted by the author. By comparing the
results with a global simulation model (which takes into account the local climate
variation) it was possible to estimate the average heat transfer coefficient for each
board. A correlation based on such results (Salin 2001) indicated that the first board
had a clearly higher coefficient, the second showed a very small increase and from
the third board onwards a constant level was found. This constant level represents a
higher value than that seen in Fig. 9.4. This is in agreement with the expected
influence of irregularities etc., in stack geometry.

At the moment the best approach seems to be to use the prediction given by Eqs.
(9.6)-(9.8) for the first board. For the second and subsequent boards the values given
by these equations should be increased in order to account for gaps, irregularities,
warp etc. This can be done, following the principle suggested in (Salin 1996b), by
replacing Rez in Eq.(9.8) by

1− F n Re z 
Re w  −n+ (z > w) (9.9)
 1− F Re w 

where
Rew - Reynolds number, based on w
w - board width
n - board number in the flow direction
F - experimentally determined correction factor
External heat and mass transfer 185

For F = 1 no change is obtained. It is possible that a value as low as F = 0.5 should


be used to account for gaps, irregularities, warp and surface roughness. The full-
scale results also indicated that the first board in the second stack of the whole kiln
load did not show as high values for the transfer coefficient as for the first board of
the first stack, although there was a considerable distance between stacks. This
indicates that a coefficient enhancement according to “entry effects” occurs only for
the first stack and mainly a “gap between boards” effect for the next stacks.

Above, only the horizontal surfaces of each board have been studied. A calculation
for the vertical surfaces in the gap between boards has been presented by (Langrish
1999) for a rather low air velocity between board layers (0.5 m/s). For gap widths 1
and 5 mm the coefficient value is 1 and 6% of the corresponding mean value for the
horizontal surfaces. For a 5 mm wide gap 3-8% values were obtained by (Salin et al.
1998). These values were obtained for an ideal case with no net flow through the
gap, which probably gives a too low value. Values for outer vertical surfaces of the
front and trailing boards are investigated below.

9.4.2 End part of kiln stack

It is quite common, especially in Scandinavia, that the kiln stacks are built from
random length boards, with board lengths dominantly much shorter than the stack
length. Normally, every second board in a layer is placed flush at one end of the
stack and the other boards flush at the other end. The centre part (in the board
direction) of such a stack has the configuration seen in Fig. 9.1 and has been dealt
with above. In the end parts of the stack every second board is however missing and
the horizontal gap between boards equals one board width. Seen from the stack end
(in the board direction) there are two possible configurations – aligned and staggered
– as illustrated in Fig. 9.5. It is obvious that the heat transfer process in this end part
differs from the centre part.

Figure 9.5. Aligned and staggered board configurations in the end part of the stack
186 Fundamentals of wood drying

The boards in the front row of the staggered case are probably not influenced very
much by the other boards. It is thus natural to first investigate heat transfer to a
single board in cross flow. This is also a rather common situation in small scale
laboratory tests.

a) Single rectangular cylinder in cross flow

Heat transfer coefficients for flow perpendicular to a rectangular cylinder have been
measured in several cases. Most cases are however for square cross sections,
(Hilpert 1933, Refai Ahmed et al. 1997) which report mean values for the
circumference and (Igarashi 1985, 1986, Goldstein et al. 1990, Yoo et al. 1993,
Tatsutani et al. 1993) that give local values also. For rectangles with width/thickness
ratios up to 1.5 results are found in (Igarashi 1987) and for a ratio of 6 in (Test et al.
1980) and both report local values.

According to these reports, the mean heat transfer coefficient for the vertical front
surface of a single board may be calculated from

Nu t = 0.71 Ret0.5 Pr1 / 3 (9.10)

where board thickness is used as the characteristic length and velocity is the
approaching air velocity. The equation seems to be valid for 1700 < Re < 53000 and
for a square, rectangle and infinite blunt plate (Kottke et al. 1977a).

The mean heat transfer coefficient for the leeward vertical surface may in the same
way be estimated from

Nu t = C Ret2 / 3 Pr1 / 3 (9.11)

where for aspect ratios 1.0, 1.33 and 1.5 the factor C is equal to 0.198, 0.183 and
0.153 respectively. The equation is valid for Re-numbers from 7500 (aspect ratio
1.5) to 53000 (square). As can be noted from Eqs. (9.10) and (9.11), the heat transfer
coefficient will be higher for the rear surface than for the front surface for Re-
numbers above 10000.

Measured values for the horizontal surfaces of a single rectangular cylinder may be
estimated from

Nu t = 0.126 Ret2 / 3 Pr1 / 3 (9.12)

which seems to be valid for aspect ratios 1.33 – 6 and 7500 < Re < 46000.

b) Multiple rectangular cylinders in cross flow

Cho et al. (1994) have investigated a situation rather similar to the first row in the
aligned configuration of Fig. 9.5. The rectangles are however rotated 90º. A
comparison of measured values to those predicted by Eqs. (9.10-9.12) indicates that
External heat and mass transfer 187

values for the array are considerably higher than for a single rectangle on all four
surfaces. This shows that all coefficients are influenced as the air is forced into a
smaller duct and the velocity is increased.

There are a few reports on transfer coefficients for two or several rectangular
cylinders arranged in a collinear array in the air flow direction. This partly resembles
the situation seen in Fig. 9.5, but the influence of adjacent horizontal board layers is
not always included. Cur et al. (1978) consider two plates (with rather low
thickness/width ratios) in series and Cur et al. (1979) extend this to eight plates.
Sparrow et al. (1980) investigated several parallel arrays of 10 plates each.

Theoretical calculations have been presented for different configurations; Sun


(2001) has studied wide gaps between adjacent boards, which actually equal the
aligned board configuration, Salin et al. (1998) present results for a few simple
aligned and staggered configurations.

It is suggested that heat transfer coefficients for the first vertical row of an aligned
configuration (Fig. 9.5), where boards are one sticker thickness apart, are estimated
as follows. Horizontal surfaces from Eq. (9.8), vertical front surface and vertical
leeward surface from Eqs. (9.10-9.11) and from the results given by Cho et al.
(1994). In the case of a staggered configuration, where boards in the first row are
two sticker thicknesses and one board thickness apart, the suggestion is that Eqs.
(9.10-9.12) are used.

When the following vertical rows in the kiln stack end part are considered, it is
natural to compare the situation to a (heat exchanger) tube bundle with circular
tubes, for which measurement data exist. From such data it is known that the second
row will have higher heat transfer coefficient values due to increased turbulence, the
third row still slightly higher values and gradually an asymptote is reach further
downstream. However, when the available data for rectangular cylinders are
analysed, this enhancement is not generally found. For the aligned configuration the
second row seems to have about 6% lower values than the first row and for the
staggered configuration only 10% higher values. Further downstream no substantial
additional change can be seen. These results may explain why no “entry effect” is
seen for the first board of the second (and subsequent) stack, although there
normally is a rather big gap between stacks in the kiln.

Finally it has to be noted that the air velocity is not the same in the centre and end
parts of the stack. The pressure drop across the stack is certainly the same for both
parts, so the air velocity depends on the difference in flow resistance. This resistance
is higher in the end part, despite more void space, but the complicated flow path
results in more turbulence. There is thus a higher air velocity in the centre part than
in the end part. Measured ratio values are in the 1.21 – 1.29 range and calculated
values are 1.34 (aligned) and 1.42 (staggered) (Salin 2001, Salin et al. 1998). A ratio
of 1.3 is probably a good choice in calculations.
188 Fundamentals of wood drying

9.5 Mass transfer coefficient

9.5.1 Analogy between heat and mass transfer

In a normal drying process, the convective heat and mass transfer between the solid
surface and the gaseous phase can be viewed as two different forms of the same
fundamental phenomenon. In both cases heat and mass has to “diffuse” through a
laminar gas side boundary layer. There is thus a connection between the heat and
mass transfer coefficients also. This “analogy between heat and mass transfer” is
discussed in detail in every textbook on the subject (Incropera and DeWitt 1996,
etc.). According to this analogy, the mass transfer coefficient may be estimated from
correlations that are obtained from the corresponding heat transfer correlations by
replacing the Nusselt-number (Nu) by the Sherwood-number (Sh) and the Prandtl-
number (Pr) by the Schmidt-number (Sc). These dimensionless numbers are defined
as Sh = hmDh/D and Sc = /D where D is the diffusion coefficient for water vapour
in the drying medium (air). It is found that the heat and mass transfer coefficient
ratio is given by

hh
= ρ c p L e1− n (9.13)
hm

where the exponent n for the Lewis number can be taken as 1/3. In most cases where
the gaseous phase is humid air, this equation can be approximated by

hh
hm = (9.14)
ρ cp

where cp represents the volumetric specific heat capacity (J/m3ºC) of the humid air
in the boundary layer.

The heat transfer coefficient has been measured and reported in the literature for an
enormous amount of cases, i.e. different geometries, different flow situations,
different fluids, etc. It should be stressed that the heat transfer coefficient is not
dependent on the material of the solid (being dried), it is describing a property of the
gas side boundary layer only. Thus the heat transfer coefficients calculated by the
equations given in the previous section 9.4 can be used for determination of the
corresponding mass transfer coefficients for wood drying processes. This is the
traditional procedure and it has been used in a lot of wood drying modelling work
reported in the literature.

9.5.2 Apparent deviations from the analogy

However, this procedure has not always been successful as a basis for wood drying
modelling. In several cases, indirectly measured mass transfer coefficients have
External heat and mass transfer 189

given values clearly lower than those predicted by Eqs. (9.13-9.14). Such cases have
been summarised in (Salin 1996a) and a few recent cases are found in (Hukka et al.
1999, Hukka 1999, Tremblay et al. 2000b). It has further been observed that models
tuned for a certain board thickness have failed to accurately predict the drying
behaviour of another board thickness, which indicates an error in the
internal/external drying resistance ratio. Finally, it is often qualitatively found that
the “constant drying rate” period is surprisingly short or even missing, although the
initial MC is very high.

These findings have often been labelled as deviations from the analogy between heat
and mass transfer. The analogy is, however, a fundamental phenomenon, that has
repeatedly been experimentally established. Thus we have an apparent deviation
from the analogy and the primary reason has to be found somewhere else. Some of
the possible reasons are discussed in the following.

9.5.3 External mass transfer above FSP

If the wood surface is above FSP then the expected evaporation rate equals the
evaporation from a free water surface and is thus relatively easily determined
theoretically. When a lower drying rate is observed experimentally, it has been
suggested that one reason could be that the wood surface is not completely wet, but
contains “dry spots” that decrease the effective area for mass transfer, but not the
area for heat transfer. It has however been shown that if the dry spots are small
compared to the boundary layer thickness, then the influence on the mass transfer
coefficient should be marginal. Further, even if there is no free water at these dry
spots, there is still bound water present and the corresponding equilibrium vapour
pressure is not very much lower than for free water - at least not zero. Thus dry spots
do not explain the often very substantial apparent reduction in the mass transfer
coefficient.

A more likely explanation above FSP is the “dry shell” concept. According to this
idea a thin dry shell with no free water is rapidly formed at the wood surface. The
wood surface is thus actually below FSP, although the MC of the piece as a whole is
far above FSP and the MC profile almost throughout the sample thickness gives an
impression (by extrapolation) of a wet surface. The situation is illustrated by the
following Figure 9.6 that presents temperature and humidity profiles across the air-
wood interface.
190 Fundamentals of wood drying

Moisture conc.
cs ce

AIR WOOD

Ts

Temperature Te
Boundary layer

Dry shell

Figure 9.6. Temperature and humidity profiles across the air-wood interface

On the left hand side of Figure 9.6 the temperature T and the moisture concentration
c (kg H2O/m3) in the bulk air flow are shown. In the boundary layer there is a
temperature gradient T – Ts that is the driving force for the heat flux to the wood
surface. The dominating part of this flux is consumed by the evaporation of moisture
and only a small part is associated with a change in wood temperature.
Correspondingly the moisture concentration gradient cs – c in the boundary layer is
the driving force for the transfer of moisture evaporated. The concentration and
temperature profiles within the boundary layer are, for the sake of simplicity,
presented as straight lines. The subscript s stands for “surface” and cs is thus the air
moisture concentration assumed to be in equilibrium with the wood surface MC at
the temperature Ts. Most of the moisture evaporation is assumed to take place at an
evaporation front, which together with the wood surface defines a “dry shell” where
there is no free water. Instead of plotting the MC profile within the wood, the
corresponding equilibrium vapour concentration is used. In this way a continuous
moisture concentration curve across the wood surface is obtained in Figure 9.6.
Within the dry shell there is a moisture gradient that is the driving force for the
migration of vapour and bound water to the surface from the evaporation front.
There is of course also a temperature gradient Ts – Te across the dry shell, which is
the driving force for the conduction of heat to the evaporation front. At least for
relatively slow drying processes below the boiling point, the interior wood
temperature is changing very slowly and the temperature on the right hand side of
Figure 9.6 can thus be assumed constant and equal to the temperature Te at the
evaporation front. Consequently, the vapour concentration ce within this range is
also constant. This completes the description of the general features of a “dry shell”.
External heat and mass transfer 191

According to the dry shell concept, the surface MC is clearly below FSP and the
corresponding air moisture concentration cs is lower than for free water, which
explains the reduction in the expected mass transfer rate. There is free water close to
the surface, but not at the surface. Based on measurements of temperatures and
drying rates reported in (Wiberg 2001, Wiberg et al. 2000) the dry shell thickness
development was calculated for a specific case in (Salin 2002a). Using the
conceptual model given above, the result was that the shell thickness increased
about linearly from zero to 1 mm during the first 24 hours of a 42 hour Scots pine
drying process from an initial MC of 100% to a 27% final MC.

If the dry shell thickness is less than 1 mm, it is of course understandable that the
shell has not been widely recognised long ago. There are however a few direct
experimental observations. CT-scanning results reported in (Wiberg 2001) showed
high MC levels close to the surface and a steep drop in the outermost part close to
the surface. Due to an insufficient spatial resolution details within the shell could not
be seen. A dry shell was rather clearly seen in experiments reported by (Tremblay et
al. 2000b) where five 0.38 mm thick slices were removed from the surface and their
MC determined. A steep MC gradient was seen within a thickness of about 2 mm.
Indications of the same kind have been reported by (McCurdy et al. 2002). Recently
an MRI (Magnetic Resonance Imaging) technique has been used for the
determination of the MC profile within a 0.3 mm thick surface layer (Rosenkilde et
al. 2002). Finally, it is well known (“kiln brown stain”) that certain chemical
compounds, initially dissolved in the free water phase, are enriched 0.5 – 2 mm
below the surface of dried boards (Kreber et al. 1999, Terziev 1996), which is an
indication that the evaporation occurs below the surface.

Modelling of the air-wood interaction during drying at mean MC levels above FSP
thus requires that the influence of the dry shell is included. The moisture migration
resistance within the shell may be described by a diffusion coefficient. However, the
shell thickness is also needed and that imposes a problem, as a detailed model for
this thickness and its changes seems not to exist. The key question is of course to
find the relevant phenomena behind the receding evaporation front behaviour. In the
following a possible process description is indicated for slow drying of wet wood
(softwood sapwood).

Immediately after sawing, there will at the surface be water-filled fibres with open
ends. This free water evaporates easily – even before the board has been loaded into
a kiln – and the meniscus in each fibre is moving inwards as a direct result of this
evaporation. This continues until a narrow throat is reached – typically a bordered
pit opening – where the meniscus stops. Further evaporation will now decrease the
radius of curvature for the meniscus surface, i.e. a suction is developing in the free
water phase. All parts of the wood void structure that are connected through the
same continuous free water phase can be said to form a free water “island”. Cell
walls and a great number of menisci bound this island. Based on measurements by
(Wiberg 2001) it may be deduced that there is only a small resistance to free water
192 Fundamentals of wood drying

flow within such an island. Thus the suction caused by evaporation from a meniscus
will act throughout the island and as a result the water surface in the largest capillary
is the one that will move. This retraction continues until the next narrow throat is
reached and during the process new throats are probably uncovered. Again, the
water surface in the largest capillary retracts and the process is repeated.

The narrow throat diameter values vary according to a stochastic distribution. As the
free water retracts from the largest throats, there is a trend that these disappear from
the distribution. However, as the water retracts new throats are uncovered and these
have diameters according to the initial distribution, i.e. include also large throats.
There is thus an increased probability that if the free water retracts at a certain spot,
then the retraction will continue at the same spot. The result is “fingers” or “trees”
that penetrate into the island. When this process continues, the penetration will
extend into the whole volume of the island and the island gradually disintegrates
from the inside. This explains why no, or very low, moisture gradients are found in
sapwood of softwood above the FSP.

It might be possible that the suction (negative pressure) created in the free water
phase is strong enough to overcome the cohesion forces between water molecules so
that a bubble is created. The process continues after that according to the same
principles as already described. Anyway, although evaporation takes place from the
menisci closest to the wood surface, the free water is disappearing at quite different
points of the island. The centre of gravity of the free water island will not move very
much in this process. However, as dry “fingers” penetrate into the island, it may be
divided into smaller islands or small islands are cut off from a big one. It is clear that
the evaporation front and the outline of the island will not be simple surfaces, but
complicated structures of fractal character. Thus the thickness of the dry shell should
in this context be viewed as the effective thickness. The gas phase relative humidity
in the “fingers” will be very close to 100% (Prat 1995) and evaporation from the
free water surfaces will thus occur almost entirely from the outer part of the island
and close to the wood surface.
External heat and mass transfer 193

2
Dry shell thickness, mm

1,5

0,5

0
40 50 60 70 80 90 100
Average MC, %

Figure 9.7. Dry shell thickness as a function of MC, calculated from measured
temperatures and drying rates

It seems plausible that there has to be a relationship between the dry shell thickness
and the amount of moisture evaporated, i.e. the change in average MC. The form of
this relationship depends on the properties of the capillary network, which are wood
species related and thus this relationship has to be determined experimentally. The
analysis mentioned above (Salin 2002a) gave the following result for the dry shell
thickness as a function of MC.

A rather linear relationship is seen down to 40-50% MC in Figure 9.7. This MC


level represents a rather low amount of free water and further drying cannot be
called “drying above FSP” any more. Drying of capillary networks has recently been
the subject of theoretical research (Prat 2000, Le Bray et al. 1999) and a better
insight into this kind of phenomena may evolve in the near future.

As long as the dry shell thickness development and its dependence on different
factors are not thoroughly understood, it has to be included in models and
calculations in the form of an approximate expression. Thus other approximate
approaches may in practice be equally good. A thin dry shell will only marginally
influence the average MC (as given by the main part of the MC profile in the board)
or other features of practical importance. Thus an obvious possibility is to include
the drying resistance of the dry shell into the external mass transfer coefficient. It is
then easily found that the mass transfer coefficient should be replaced by the
expression:
194 Fundamentals of wood drying

hm
hm/ = (9.15)
hm
1+
D/s
where
s - effective thickness of the dry shell
D - average diffusion coefficient within the dry shell
hm - mass transfer coefficient for the corresponding case with a wet surface

If s is introduced as a linear function of MC change (c.f. Fig. 9.7) and D is seen as


an adjustable parameter (perhaps including an Arrhenius type temperature
correction) then a rather good over all modelling result should be achieved. This
approach may be used for the initial drying period, as long as the “extrapolated” MC
indicates a surface MC above FSP. It has to be stressed that hm/ is an apparent mass
transfer coefficient.

9.5.4 External mass transfer below FSP

As mentioned above, (apparent) deviations from the analogy between heat and mass
transfer have been reported and especially for situations below FSP. There is thus a
reason to analyse how these low values for the mass transfer coefficient have been
obtained. Generally, all values have been calculated from the transfer Equation (9.2)
written in the form:

mɺ / A
hm = (9.16)
cs − c

The moisture flux has in experiments normally been measured by weighing the
sample or by other rather direct methods. The moisture concentration in the
surrounding drying air, c, has been determined by standard methods. The crucial
point is thus the determination of the moisture concentration cs of the air in contact
with the wood surface. A direct measurement is extremely difficult. The normal
procedure is to assume equilibrium between the wood surface and this air. Then the
surface moisture concentration has to be determined instead, which is also a very
difficult task. Normally indirect methods are used. However, if the mass transfer
coefficient is needed for a numerical simulation model, then this model calculates
the surface moisture concentration and temperature from mass and energy balances.
This means that the mass transfer coefficient in this case can be seen as an
adjustable parameter used for tuning of the model. Such a model may predict other
situations with a good accuracy, but the mass transfer coefficient determined in this
way does not necessarily equal the correct value for the external boundary layer.
Thus an apparent deviation from the analogy may occur. If the calculation of the
surface moisture concentration and temperature includes approximations, then the
mass transfer coefficient may be affected, but this is not the kind of problem we are
concerned with here.
External heat and mass transfer 195

The reason for an apparent deviation is that there are several weak points in the
chain of calculations above. First, the equilibrium values are often taken from
sorption curves (or corresponding numerical expressions) found in most textbooks
on the subject. Normally these are initially based on measurements made long ago
with Sitka spruce. These curves may very well differ from the correct curves for the
actual wood species. This, however, is normally not a big problem. More important
is the fact that the sorption hystereris is often neglected, i.e. the average sorption
curve is used instead of the desorption curve in a drying process. This introduces (in
drying) a too high cs value into Eq. (9.16) and the calculated hm value becomes too
low. A drying simulation model tuned in this way may give rather correct values for
desorption, but will fail for absorption processes. Finally, it is known that sorption
processes can be very slow (Wadsö 1993) and it is possible that a complete
equilibrium between the wood surface and the air in contact with it is not achieved
in a dynamic process. This additional resistance is then seen as an apparent decrease
in the mass transfer coefficient value. An important question is now how much of
the apparent decrease in the mass transfer coefficient value that can be attributed to
these possible reasons.

Preliminary calculations (Salin 2002b) regarding the influence of sorption hysteresis


have indicated that at least a considerable part of the observed apparent deviation
from the analogy could be explained in this way. The hysteresis effect was
accounted for by calculating the desorption EMC from the average EMC by dividing
by 0.9 and the absorption EMC correspondingly by multiplying by 0.9 regardless of
air temperature and relative humidity. This equals a A/D-ratio of 0.81 which is in
accordance with reported experimental values (Hartley et al. 2002). Measured
drying behaviour was compared to simulated results for full-scale boards of Scots
pine (Pinus silvestris) and Norway spruce (Picea abies). For Norway spruce the
deviation from the analogy below FSP was almost completely explained in this way
(within the measurement and calculation accuracy). For pine an unexplained part
remained. However, by slightly changing the factor 0.9 to 0.87 a good fit was
obtained. These results show clearly the importance of the sorption hysteresis in
wood drying modelling. Whether any additional explanations (such as a dynamic
non-equilibrium between bound and gaseous phases) are required, remains an open
question.

In summary, it seems that the observed deviations from the analogy between heat
and mass transfer really are apparent deviations that disappear if the drying process
and especially the air/wood interaction phenomena are modelled carefully. The
correct procedure would then be to first determine the heat transfer coefficient from
correlations like those given in section 9.4. Then the corresponding mass transfer
coefficient is determined from the analogy (Eqs. (9.13) or (9.14)). These transfer
coefficients are then used in combination with a model for the dry shell development
and with a model for the sorption hysteresis phenomenon. However, as long as all
details are not available (especially regarding the dry shell thickness development)
an alternative method is justified. In that case the dry shell and/or the sorption
hysteresis is neglected and a correction factor for the mass transfer coefficient is
196 Fundamentals of wood drying

introduced instead. This correction factor has to be established from experiments


with conditions that are reasonably close to the actual situation.

9.6 The surface emission concept

The surface emission approach to the external mass transfer is based on the
following single equation.

mɺ / A = S ( X s − X e ) (9.17)

where
Xs - surface MC (kg/kg)
Xe - equilibrium moisture content (kg/kg)
S - surface emission coefficient (kg/m2s)

The “driving force”, Xs - Xe, reflects that the wood MC ultimately approaches
equilibrium with the surrounding climate, but there is no other direct physical
justification for this expression. Eq. (9.17) seems to have been introduced by
(Newman 1931b) in order to have a simple boundary condition that enabled
analytical solutions to Fick’s equation. This is still the main benefit.

It should be noted that the temperature is not explicitly involved in Eq. (9.17). If, as
an example, a cold and wet board is brought in contact with hot, humid air, then the
wood surface will be wetted (at least if the wood temperature is below the dew
point), but Eq. (9.17) predicts drying. This simple example shows that Eq. (9.17)
cannot be used in dynamic situations, where the climate is changing.

A survey of S-values reported in the literature is given in (Söderström and Salin


1993). An enormous variation was found – the ratio of the highest and the lowest
value was 500000. One reason may be that S by many researchers has been
considered as a wood material property only. In reality S is a mixture of wood
properties and external conditions (flow velocity and pattern, etc.) There is thus no
reliable method available for the prediction of numerical S-values – other than a
transformation of the corresponding hm-value for the actual situation (using the
sorption curve for a recalculation of the driving force).

In summary, the surface emission concept can be used for very preliminary
simplified analyses, but it is strongly recommended that it is avoided in serious
modelling work.

9.7 The influence of the drying method

The low temperature heat and vent drying process is still the dominating method at
sawmills, although other drying methods are used industrially also, at least for some
specific products. In the heat and vent drying method, energy is transferred from the
drying medium to the wood surface by a process that can be characterised as forced
convection heat transfer. This is at least partly the case for many other drying
External heat and mass transfer 197

methods also. This transfer process is described by the correlations given in section
9.4. Some aspects related to high temperature and vacuum drying should however be
considered. In other drying methods, quite different energy transfer processes may
be dominating and these are also briefly listed in the following.

High temperature drying is often defined as drying above 100ºC. A physically more
correct definition is drying above the boiling point of water. In this sense vacuum
drying (in superheated steam) is a high temperature drying process although the
temperature may be below 100ºC. A characteristic of high temperature drying is the
development of internal pressure gradients and thus pressure driven moisture
migration. When such a pressure driven gaseous flow reaches the wood surface it
will continue directly into the boundary layer, as a result of the momentum. This
flow will certainly influence the boundary layer transfer processes, but for moderate
drying rates the two mass fluxes may be directly added. This feature of high
temperature has been especially pointed out by Hukka (1996) and others, but has in
many cases been neglected.

For elevated temperatures or low pressures (vacuum drying) the air and water
vapour properties are affected and thus also the Nu-, Re-, Pr-, Sh- and Sc-numbers
involved in the correlations for the transfer coefficients. The kinematic viscosity
increases as the temperature increases. This dependence can be expressed as ~Tn
where T is the absolute temperature and n is about 1.74 for pure air and about 1.95
for pure steam. According to theory the kinematic viscosity is inversely proportional
to the absolute pressure, p. The thermal conductivity increases as the temperature
increases, and using the same formula, n is about 0.87 for pure air and about 1.31 for
pure steam. The thermal conductivity is almost independent of the pressure. The
influence of temperature and pressure on the Prandtl-number is rather small. These
dependences are approximately valid within the temperature and pressure intervals
normally used in wood drying.

These dependences can now be inserted into the correlations of section 9.4. As an
example, if Nu~Re0.7Pr1/3 represents a typical correlation, then the influence of
temperature and pressure on the heat transfer coefficient is described as follows

p 0.7
hh ~ (pure air)
T 0.35

p 0.7
hh ~ (pure steam) (9.18)
T 0.06
As the relative variation in the absolute temperature is rather limited, the influence
of the temperature on the heat transfer coefficient is not very important. Reduced
pressure (vacuum drying) will, on the contrary, clearly influence the transfer
coefficient. This explains the need for high velocities in vacuum drying.
198 Fundamentals of wood drying

In superheated steam drying the mass transfer coefficient has no relevance. For
drying in air, however, the influence of temperature and pressure on the mass
transfer coefficient may be investigated in the same way as above. The diffusion
coefficient for water vapour in air is approximately proportional to T1.77/p and the
Schmidt-number is almost a constant. For Sh~Re0.7Sc1/3 which corresponds to the
heat transfer case above, we get

T 0.55
hm ~ (pure air) (9.19)
p 0.3
Eqs. (9.18-9.19) should be regarded as approximate. It should be noticed that many
of the properties of humid air are not linear functions of the properties of pure air
and pure steam.

So far the main interest has been in forced convection transfer processes. In
timberyard drying and especially considering the wood-air interaction for indoor
conditions, free convection may play an important role. Cold air is heavier than
warm air, and humid air is lighter than dry. It is thus not self-evident whether the
cold and humid air in the boundary layer will flow upwards or downwards in a
drying process. For lower temperature levels the vapour content (kg/m3) in the air is
very low and the temperature effect will thus in such cases be dominating, i.e., air
flows downwards for moisture desorption. Correlations for the prediction of heat
transfer coefficients for free convection cases are found in all handbooks on the
subject (Incropera and DeWitt 1996, etc.) When these correlations are used, both
temperature and humidity effects on the density should be taken into account, not
temperature only as in the normal situation. It seems that that heat transfer
coefficient values for indoor/outdoor pure free convection conditions are found
within the 0.5…3 W/m2K range.

There are several drying methods where energy is transferred to the wood by other
means than forced (or free) convection. Such methods are using electromagnetic
fields (radio frequency, microwave, infra-red, etc.) or direct solid/solid contact
(press drying). These methods are, however, outside the scope of this presentation.
The reader is referred to surveys found in (Keey et al. 2000, Chen et al. 1997, Turner
et al. 1998, Avramidis 1999, Sandland et al. 2002, Zombori et al. 2003) and
literature cited in these. It should be noted that forced convection heat and mass
transfer may represent an important part in these more sophisticated drying methods
also.

9.8 Some pitfalls regarding external heat and mass transfer

In almost all experiments related to wood drying, heat and mass transfer between the
wood surface and the surrounding drying medium is involved. In surprisingly many
cases, especially in those where the experiment is focused on internal behaviour, the
external part has not been given the attention it deserves and the experimental result
may have been biased. In the following a list is given of some pitfalls and other
External heat and mass transfer 199

important considerations that may influence the result, if the experiment is not
carefully designed.

a) It is a very common situation that some surfaces of a wood sample are sealed by
a non-permeable coating in order to obtain an unidirectional moisture flow
within the sample. The mass transfer through these surfaces is thus zero, but as
the coating normally is relatively thin, the heat transfer is only marginally
influenced. This means that the surface area for heat transfer is larger than the
corresponding area for mass transfer. This sample will thus attain a higher
temperature than, for instance, a full size normal board in the same climate. As
many properties are strongly dependent on the temperature, all such sealed
surfaces should be heat insulated also, in order to give a correct result. This
pitfall was pointed out by Sherwood – the grand old man of drying theory –
already in 1929 but is still frequently ignored.

b) The correlations for the heat transfer coefficient presented and discussed in
section 9.4, clearly show that the coefficient is strongly influenced by the flow
geometry and the stack configuration. Laboratory drying tests are – for practical
reasons – normally performed using a few or even a single “board” only. It is
obvious that the average heat (and mass) transfer coefficient for such a board,
may differ considerably from the coefficient for a specific board in a real kiln
stack, even though the air velocity is equal in both cases. A “similarity”
requirement should thus be based on equal transfer coefficients, not on equal air
velocities.

The same situation is found in scale-up procedures from laboratory size samples
and equipment, to full size kiln configurations. The characteristic length, d, is
included in both the Nu- and the Re-numbers of the correlations given above.
For instance, a proportionality Nu ~ Re0.7 implies that hh ~ d-0.3 when all other
variables are constant. If the characteristic length (board thickness or distance
between board layers etc) is doubled, then the external transfer coefficient will
decrease by 19%. Considering that the internal drying resistance, through the
Fourier number, generally is proportional to the square of the sample thickness,
scale-up situations are difficult to handle if “similar” conditions are required in
both laboratory and full scale.

Accordingly, it is strongly suggested that the external conditions (air velocity,


flow geometry, stack configuration etc.) should be given in detail when
experimental results are published. Very general statements like “the air velocity
was about 3 m/s” without any further details restrict the use of otherwise
valuable results by other researchers.

c) When the final MC is measured with a handheld meter in a full-scale kiln, the
measurement points are, for practical reasons, often located in boards at the
front (or back) edge of the kiln charge. The points are further often selected
from the centre part (in the board direction) of the stack. As seen in Figure 9.4,
200 Fundamentals of wood drying

the first board on the air entry side will have higher external transfer
coefficients for the horizontal surfaces than boards further downstream in the air
flow direction. In addition, the vertical front surface of such a board is an
“active” drying surface, contrary to vertical surfaces further in.

Even with rather frequent fan reversals, the average climate will not be constant
throughout the kiln charge. A higher drying rate is found for boards close to the
edge of the charge, compared to boards in the centre of the charge (Salin 2001).
When these effects are combined, it is obvious that MC values especially for
boards in the first position at the front or back edge of the charge, do not reflect
the average final MC of the whole charge.

If the final MC is measured gravimetrically instead, the samples are preferably


taken close to the end of the board in order not to completely destroy too many
boards. Standards normally require that samples are taken at least 30 or 50 cm
from the end. If the kiln stack consists of random length boards, these samples
are thus taken from the part which in section 9.4.2 was referred to as “end part
of the kiln stack”, where a rather different flow pattern (and drying behaviour)
compared to the “centre part” is found. As the total timber volume is roughly
equally divided between these two parts, such a measurement does not
necessarily reflect the average MC of the whole charge. It is also remarkable
that measurements with a handheld meter and MC values determined
gravimetrically, may, for practical reasons mentioned, give different results.

In summary, in scientific research work the MC sampling and measurement


procedures have to be carefully considered in order to achieve a correct and
differentiated picture of the global result in a kiln.

d) When small wood samples for laboratory tests are prepared, sample surfaces are
normally aligned strictly according to the principal directions of the wood
material (longitudinal, radial, tangential). Further, surfaces are made very
smooth.

The surface of a real board in a kiln is most often a sawn, but not planed, surface
and is thus rather rough, in some cases very rough. This will certainly influence
the boundary layer and thus the external heat and mass transfer coefficients. The
laboratory test is in this sense not completely relevant regarding the industrial
situation. There seems to be rather few investigations of this issue reported in
the literature, i.e. the difference between sawn and planed surfaces.

It is well known that a real board surface represents a mixture of the radial and
tangential directions. However, due to the spiral grain feature found for many
commercially important wood species, the surfaces are not completely oriented
in the longitudinal direction either. This fact, that a real board is not sawn
completely in the principal directions of the wood structure, is often neglected.
External heat and mass transfer 201

It is of utmost importance regarding internal moisture migration, as the diffusion


coefficient is much higher in the fibre direction.

The external heat and mass transfer coefficients describe properties of the air-
side boundary layer and are thus by definition independent of the orientation of
the wood surface in relation to the fibre matrix. Secondary effects, like the
influence of differences in surface smoothness in different directions, should of
course not be included in this context. In addition, wood species or other
properties that are related to the wood material only should, for the same reason,
not influence these coefficients. Despite this, external transfer coefficients have
in some cases been reported in the literature as functions of the principal
directions, wood species, etc. Such results have to indicate errors in the
experimental set-up or other problems.

However, if the “dry shell” resistance is combined with the real mass transfer
coefficient to form an apparent coefficient, as discussed in section 9.5.3, then
the situation changes. The dry shell thickness development is probably
dependent on the surface orientation in relation to the capillary network, on the
wood species and even on the sample thickness. Even though the mass transfer
coefficient is a strictly external property, the apparent mass transfer coefficient
may thus very well be dependent on several internal wood properties.
10 Coupled heat and mass transfer
Patrick Perré1 and Ian Turner2
1
LERMAB
Laboratoire d´Etude et de Recherche sur le Matériau Bois
UMR1093 INRA, ENGREF, UHP,
ENGREF 14, rue Girardet 54 042 Nancy Cedex, France
perre@nancy-engref.inra.fr
2
School of Mathematical Sciences,
Queensland University of Technology,
GPO Box 2434, Brisbane, Q4001, Australia
i.turner@qut.edu.au

ABSTRACT. This chapter is devoted to studying the coupled heat and mass transfer
phenomena involved in wood drying. The important physical mechanisms and related
coupling are explained and the main physical formulations proposed in the literature to
model these phenomena are presented and compared. Specific case studies have been chosen
to highlight the possibilities and limitations offered by these different formulations.

KEYWORDS: Wood, Heat transfer, Mass transfer, Coupled phenomena, Physical


formulation, Numerical simulation.

Fundamentals of wood drying, Patrick Perré editor, pages 203 to 241.


204 Fundamentals of wood drying

10.1 Introduction

Drying is one of the most energy intensive industrial processes finding application in
a wide variety of sectors in science and technology. It is a topic that embraces at its
core, simultaneous heat and mass transfer in porous media and its subject matter
spans a number of disciplines of science and engineering. Over the last three
decades the food, timber and construction industries have injected substantial
funding into fundamental drying research that has ranged from the kiln scale to the
complicated transport phenomena evolving at the macroscopic and microscopic
scales within the product. The knowledge gained from this fundamental work has
led to a better understanding of the drying process, which has enabled highly
optimised drying systems to be investigated and developed by drying engineers and
practitioners worldwide. Mathematical modelling has been and still remains an
integral component of this research work. The advancements made in experimental
work that has paralleled and motivated this modelling has facilitated the
development of more accurate and realistic mathematical formulations. In particular,
the important physical property correlations and parameters necessary within the
framework of the drying conservation equations have been determined using a blend
of sophisticated strategies that include image analysis and homogenisation
techniques. Such advancements have ensured that the proposed models well
represent reality and this fact is becoming increasingly evident in the literature,
where good agreement between theory and experiment is now often reported.

During drying several transfer phenomena are involved, such as heat transfer, water
vapour diffusion, bound water diffusion, capillary migration and bulk liquid and gas
flow. Because all of these phenomena are nonlinear and coupled in nature,
numerical simulation becomes the perfect forum through which to explain and
exemplify these coupled phenomena. Although numerical simulation is an excellent
tool that can be used to resolve these issues, the complexity of the numerical model
must remain consistent with a global project of wood drying improvement or kiln
control. In particular, the final choice must account for:
- the aim of the numerical model - what is the refinement expected for the
process description?
- the knowledge of the product physical properties - are all required values
available? Is it possible to perform or propose experimental strategies to
measure the unknown values?
- the possibilities of software development - what is the numerical expertise
of the team in charge of the project?
Furthermore, numerous other possibilities and options manifest with the
development of a computer code for wood drying simulation, including:
- the number of dependent variables to be used in the physical formulation,
- the number of spatial dimensions to be treated in the numerical solution,
- the most consistent way to deal with drying stresses and deformations.

This chapter is devoted to the coupled heat and mass transfer phenomena that occur
during wood drying. Simple diagrams are used to elucidate the important physical
Coupled heat and mass transfer 205

mechanisms and related couplings involved in the process. An entire set of


experimental data is proposed to gain more understanding of the subtle coupling that
exists between heat, mass and momentum transfer during drying. The effect of
drying conditions as well as the effect of the wood properties will be particularly
exemplified.

In order to capture this complexity, the most common formulations encountered in


the literature are presented and compared using the results of several numerical
simulations. The conclusions of the work are summarised in a table that highlights
the possibilities offered by a computational model that is a function of the number of
variables used in the physical formulation and the number of spatial dimensions
used for the simulation. Three recent wood drying case studies for high temperature
convective drying, combined microwave and convective drying and vacuum drying
with radiative heating have been chosen to highlight the important mechanisms of
drying.

10.2 Description of heat and mass transfer during drying

10.2.1 Low temperature convective drying

Low temperature convective drying is one of the most widespread industrial


conventional kiln drying processes. In this case, the role of internal gaseous pressure
is almost negligible and transfer occurs mainly in the direction of the board
thickness. Concerning drying mechanisms, two periods must be distinguished:

a) The constant drying rate period

This stage of drying is very common for some porous media, especially for example
with aerated concrete, however it is more difficult to observe for wood. In fact, the
constant rate drying period exists almost always for fresh boards consisting of
sapwood that are dried in moderate conditions (Perré et al. 1993, Perré and Martin
1994). During this period, the exchange surface of the board is still above the fibre
saturation point (FSP) and as a result, the vapour pressure at the surface equals the
saturated vapour pressure, which is a function of only the surface temperature.

Crossed heat and vapour transfer occurs in the boundary layer during this period,
and the heat flux supplied by the airflow is used solely for transforming the liquid
water into vapour. During this stage, the drying rate is constant and depends only on
the external conditions (temperature, relative humidity, velocity and the flow
configuration). The temperature at the surface of the product is equal to the wet bulb
temperature. Moreover, because no energy transfer occurs within the medium during
this period, the whole temperature of the board remains at the wet bulb temperature.

The exchange surface is supplied with liquid water, which is transported from the
inside of the board by capillary action - the liquid migrates from regions with high
206 Fundamentals of wood drying

moisture content (liquid-gas interfaces in large pores) towards regions with low
moisture content (liquid-gas interfaces in small pores).

Boundary layers

T Pv
EXTERNAL FLOW

VAPOUR
HEAT

Capillary migration low MC


=
small radius

LIQUID FLOW

WOOD high MC
=
large radius

Figure 10.1. Constant drying rate period: the moisture migrates inside the medium
mostly by capillary forces, evaporation occurs at the exchange surface with a
dynamical equilibrium within the boundary layers between the heat and vapour
flows

b) Effect of external drying conditions

During the constant drying rate period the airflow supplies heat to the product,
allowing the water evaporated at the surface to be evacuated. In spite of the
complexity of the heat, mass and momentum transfers that exist within the boundary
layers, the heat and mass transfer coefficients are usually used to obtain a global
formulation of transfer between the board surface and the surrounding air. When the
vapour pressure remains much smaller than the total pressure, the following
expressions are obtained:
! !
qh .n = hh (Tsurf - T¥ ) (7.28)

! !
qv .n = hm cM v ( xv surf - xv ¥ ) (7.29)
!
In equations (7.28) and (7.29), qh is the heat flux at the exchange surface (W m-2),
!
n is the external unit vector, hh is the heat exchange coefficient (W m-2 K-1), T the
!
temperature, qv is the vapour flux at the exchange surface (kg s-1 m-2), hm the mass
Coupled heat and mass transfer 207

transfer coefficient (kg s-1) and xv the mole fraction of vapour. The indices surf and
∞ denote respectively the value at the product surface and the value of the
surrounding air (outside of the boundary layer).

During the constant rate stage the vapour pressure at the surface of the medium is
equal to the saturated water vapour:

Pvsurf = Pvs (Tsurf ) (7.30)

Equation (7.30), together with equations (7.28) and (7.29), lead to a dynamical
equilibrium where heat and mass fluxes are balanced through the latent heat of
vaporisation:

hh (Tsurf − T∞ ) = Lv × hm cM v ( xvsurf − xv∞ ) (7.31)

For low mass transfer rates, a simple relation exists between the heat transfer and
mass transfer coefficients (Welty et al. 1969). The following expression is a good
approximation of this analogy in the case of air/vapour mixture:

hh
= ρC p (7.32)
hm

This condition is usually fulfilled at atmospheric pressure and low temperature: in


such configurations, the vapour pressure is very small compared to the gaseous
pressure, hence mass transfer occurs mainly by diffusion of vapour molecules within
the boundary layer, as for heat diffusion. Using the analogy between heat and mass
transfer, equations (7.31) and (7.32) can be considered as a nonlinear set of
equations that defines the wet bulb temperature. Provided equation (7.30) remains
valid, the product temperature is equal to the wet bulb temperature and the mass flux
qv has a simple expression based on the heat supplied to the product:

hh (Td − Tw )
qv = (7.33)
Lv

where the indices d and w stand for dry bulb and wet bulb respectively and Lv is the
latent heat of vaporisation (J kg-1). Equation (7.33) demonstrates that, during the
constant rate stage, the mass flux is constant and determined by the external
conditions only (airflow, which determines the value of hh, temperature and relative
humidity of the air). Note that the nature of the product does not appear in this
expression.

Figure 10.2 depicts a graphical determination of the dynamic equilibrium at the


surface during the constant drying rate period. The corresponding point is obtained
as the intercept of the saturated vapour curve and the straight line coming from the
point representative of the external conditions with the negative slope given by the
analogy between heat and mass transfer in the boundary layer (equations (7.31) and
(7.32)). Once the wet bulb temperature is obtained, the magnitude of the drying rate
208 Fundamentals of wood drying

is directly proportional to the difference between the wet and the dry bulb
temperatures (equation (7.33)). Then, it is easy to anticipate the effect of external
conditions on the value of the drying rate, which:
- increases with an increase in temperature or a decrease in relative humidity,
- decreases and tends towards zero when the external point approaches the
saturated vapour curve.
Vapour concentration Saturated
vapour
curve

Slope given by the


analogy between heat
Corresponding point of the and mass transfer
borad surface during the
constant drying rate period

Point characteristic of the


external air conditions

cv∞∞

T d – Tw

Dew point Tw Temperature


Td

Figure 10.2. Graphical determination of the dynamic equilibrium obtained during


the constant drying rate period

c) The decreasing drying rate period

The constant drying rate period continues while the surface is supplied with liquid
and its duration depends strongly on both the drying conditions (magnitude of the
external flux) and the properties of the medium. The internal liquid flow can be
modelled by Darcy's law (permeability × gradient of liquid pressure). When the
moisture content decreases, the capillary forces usually increase, however the
permeability to the liquid phase decreases dramatically. On the whole, the liquid
flow tends to decrease for the same gradient of moisture content. Consequently, the
moisture content profile becomes steeper and steeper as the drying progresses, and
continues to do so until the liquid flow directed towards the surface ceases, which
marks the end of the first drying period. The time required for attaining this instant
depends on both the drying conditions and the board properties: a high initial
moisture content, a low drying rate and an uninhibited internal moisture migration
are conducive for a long constant drying rate period. Note also that the temperature
level of the board during the constant drying rate period (which is equal to the wet
bulb temperature) has an indirect effect: the higher the temperature, the easier the
internal moisture migration.
Coupled heat and mass transfer 209

Figure 10.3. Second drying period: A region in the hygroscopic range develops
from the exchange surface. In that region, both vapour diffusion and bound water
diffusion act. Evaporation takes place partly inside the medium and as a
consequence, a heat flux is driven towards the inner part of the board by conduction

Once the surface attains the hygroscopic range the vapour pressure becomes smaller
than the saturated vapour pressure. As a consequence, the external vapour flux is
reduced and the heat flux supplied to the medium is temporarily greater than that
necessary for liquid evaporation. The excess energy is used to heat the board via the
surface at first, and then throughout the inner part by conduction. A new, more
subtle, dynamical equilibrium takes place. The surface vapour pressure, and hence
the external vapour flow, depends on both temperature and moisture content. In
order to maintain the energy balance, the surface temperature increases as the
surface moisture content decreases. This leads to a decreasing drying rate and the
heat supplied by the airflow diminishes.

Two zones develop inside the porous medium: an inner zone where liquid migration
prevails and a surface zone, which is less efficient, where both bound water and
water vapour diffusion take place (Fig. 10.3). During this period, a conductive heat
flux must exist inside the board to increase the temperature and to evaporate the
liquid driven by gaseous diffusion.

The region of liquid migration naturally reduces as the drying progresses and finally
disappears. The process is finished when the temperature and the moisture content
attain, respectively, the outside air temperature and the equilibrium moisture
content.
210 Fundamentals of wood drying

120

Dry bulb temperature


Temperature (°C) or Moisture content (%)

100

80

Dew point temperature

60
Surface temperature
Moisture content Centre
40 temperature

20

0
0 2 4 6 8 10 12 14 16 18
Time (hour)

Figure 10.4. Convective drying below the boiling point of water: experimental
evolution of external drying conditions, average moisture content and internal
temperature. The so-called “surface temperature” is collected 2 mm below the
exchange surface (Spruce, sapwood, 32 mm thick)

Figure 10.4 summarises this coupling between heat and mass transfer in the case of
convective drying at low temperature (below the boiling point of water) when drying
a sapwood board of spruce 32 mm thick. This test was obtained in an accurate
laboratory kiln (Perré et al. 2000) in which the relative humidity is controlled
through the temperature of a water bath. This test distinguishes three periods:
- a heating period, with a linear increase in time of the heating air and the
board moisture content,
- a plateau at a constant temperature, with saturated conditions (the kiln and
the load are heated by the water bath, which is slightly higher than the air
temperature during this phase),
- a drying period.
During the plateau period all temperatures have almost the same value. Note that,
due to the saturated conditions, the dew point temperature is equal to the wet bulb
temperature and the moisture content is constant during this period. When drying
starts, the surface of the board is fully saturated: the constant drying rate period is
clearly observable in this case. Consistently, the wet bulb temperature is higher than
the water temperature, which imposes the dew point temperature. Nevertheless, one
can see a slight increase in temperature after some hours, which can be explained by
the concept of a thin (or microscopic) dry layer at the surface (see chapter 9). At 14
hours the board temperature markedly increases and from that point a macroscopic
dry layer, as depicted in figure 10.3, develops.
Coupled heat and mass transfer 211

d) The effect of internal pressure on mass transfer

In order to reduce the drying time without decreasing the quality of the dried
product, the drying conditions must be such that the temperature of the product is
above the boiling point of water. Such conditions ensure that an overpressure exists
within the material, which implies that a pressure gradient drives the moisture
(liquid and/or vapour) towards the exchange surfaces (Lowery 1979, Kamke and
Casey 1988). At the atmospheric pressure, the boiling point of water equals 100°C.
Consequently, in order to obtain an internal overpressure, the temperature of the
porous medium must be above that level at some point of the process. This is exactly
the aim of convective drying at high temperature (moist air or superheated steam)
and a possible aim of contact drying or dielectric drying (microwave or high
frequency). However, as shown in figure 10.5, it is possible to reduce the boiling
point of water by decreasing the external pressure and, consequently, to obtain a
high temperature configuration with relatively moderate drying conditions. This is
the principle of vacuum drying, which is particularly relevant for materials that
would be damaged by high levels of temperature.

When an overpressure exists inside a board during drying the high anisotropy ratios
induce bi-dimensional heat and mass transfer phenomena, whereby heat is often
supplied across the thickness while, in spite of the length, the effect of the pressure
gradient on gaseous (important for low MC) or liquid migration (important for high
MC) takes place in the longitudinal direction (Fig. 10.6). This is obviously a result
of the anatomical features of wood. In the case of very intensive internal transfer the
endpiece can be fully saturated and sometimes, moisture can leave the sample in the
liquid state, which is quite easy to observe during for example, microwave heating.

S a tu ra te d v a p o u r
A tm o s p h e r ic p re s s u r e
100
Pressure (kPa)

B o ilin g te m p e ra tu re
50

E x te rn a l p re s s u re

0
0 20 40 60 80 100 120
T e m p e ra tu re (°C )

Figure 10.5. Vacuum drying seeks to reduce the boiling point of water to generate a
high temperature configuration with moderate drying conditions
212 Fundamentals of wood drying

HEAT Vessel or
tracheid
VAPOUR

LIQUID

Pits

Endpiece
fully saturated

Liquid evacuation
possible in
OVERPRESSURE microwave heating

Figure 10.6. Depending on the MC, an overpressure within the board induces liquid
and/or gaseous flow; In addition, wood being strongly anisotropic, the majority of
the flow occurs in the longitudinal direction (see the magnified views)

10.2.2 Drying with internal vaporisation

Species like Oak play an important role in drying innovation as a result of the very
fact that their drying is quite a difficult procedure. On the other hand, species prone
to fast and good drying allow basically any audacious technology to be invented.
Several technological innovations are currently proposed to obtain a fast and good
(in terms of cost and quality) drying process. Note that most of them originate from
ideas patented several decades ago:
- development of pre-drying kilns able to replace the storage operation with a
better control,
- use of a vacuum superheated steam drier where good results were obtained with
27 mm thick boards, but larger sections remain problematic (Ressel, 1994, Joyet
and Meunier, 1996),
- use of HF/Vacuum dryers, which allows the combination of two major
advantages a) the enhancement of the internal moisture transfer by exceeding
the boiling point of water (role of vacuum) and b) providing a good supply of
energy to the wood pieces, whatever their thickness. Several prototypes are
available in laboratories. Some industrial prototypes not only exist for softwood
but also for Oak (Smith et al., 1994, Avramidis, 1999, Kobayashi et al., 1999,
Resch and Gautsch, 2000),
Coupled heat and mass transfer 213

- development of very fast drying procedures able to replace the conventional


batch process with a continuous drying method (Antti and Perré, 1999),
- use of pressure cycles to enhance the departure of liquid and vapour from the
wood pieces (Hayashi et al. 1994, Schill et al., 1994).

In chapter 1 it was apparent that the drying of wood involves many complex and
coupled mechanisms. Consequently, the improvement of drying lies in a subtle
compromise between several outputs involving time, cost and product quality. For
this reason the heat and mass transfer mechanisms are able to provide answers to
only one part of the question. Note however, that most of the efficient processes
described above are based on the same principle: to generate a gradient of internal
pressure in the board to enhance moisture migration, which is possible as a result of
the large permeability values of most wood species in the longitudinal direction.

a) Convective drying at high temperature

The simplest way to induce an internal overpressure within a board is to use


convective drying at a temperature higher than the boiling point of water (100°C).
However, one has to keep in mind that two conditions are required for an
overpressure to develop inside the board: the presence of free water along with a
temperature above 100°C. This explains why drying with saturated steam is better
than with moist air and why the overpressure develops only during the second
drying period (Perré et al. 1993). Using a temperature of around 120°C, the total
drying time can be reduced to 24 hours for 27 mm thick broads of softwood, while
using 150°C, the time can be reduced to about 6 hours (Basilico and Martin, 1984).

Figure 10.7 exhibits drying experiments carried out on spruce with superheated
steam at 150°C that are representative of the trends generally observed elsewhere
(Salin 1989, Pang et al. 1994, Perré and Martin, 1994). Notice the very long first
drying period observed for sapwood and the absence of this period for heartwood.
The test on sapwood also confirms that the internal overpressure develops only
during the second drying period. At this time (around 350 minutes) an important
overpressure follows the temperature increase, which begins to disappear once the
medium enters the hygroscopic range.

The results obtained for heartwood are quite different because no first drying period
can be observed, with a short plateau at the boiling point detectable at the rear end of
the board (T8). Consequently, the overpressure remains high (especially for the
centre pressure P3) up to the end of the drying. Note further that the maximum
pressure is higher for heartwood than for sapwood. The drying kinetics obtained for
these two experiments (Fig. 10.8) highlights the much higher initial moisture content
(170% against 60 %) and the long constant drying rate period observable for
sapwood. These two combined features allow the drying time to be almost the same
for the two broads: the drying curves cross each other at 450 minutes of drying. This
trend is quite common for softwood (Salin 1989).
214 Fundamentals of wood drying

150 1.0
T1
T8
130 0.8
T5
T3

Overpressure/Patm
Temperature ( C)
o

110 0.6

P3
90 0.4

P5
70 0.2

50 0
0 100 200 300 400 500 600 700

TIME (min)
D ifferent locations
A ir flow in the sam ple
5
4
1 3 30 m m
0 8
1.0 m

150 1.25
T1
T8
130 1.00
T5
T3
Overpressure/Patm
Temperature ( C)
o

110 0.75

P3
90 0.50

70 0.25
P5

50 0
0 100 200 300 400 500

TIME (minutes)

Figure 10.7. Experiment on sapwood (Spruce dried with superheated steam at


150°C). Temperature and internal pressure at different locations. Sapwood (top)
and heartwood (bottom)
Coupled heat and mass transfer 215

150
Sapwood
Moisture content (%)

100

Heartwood
50

0
0 200 400 600

Time (minutes)

Figure 10.8. Drying kinetics for same test shown in figure 10.7

b) Vacuum drying

High temperature levels experienced during drying are sometimes unacceptable


because collapse or discoloration may affect the end product quality. For species
sensitive to temperature, the sought-after pressure gradient can be obtained at lower
temperatures simply by reducing the boiling point of water. This possibility opened
the door to several vacuum drying techniques. Due to the reduced pressure level, the
problem in this case is to supply energy to the board (Perré et al., 1995). Therefore,
the vacuum drier must achieve two main objectives: to ensure an underpressure and
to supply the necessary energy to the stack. The method by which thermal energy is
supplied to the material characterises the vacuum drying process.
- Heating of wood through convection during periods at the atmospheric pressure
(discontinuous vacuum kilns),
- Heating of wood through conduction using heated plates,
- Heating of wood through radiation using infrared emitters,
- Heating of wood through convection with superheated steam (moderate level of
vacuum with a very high gas velocity),
- Heating of wood through volumetric heating (radio-frequency or microwave),
In the later case, one secondary effect can be of importance: due to volumetric
heating the thermal gradient, which gives rise to thermo-migration, can be positive,
(with the core warmer than the surface) rather than negative in the case of
convective heating (Fig. 10.3).

Vacuum drying is often used for high valuated species or for large sections
(Kanagawa 1993, Perré et al. 1995, Wengert and Denig 1995). The physics of
216 Fundamentals of wood drying

vacuum drying is well known and can also be included in the model. The most
important features of heat and mass transfer in partial vacuum are:
- the way to supply energy to the product,
- the regime of gas migration in the porous medium.

We present here an example of vacuum drying with radiative heating (Fig. 10.9).
Details of the experimental procedure and other experiments can be found elsewhere
(Perré et al. 2004). The present test has been carried out on a sapwood board of fir
(Pchamber = 0.2 Patm). Due to the low pressure, the boiling point of water is reduced
(to around 60°C here) and it can be observed that after the first decrease of pressure
from the atmospheric value, all of the curves exhibit the same trends as those noted
for high temperature drying with superheated steam.
180 60
Moisture content (%) or Temperature (°C)

Surface temperature
150 50
Moisture content Center temperature

Overpressure (kPa)
120 40
Center pressure

90 30

60 20
Surface pressure
30 10

0 0
0 100 200 300 400 500

Time (minutes)

Figure 10.9. Vacuum drying of fir (Abies alba) with radiative heating. Drying
kinetics, temperature and pressure at different locations

c) Microwave heating at atmospheric pressure

In the case of microwave or radio-frequency drying, the energy is supplied directly


to the core of the sample by electromagnetic heating. Consequently, the medium
temperature is no longer determined by the external conditions. No plateau can be
observed on the evolution of the temperature. In addition, the heat flux does not
depend on the thermal gradient: important drying rates with internal vaporisation
can be obtained without great differences between the surface and inner
temperatures. Indeed, with microwave heating, the surface is often cooled by the air
flux, leading to a positive thermal gradient accelerating the mass migration towards
the surface. Because two driving forces act in the same direction (moisture gradient
and thermal gradient), the drying rate might be much higher in the hygroscopic
range, as compared to convective drying, without necessarily attaining the boiling
point of water in the board (Constant et al. 1992).
Coupled heat and mass transfer 217

Typical tests carried out with microwave heating distinguish four drying periods:
- a heating period: at first the energy supplied heats the medium without mass loss,
- a “streaming” period: when the temperature reaches and overtakes the boiling point
of water the internal pressure becomes high enough for draining liquid water out of
the medium (pumping effect),
- an enthalpic period: when the moisture content decreases, vapour migration takes
the place of liquid migration. The mechanism becomes simpler whereby the vapour
flux sustained through vaporisation by the amount of energy leaves the medium by
diffusion-convection (or pure convection). The analytical model proposed in Perré et
al. (1993) can be used in this case to provide an explanation of the mass migration
mechanisms and internal pressure prediction.
- a “burning” period: finally, when insufficient water remains inside the medium, the
temperature increase can be quite spectacular. Indeed, the dielectric loss factor
(imaginary component of the permitivity) often increases with temperature implying
that the drier part of the medium becomes hotter and receives more energy, which
eventually leads to a thermal runaway effect.

The four drying periods can be clearly observed from the experimental test carried
out on a small spruce sample (Fig. 10.10). Due to the intensive volumetric heating
the pressure level is much higher than the one observed for convective drying at
high temperature. In fact pressure levels as high as 3 times the atmospheric pressure
are obtained for softwood.

140 5
Center temperature

120
4
Temperature (oC) or MC (%)

100
Pressure/Patm

3
80

Surface temperature
60
Center pressure Endpiece temperature 2

40 Mid-length Pressure
1
Patm
20
Average MC

0 0
0 20 40 60

Time (min)

Figure 10.10. Microwave drying of spruce (Experiment; incident power close to


100 W)
218 Fundamentals of wood drying

10.3 Physical formulation

A hygroscopic porous medium such as wood consists of solid, liquid, gas and bound
water phases. The transfer mechanisms must be derived for each phase separately by
carefully taking into account the complex interactions that exist between them.
Typically, the geometry of the porous medium is so complex that the equations must
be written at the macroscopic scale, which leads to the definition of empirical laws
of migration that can be expressed by averaging over representative volumes
(REVs). At this level, the porous medium can be perceived as a fictitious,
continuous medium and the conservation laws obtained at this scale appear similar
to those valid for a continuous medium, with the exception that effective coefficients
are used in their definitions (Whitaker 1977, 1998).

LIQUID DRY AIR


+
W ATER VAPOR

SOLID
+
BOUND W ATER

REPRESENTATIVE ELEMENTARY VOLUME

Figure 10.11. Schematic representation of a porous medium

Several versions of the macroscopic drying equations have been proposed in the
literature for simulating the wood drying process. The first fundamental difference
in these formulations is the number of state variables used to describe the medium:
* one variable: moisture content (MC), or an equivalent variable such as
saturation or water potential,
* two variables: MC or equivalent, and temperature (T), or an equivalent
variable such as enthalpy,
* three variables: MC or equivalent, T or equivalent and gaseous pressure
(Pg), or an equivalent variable such as air density, or intrinsic air density.
Coupled heat and mass transfer 219

a) The Comprehensive three equation model – (Model 1)

The three corresponding sets of equations are now presented, starting with the most
comprehensive macroscopic formulation, which will be referred to as Model 1 -
three state variables. At present, researchers using model 1 agree with the
macroscopic formulation to be used. The governing set of equations originates for
the most part from Whitaker’s research (Whitaker 1977, 1980) with minor
modifications required to account for the wood properties and drying with internal
overpressure (Perré and Degiovanni 1990). In particular, the reader must be aware
that all variables are averaged over the REV, hence the expression “macroscopic”.
This derivation assumes the existence of a representative volume, which is large
enough for the averaged quantities to be defined and small enough to avoid
variations due to macroscopic gradients and non-equilibrium configurations at the
microscopic level. This is why that although the formulation can be used to simulate
a number of different configurations, this set of equations does have some
limitations (Perré 1998).

Water Conservation


∂t
(ε w ρw + ε g ρv + ρ b ) + ∇ ⋅ ( ρ w vw + ρv vg + ρb vb ) = ∇ ⋅ ( ρ g Deff ∇ωv ) (7.34)

Energy Conservation


∂t
(ε w ρw hw + ε g (ρv hv + ρa ha ) + ρ b hb + ρo hs − ε g Pg )
(
+ ∇ ⋅ ρ w hw vw + ( ρv hv + ρ a ha ) v g + hb ρb vb ) (7.35)

( )
= ∇ ⋅ ρ g D eff ( hv ∇ωv + ha ∇ωa ) + λ eff ∇T + Φ

Air Conservation


∂t
(ε g ρa ) + ∇ ⋅ ( ρ a vg ) = ∇ ⋅ ( ρ g Deff ∇ωa ) (7.36)

where the gas and liquid phase velocities are given by the Generalised Darcy’s Law:

K ℓ kℓ
vℓ = − ∇ϕℓ , ∇ϕℓ =∇Pℓ − ρℓ g ∇χ where ℓ = w, g (7.37)
µℓ

the quantities ϕ are known as the phase potentials and χ is the depth scalar. Φ is
the volumetric power density term that is used in the instance of microwave heating.
All other symbols have their usual meaning.
220 Fundamentals of wood drying

A more detailed description of these equations and the related assumptions can be
found elsewhere (Perré 1996, Turner and Perré 1996, Perré and Turner 1999). Note
that because this formulation accounts for the evolution of internal pressure through
the air balance (equation (7.36)), the set of equations proves to be very powerful,
and is able to deal with numerous configurations involving intense transfers,
including high temperature convective drying, vacuum drying and dielectric drying.

Boundary conditions

For the external drying surfaces of the sample the boundary conditions are assumed
to be of the following form:

 1 − x∞ 
J w x =0+ ⋅ nˆ = hm cM v ln  
1− x 
 v x = 0+ 
J e x =0+ ⋅ nˆ = h(T x =0+ − T∞ ) (7.38)
Pg = Patm
x =0+

where J w and J e represent respectively the fluxes of total moisture and total
enthalpy at the boundary respectively, and x denotes the position from the boundary
along the external unit normal.

b) The two equation model for heat and mass transfer – (Model 2)

Model 2 describes that drying of a porous medium that is represented by only two
dependent variables, namely temperature or its equivalent and moisture content or
its equivalent. This model can be analysed as a simplification of model 1. There is
no account taken of the internal gaseous pressure and equation (7.36) becomes
redundant. Darcy's law (equation (7.37)) subsists solitarily for the liquid phase, in
which only capillary forces act. Assuming the capillary pressure to be a function
dependent only upon moisture content, instead of moisture and temperature, both
vapour and liquid migration can be expressed as diffusion mechanisms that involve
so-called global “pseudo-” diffusion coefficients K v and K ℓ . This model is
obviously unable to deal with the effect of internal pressure on liquid and gaseous
migration and hence is not suitable for drying configurations involving intense
migration. However, through the difference of enthalpy between the vapour and
liquid phases, this model does account for the latent heat of vaporisation and is
therefore able to deal with the most important feature of drying, which is the
coupling between heat and mass transfer, i.e., latent heat of vaporisation and the
relation between temperature and saturated vapour pressure.
Coupled heat and mass transfer 221

Water Conservation


∂t
( ρ0 X ) = ∇ ⋅ ( ρ K ∇X ) + ∇ ⋅ ( ρ K ∇X )
v v ℓ ℓ (7.39)

Energy Conservation


(
 ρ0 ( h0 + X .hℓ )  + ∇ ⋅ ( ρ ℓ hℓ qℓ + ρ v hv qv ) = ∇ ⋅ λ eff ∇T + Φ
∂t 
) (7.40)

where (
qv = ∇ ⋅ ρv K v ∇X ; ) (
qℓ = ∇ ⋅ ρℓ K ℓ ∇X )
Boundary conditions

 1 − x∞ 
J w x =0+ ⋅ nˆ = hm cM v ln  
1− x 
 v x = 0+ 
J e x =0+ ⋅ nˆ = h(T x =0+ − T∞ ) (7.41)
Pg = Patm
x =0+

c) The one equation model for mass transfer only – (Model 3)

From model 2, it is still possible to imagine one further simplification of the theory,
whereby the product temperature is assumed constant and equation (7.35) now
becomes redundant. In this case the product temperature is supposed to follow the
air flow temperature evident in the kiln. It is no longer necessary to differentiate
liquid flow from vapour flow and the use of one global diffusion coefficient
becomes possible, which leads to Model 3.

Water Conservation

∂X
∂t
(
= ∇ ⋅ K ∇X ) (7.42)

Boundary conditions

(
J w x =0+ ⋅ nˆ = α cv x = 0+ − c∞ ) (7.43)

In equation (7.43) the mass flux at the interface is written in a specific manner.
Indeed, because the surface temperature cannot be known using this one-equation
model, the external mass flux is simply obtained as the product of one coefficient
222 Fundamentals of wood drying

(often called an “emission factor”) by the difference between two “concentrations”,


i.e. moisture content at the surface and the equilibrium moisture content.

Recalling that the objective of drying is to remove the moisture from the product,
the concept of a “one equation” model appears attractive since it accounts for the
moisture migration within the medium and the moisture flux at the boundary.
Unfortunately, this simplistic argument encouraged, in our opinion, too many
scientists to adopt model 3. Figure 10.12 depicts the one-dimensional computational
results obtained for convective drying at low temperature (Dry bulb = 50°C, dew
point = 30°C, board thickness = 30 mm). Model 2 (two equations) exhibits a
constant drying rate period during which the board temperature equals the wet bulb
temperature. Model 3 (one equation) is not capable of capturing the reduction of
vapour pressure at the surface, which results from the surface cooling due to
evaporation. Consequently, for Model 3, the drying process is much faster than it
should be when reasonable physical values are used. Interestingly, this error can be
corrected, or at least the evolution of the averaged moisture content can be made
more realistic, by fitting the mass exchange coefficient hm (see figure 10.12, curve
obtained with hm divided by a factor of 10).

160 60

One equation (mass)


One equation (hm divided by 10)
Two equations (heat and mass)

120 50
Moisture content (kg/kg)

Temperature
Temperature (°C)

80 40

Averaged moisture content


40 30

0 20
0 20 40 60 80 100

Time (hours)

Figure 10.12. Convective drying at low temperature (Dry bulb = 50°C, dew point =
30°C, board thickness = 30 mm, h = 14 W.m-2.°C -1; hm = 0.014 m.s-1): Model 2
versus model 3

However, this scaling implies that the mass transfer coefficient no longer retains its
physical meaning and, perhaps more crucial, a new value has to be fitted each time
the drying conditions change. For example, the factor 10 that allowed a reasonable
drying curve to be obtained at 50°C is no longer valid for a dry bulb temperature of
Coupled heat and mass transfer 223

70°C (Perré 1999). In conclusion, because one variable prevents the strong coupling
that always exists between heat and mass transfer within the medium to be captured,
Model 3 is a crude choice that the authors believe must absolutely be discarded,
even for very simple drying configurations.

Model 2 (two equations) provides a much more realistic account of the drying
process and allows numerous configurations to be simulated, including conventional
kiln drying. The authors recommend that this choice is the correct one to begin with
when simulating drying at moderate temperatures, or for providing on-line help in
kiln control. Nevertheless, the users of such a model must be aware of its
limitations. From figure 10.6, it is easy to understand that this formulation fails as
soon as the internal pressure has a significant impact on transfer. Model 2 provides
no mechanism by which to simulate high temperature drying, vacuum drying,
microwave heating, or, simply, full saturation of a wood sample by cycles of
vacuum and pressure.

Figure 10.6 provides the perfect illustration of why transport in the longitudinal
direction must be taken into account in the modelling. Numerous configurations
have been computed with great success using model 1 (comprehensive set) with two
spatial dimensions (length and thickness of the board) and several examples have
been published that provide thorough comparisons of theoretical and experimental
results, including temperature and pressure measurements within the board (Stanish
et al. 1986, Constant et al. 1996, Perré 1996, Perré and Turner 1996, Johansson et
al. 1997). One example of high temperature convective drying (Tdry = 140°C; Tdew =
85°C) has been chosen here to elucidate the possibilities offered by model 1 with
two spatial dimensions. This example depicts all of the trends observed during
drying of wood when the internal pressure is important. Specifically, at the end of
the constant drying rate period (Fig. 10.13, around 3 hours), the surface temperature
increases at first, followed thereafter by the centre temperature. Once the boiling
point of water is attained, an overpressure develops within the board and because the
longitudinal permeability of wood is much higher than the transverse permeability
(by a factor 1000 in this simulation), this overpressure gives rise to an important
longitudinal moisture migration towards the endpiece. The latter is re-saturated and
remains fully saturated for a long period (3.5 to 7.5 hours). Typical bi-dimensional
transfer occurs - heat is supplied to the product mainly through the thickness and
moisture partially moves along the longitudinal direction (Fig. 10.14).
224 Fundamentals of wood drying

1m

end
30 mm

centre surface

160 2.0
Tsurf Tend
MCend
Temperature (°C) or MC (kg/kg)

120

Internal pressure (P/Patm)


Tcentre

80 1.5

MCaverage MCcentre
Pcentre
MCsurf

40

0 1.0
0 4 8 12 16 20

Time (hours)

Figure 10.13. Convective drying at high temperature (Dry bulb = 140°C, dew point
= 85°C, board thickness = 30 mm, board length 1 m): model 1, two spatial
dimensions

As a consequence of this moisture migration, the endpiece temperature remains


close to the wet bulb temperature for a corresponding duration. This effect has often
been observed during experiments carried out using thermocouples (Perré et al.
1993). During the last stage of drying the overpressure remains important with a
diminishing amount of liquid water inside the product. However, longitudinal
migration of moisture is still an important transport mechanism as a result of the
high percentage of vapour that exists in the gaseous flux.
Coupled heat and mass transfer 225

50

MC
40
1.0
30

Len
gth 20
0.5
( cm)

10 0.0
2 )
m
1 (c
h
0 0 idt
W

50
150

Temp
40
125

30
100
Len
gth 20
(cm 75
)
10 50
2 )
m
1 (c
h
0 0 idt
W

50
Pres

40 1.8

30 1.6

Len 1.4
gth 20
( cm) 1.2

10 1.0
2 )
m
1 (c
h
0 0 idt
W

Figure 10.14. High temperature drying (140°C/85°C) of sapwood. Carpet plot at 5


hours of drying. Model 1 (Comprehensive physical formulation), two spatial
dimensions. Note the internal overpressure, re-saturation of the endpiece, thermal
conduction along the thickness and the endpiece close to the wet bulb temperature

Because a material model based on a comprehensive physical formulation has been


used, the differences between sapwood and heartwood can be simulated by changing
the material properties (Perré and Martin 1994). These differences are realised
through two sets of parameters - the permeabilities and the initial moisture content.
226 Fundamentals of wood drying

50

MC
40
1.0
30

Len
gth 20
0.5
(cm
)
10 0.0
2 )
cm
h(
1
0 0 idt
W

50
150

Temp
40
125

30
100
Len
gth 20
( cm) 75

10 50
2 )
m
1 (c
h
0 0 idt
W

50
Pres

40 1.8

30 1.6

Len 1.4
gth 20
( cm) 1.2

10 1.0
2 )
m
1 (c
h
0 0 idt
W

Figure 10.15. High temperature drying (140°C/85°C) of heartwood. Carpet plot at


5 hours of drying. Model 1 (Comprehensive physical formulation), two spatial
dimensions. Note the high value of internal pressure and the absence of endpiece re-
saturation

Here the values of initial moisture content can be obtained directly from
measurements (150% to 200% for sapwood, 40% to 80% for heartwood) and the
permeability the values used to differentiate sapwood from heartwood are based on
considerations concerning pit aspiration (see chapter 7).
Coupled heat and mass transfer 227

Putting these new values in the model, all of the experimentally observed trends
(Salin 1989, Pang et al. 1994, Perré and Martin, 1994) can be captured during the
simulations (Figs 10.14 and 10.15). In particular, no re-saturation of the endpiece is
observed and the constant drying rate period disappears. The overpressure develops
from the beginning of the process and remains high up to the end of drying. The
maximum pressure is higher for heartwood than for sapwood. These results are in
entire agreement with experimental observations (Fig. 10.16).

Sapwood Heartwood

Figure 10.16. A stack of boards after some hours of drying at high temperature

Using the same formulation (Model 1) with only one spatial dimension (thickness),
enables the longitudinal mass transfer to be computed (Fig. 10.17). The internal
overpressure develops without an important effect on mass transfer (the transverse
permeability is too small for a significant flux to be obtained). Both internal
temperature and internal pressure increase faster than for a two-dimensional
simulation, however the global drying curve is much slower.

On the other hand, when using model 2 for the simulation, the internal pressure is
assumed to have the same value as the external pressure. Consequently, no mass
transfer can be induced by internal vaporisation. Even if supplied by capillary
action, which explains the plateau near the wet bulb temperature, the endpiece
moisture content continuously decreases throughout the process (Fig. 10.18). Due to
the lack of important longitudinal mass transfer, the drying curve is much slower
than the one depicted in figure 10.13.
228 Fundamentals of wood drying

160 3.0
Tsurf
Temperature (°C) or MC (kg/kg)

120 2.5

Internal pressure (P/Patm)


Tcentre

Pcentre
MCcentre
80 2.0
MCaverage

MCsurf

40 1.5

0 1.0
0 4 8 12 16 20

Time (hours)

Figure 10.17. Convective drying at high temperature (Dry bulb = 140°C, dew point
= 85°C, board thickness = 30 mm, board length 1 m): Model 1, one spatial
dimension

160 2.0

Tsurf
Tcentre
Temperature (°C) or MC (kg/kg)

120
Internal pressure (P/Patm)

Tend

80 1.5
MCcentre

MCsurf MCaverage

40
MCend

Pcentre

0 1.0
0 4 8 12 16 20

Time (hours)

Figure 10.18. Convective drying at high temperature (Dry bulb = 140°C, dew point
= 85°C, board thickness = 30 mm, board length 1 m): model 2, two spatial
dimensions

The same physical formulation has also been computed in three dimensions (Model
1, three spatial dimensions), which permitted new wood drying mechanisms to be
Coupled heat and mass transfer 229

exhibited and investigated (Perré and Turner 1999). The result of this computational
model is a comprehensive description of both physical and geometrical aspects of
the wood drying process. The authors believe that such a model has probably great
potential to play an important role in the advancement of wood drying in the future,
provided the computational time can be dramatically reduced.

To conclude this section devoted to the choice of the relevant physical formulation
to simulate heat and mass transfer during drying, a summary of the main findings
drawn from the simulations proposed above and from the literature is summarised in
table 10.1. Finally, one must keep in mind that the computational effort dramatically
changes according to these choices (Table 10.2). One must also note that because all
of the drying equations are strongly coupled and highly nonlinear, the use of non-
efficient computational strategies can lead to computational times orders of
magnitude (a factor 10, possibly 100) higher than the values reported in table 10.2.

Table 10.1. Heat and mass transfer during wood drying: simple guideline to help
with the choice of a physical formulation and geometrical description

dependent Model 3 Model 2 Model 1


variables One variable Two variables Three variables
(MC) (MC + T) (MC + T + Pg)
spatial dimension
1-D Comprehensive physics,
Simple but very poor
(Thickness) Suitable for a coarse but unrealistic effect of the
physics (no coupling
description of convective internal pressure (no
between heat and mass
drying at low longitudinal direction)
transfer) hence
(1) temperature misleading results at high
misleading results
temperature
2-D Correct description of Comprehensive physics,
Correct description of
(Thickness + the section shape but but unrealistic effect of the
the section shape.
Width) very poor physics (no internal pressure (no
Suitable for convective
coupling between heat longitudinal direction)
drying at low
and mass transfer) hence misleading results at high
temperature
misleading results (1) temperature
2-D Useless additional
(Thickness + computation effort Comprehensive physics
Length) compared to 1-D Very Superfluous additional and good geometrical
poor physics (no computational effort description. Suitable for
coupling between heat compared to 1-D almost all drying processes
and mass transfer) and (effect of internal pressure)
misleading results
3-D Good geometrical
Caution: complex The best possible
description. Can be
numerical calculation for description at the
useful for calculating the
very poor physics and macroscopic level
global deformation of
misleading results (geometry and physics)
entire boards
(1) : Suitable only for diffusion in the hygroscopic range of very thick boards.
230 Fundamentals of wood drying

Table 10.2. Order of magnitude of the computer time required for a complete drying
simulation (Personal computer, Pentium IV 3GHz, state-of-the-art numerical
techniques)

dependent Model 3 Model 2 Model 1


variables One variable Two variables Three variables
Space (MC) (MC + T) (MC + T + Pg)
dimension
1-D A mouse-click Around 0.1 second Less than seconds

2-D Around one second Some seconds Less than one minute

3-D Around one minute Some minutes Less than one hour

10.4 Case studies in wood drying

In this section, some of the drying modelling work reported in the literature by the
authors over the last five years is briefly summarised. In particular, three recent
wood drying case studies are presented that concern low temperature convective
drying using the heterogeneous formulation for an RT cross-section; combined
microwave and convective drying; and vacuum drying with radiative heating. These
case studies have been chosen specifically to compare theoretical results obtained
using Model 1 with the experimental observation given in section 10.2.

10.4.1 Case Study 1 – The heterogeneous model

As stated previously, the properties of wood depend not only on the species of tree
but also on position within the tree. A common approach adopted in porous media
research to overcome this variability is to utilise homogenisation techniques
(Hornung 1996). Perré and Turner (2001b) used exactly this approach to determine
the important macroscopic wood properties based on a combination of microscopic
observation and knowledge of the morphology of the porous structure. In that work
the geometrical description of the pore structure was taken directly from image
analysis (Perré and Badel, 2000, Perré 2005) performed on anatomical views and the
average effect of structural variations within the annual ring of softwood was
captured.

This approach enabled a set of correlations for capillary pressure Pc ( ρ0 , X w , T ) ,


bound liquid diffusivity Db ( ρ0 , X b , T ) , thermal conductivity K eff ( ρ0 , X , T ) and
absolute permeability k w ( ρ0 ) , k g ( ρ0 ) to be postulated (we refer the reader to the
two part paper by Perré and Turner, 2001ab for complete details). In light of this
foundation research work, the drying equations presented above in equations (7.34)–
(7.37) were extended to account for material heterogeneity through the apparent
Coupled heat and mass transfer 231

density of the porous medium ρo ( x ) and via the density variation of the material
properties (Perré and Turner 2002).

Simulation results using the heterogeneous version of the drying model formulation
are reported for the drying of a sample of softwood (RT cross-section 6 cm × 2 cm)
that contains nine growth rings. The initial average moisture content of the sample
was 170% and the average porosity and apparent density were 0.703 and 455 kg m-3
respectively, with the dry bulb and wet bulb temperatures 80ºC and 65ºC. The air
velocity used for the simulations was 3 m/s, yielding approximate heat and mass
exchange coefficients of 25 W m-2 K-1 and 0.025 m s-1 respectively. The board
sample represents a quarter sawn section of maritime pine depicted in figure
10.19(a).

(a)

(b)
Figure 10.19. (a) Image of maritime pine depicting nine growth rings, (b)
Computational mesh constructed from image.

In this case of heterogeneous model, the initial moisture content field is not uniform
and has to be computed prior to the commencement of the drying process (Perré and
Turner, 2002). Indeed, due to the large density variation that can be observed in
figure 10.19, it can be concluded that the pores in the latewood component of the
annual ring are smaller than in the earlywood component. Consequently, stronger
capillary forces become evident in latewood.

To account for these anatomical features, the capillary pressure curve supplied to the
model depends not only on moisture content and temperature, but also on the local
232 Fundamentals of wood drying

variation of density (Perré and Turner 2001a). This dependency allows the main
features of softwood regarding liquid pressure to be captured:
- low values of capillary pressure provide a moisture content higher in early
wood than in late wood (i.e. when approaching full saturation).
- high values of capillary pressure provide a moisture content higher in latewood
than in earlywood because of the smaller size of the pores in this part of the
annual ring (when the liquid content decreases towards the fibre saturation
point).

By noting further that all capillary forces must be in equilibrium when the medium
is in its initial state, the capillary pressures at each node of the computational mesh
must be equal. Let Pco be this equilibrium capillary pressure, then the following
equation must hold for each control volume:

PcP ( S w , T , ρo ) = Pco , ∀p = 1, 2,… , N . (7.44)

Figure 10.20 depicts the initial moisture content field obtained using equation (7.44)
for the present section (Fig. 10.19). In order to approach full saturation, a small
value of Pco is chosen in this equation.

Initial MC field

0
Moisture
0.01 2.30
2
0.02 2.04
Moisture

1 1.79
0.03 1.53
dth 1.28
0
0 0.04 Wi 1.02
0.005
0.77
Th 0.01 0.05
ick 0.015 0.51
ne
ss 0.02 0.26
0.06 0.00

Figure 10.20. Initial moisture content field, computed as an isovalue of the liquid
pressure

In this section the simulation results for an 80°C/65°C (dry bulb and wet bulb
temperatures respectively) drying schedule are presented and the important aspects
of the low temperature drying process exhibited.
Coupled heat and mass transfer 233

2 hours

0
0.01
2
0.02
Moisture

1
0.03
dth
0
0 0.04 Wi
0.005
T h 0.01 0.05
ick 0.015
ne
ss 0.02
0.06

5 hours

0
0.01
2
0.02
Moisture

1
0.03
dth
0
0 0.04 Wi
0.005
T h 0.01 0.05
ick 0.015
ne
ss 0.02
0.06

10 hours

0
0.01
2
0.02
Moisture

1
0.03
dth
0
0 0.04 Wi
0.005
T h 0.01 0.05
ick 0.015
ne
ss 0.02
0.06

Figure 10.21. Evolution of the wood moisture content throughout drying (2, 5 and
10 hours of drying)
234 Fundamentals of wood drying

As mentioned previously, two distinct periods are evident during drying, namely the
constant rate period and the falling rate period. For the maritime pine sample and the
drying conditions studied here, the constant rate period appears to be 5 hours in
duration, and thereafter the falling rate period commences and continues throughout
drying. During the constant rate period, heat supplied by the drying air is used
primarily to transform water into vapour and the board temperature remains close to
the wet bulb temperature of 65°C. Liquid migration is primarily due to capillarity
and in figure 10.21 this phenomenon can be observed to distinctly follow the growth
ring pattern evident in the virtual board, draining at first from regions with high
initial moisture content and low density (earlywood), followed by the regions with
low initial moisture content and high density (latewood).

Once the surface moisture content reaches the hygroscopic range a reduction in the
external vapour flux causes the overall drying rate to slow down. The drying air now
begins to heat the board and the sample case appears dry (see figure 10.21 at 10
hours). The board has now entered the falling rate drying period whereby two zones
are clearly evident - an inner zone where liquid migration again distinctly follows
the growth ring pattern, and a surface zone where only bound water migration and
water vapour diffusion occur. These phenomena can be unmistakably seen in figure
10.22 at the exhibited times from 15 to 25 hours.

One can observe on the carpet plots in figure 10.22 that all internal regions with a
high moisture content are located in the latewood part. In fact, in the core of the
section the moisture flux remains small enough to analyse the moisture content field
as an isovalue of the liquid pressure (roughly the capillary pressure in this
configuration of low temperature drying). For this range of drying times, the
capillary pressure inside the section is such that the moisture content in latewood is
higher that the moisture content in earlywood. Indeed, during this period of drying, a
macroscopic moisture content gradient is evident (from the inner part of the section
towards the exchange surfaces) and a “microscopic” moisture gradient due to the
local heterogeneities of the porous medium.

The region of liquid migration diminishes as the drying progresses and finally at 30-
40 hours, the drying process is nearing completion once the temperature attains the
dry bulb temperature and the moisture content reaches the equilibrium moisture
content. Note finally that although there is a considerable impact on the moisture
evolution due to the local heterogeneities in the virtual board, this impact is almost
unnoticeable in the temperature fields. For this reason, they are not included here.

The internal gaseous pressure whilst remaining effectively close to the atmospheric
value throughout drying is impacted by the heterogeneity of the virtual board.
Specifically, one notes the distinct influence of the growth ring locations where
undulations in the pressure distribution are evident. As water is removed from the
board initially an initial under-pressure during the constant rate period appears due
to the extension of the gaseous phase, which evolves to a slight over-pressure as the
temperature in the board slowly increases throughout the falling rate period.
Coupled heat and mass transfer 235

15 hours

0
0.01
2
0.02
Moisture

1
0.03
dth
0
0 0.04 Wi
0.005
T h 0.01 0.05
ick 0.015
ne
ss 0.02
0.06

20 hours

0
0.01
2
0.02
Moisture

1
0.03
dth
0
0 0.04 Wi
0.005
T h 0.01 0.05
ick 0.015
ne
ss 0.02
0.06

25 hours

0
0.01
2
0.02
Moisture

1
0.03
dth
0
0 0.04 Wi
0.005
T h 0.01 0.05
ick 0.015
ne
ss 0.02
0.06

Figure 10.22. Evolution of the wood moisture content throughout drying (15, 20
and 25 hours of drying)
236 Fundamentals of wood drying

10.4.2 Case Study 2 – Dielectric Drying

It is now well recognised that dielectrically assisted drying processes can provide
numerous advantages over classical convective drying techniques, including
volumetric heating, moisture levelling, increased liquid and vapour migration from
the interior to the surface of the product, and in some cases, superior end product
quality.

In this case study a completely coupled drying model (see Perré and Turner 1997,
2000 for the complete details), which combines Maxwell’s electromagnetic
equations with the comprehensive heat and mass transfer model, is used to
investigate some important aspects of microwave enhanced convective drying of
softwood in an over-sized waveguide. The solution of the Maxwell equations is
performed in the time domain using the Finite Volume Time Domain (FV-TD)
technique (Zhao and Turner 1996, 2000, Zhao et al. 1998, ) and the resolution of the
heat and mass transport model is carried out again using TransPore.

The oversized waveguide system exhibited in figure 10.23(a) operates by


propagating travelling waves along the z-direction of the waveguide towards a water
load that is situated at the end of the guide. The water load ensures that only a
minimal amount of energy is reflected back to the sample, while an iso-circulator,
which is located at the front end of the guide, is used to trap any energy reflected
from the sample damaging the magnetron. The sample of wood is inserted in the
central over-sized section of the guide. As the microwave heating proceeds, the
wood absorbs a certain amount of energy from the electromagnetic fields evolving
within the waveguide.

At first a three-dimensional study was performed on this configuration using the


mesh depicted in figure 10.23(b) in order to ascertain some fundamental
understanding of the electromagnetic phenomena evolving in the guide when a
block of lossy material with constant dielectric properties representative of wood
was located within it. The computed electric field intensity in the xz- and yz-planes
is shown in figures 10.23(c) and (d) respectively. It is clear that the dominate TE10
input microwave field changes to a multimodal nature inside the oversized section of
the guide. Most of the energy is either absorbed, or reflected, by the wood and a
standing wave pattern is established around the leading edge of the sample. It can
also be observed that very little energy is transmitted to the water load and one can
conclude that this particular design is effective in delivering the majority of energy
directly to the wood sample for the purposes of heating.

A comparison now is made with the experimental data set depicted in figure 10.10
for the microwave enhanced convective drying of spruce heartwood (refer to figure
1
10.24). The volumetric power density term is given by Φ = σ e E 2 , where E is
2
the electric field and σ e the effective conductivity. The coupling between the heat
Coupled heat and mass transfer 237

and mass transfer mechanisms and the electromagnetic fields arises due to the
variation of the relative dielectric constant ε ′( X , T ) and the relative dielectric loss
factor ε ′′( X , T ) with both moisture content and temperature throughout drying. The
correlations for ε ′ and ε ′′ were taken directly from the literature (Chen. and Pei,
1987). Because of the increase in the dielectric loss factor with increasing moisture
content, wet zones within the sample of wood are heated faster than dry zones,
inducing a moisture levelling phenomenon. However, because the loss factor also
increases with temperature at low moisture contents, a thermal runaway effect can
arise at locations where the standing wave configuration established within the
material maximises the absorbed power. This effect is one that is highly undesirable
and can be detrimental to the final product quality.

Figure 10.23. (a) Schematic of the Over-sized Waveguide used for this analysis,
(b) Computational mesh, (c) Electric field intensity in xz-plane and (d) yz-plan.
Pictures courtesy of Dr. Huawei Zhao
238 Fundamentals of wood drying

It is apparent from figure 10.24 that the match between the experimental and
theoretical results is qualitatively consistent, with the simulations exhibiting the
same overall characteristic shape as the experimental profiles. Although there are
some differences between theory and experiment, the coupled model captures reality
quite well and further observation of the kinetic curves indicates that the overall
drying times are approximately equivalent: the magnitude of the temperatures and
pressures predicted within the sample are all close to the experimentally determined
values. Further details of the experimental results and numerical simulations can be
found in Perré and Turner (1997, 2000).
1 .1 8 c m E n d p ie c e
y 1 7
2 .7 5 c m
C V lo c a tio n s o f th e
2 4 5 6 p lo tte d v a lu e s
4 .0 5 c m
z
3

Figure 10.24. Simulation results for combined microwave and convective drying of
spruce heartwood

10.4.3 Case Study 3 – Vacuum drying with radiative heating

In this case study the vacuum drying of a board placed in a pressure tank between
two IR emitters with reflectors as shown in figure 10.25 is presented. Far IR
radiation (temperatures close to 600 K) was obtained by supplying classical quartz
tubes with low voltage. This low temperature of radiation was necessary to have
good control over the process by ensuring a significant reduction in supplied power
when the surface temperature increased.
Coupled heat and mass transfer 239

D ry a ir o u tle t
vapour
F lo w o f c o ld w a te r c o n d e n sa tio n
P e rs o n a l
c o m p u te r
R e c o v e re d w a te r
D a ta
B a la n c e T h e rm o c o u p le s re c o rd in g

V acuum pum p
w ith p re s su re re g u la to r
P re ss u re
ta n k

P re ss u re
gauges

In fra re d e m itte rs

S a m p le

Figure 10.25. Schematic of vacuum drying experiment

The experimental results exhibited in figure 10.9 have been performed at a


surrounding pressure of 20 kPa (≈ 0.2 Patm). This level reduces the boiling point of
water to approximately 60°C. The sapwood part of the Fir tree has a high value of
initial moisture content (160%). The sample dimensions were 1.3 cm × 22.5 cm
× 15 cm. The volume of the chamber was 0.1 m3 and the pump flow rate was
3 m3·h-1. The overall drying kinetics computed by the computational model are
exhibited in figure 10.26. Clearly the trends exhibited by the model are consistent
with those observed experimentally, with the drying time, overpressure and
temperature evolution consistent with the experimental results.
240 Fundamentals of wood drying

180 1.0

160 0.9
MC
End piece
140 0.8
Temperature (°C) or MC (%)

120 0.7

Pressure (P/Patm)
T surface

100 0.6
T 4 mm T centre

80 0.5

60 0.4
P centre
P 4mm

40 0.3

20 0.2

0 0.1
0 100 200 300 400 500

Time (minutes)

Figure 10.26. Simulation results obtained with the comprehensive model


TransPore: sapwood part of Fir. Averaged MC, temperature and total pressure at
different positions plotted versus time

10.5 Conclusion

An exposition of the coupled heat and mass transfer mechanisms evident during the
drying of wood has been presented in this chapter. The physics of drying has been
carefully reviewed, specifically in the context of a complex hygroscopic biological
porous medium such as wood, and the knowledge gained from this study has been
used to elucidate the effectiveness of some important industrial drying
configurations, which includes both low temperature convective drying and drying
with internal vaporisation. The latter case is of considerable importance to the wood
drying industry because of its relevance to a variety of state-of-the-art wood drying
operations, such as high temperature convective drying with superheated steam,
vacuum drying and dielectric drying.

We have also surveyed and assessed the main drying models reported in the
literature over the last twenty years for simulating wood drying processes and have
discussed the motivation for the choice of these formulations and highlighted the
limitations of some of these models for simulating high temperature configurations
that involve internal vaporisation. We have used this study to propose a simple
guideline for the choice of formulation and geometrical description that we hope
will be useful in assisting drying practitioners with their model decision making in
the future when simulating the drying of wood.
Coupled heat and mass transfer 241

Finally, we have presented three wood drying case studies that exemplify the
importance and insight that can be gained from computational modelling with
understanding the fundamental physical processes involved in wood drying. We
believe that it is only with the combination of advanced computational modelling
and sophisticated experimental work that we are now very close to bridging the once
wide gap that existed between theoretical and experimental observation. The future
prospects of this combined approach are indeed exciting, with three-dimensional
simulations of virtual boards now close to realisation. This methodology will
undoubtedly enable further optimisation of existing technologies, and hopefully lead
to the discovery of novel and innovative drying strategies that can be applied not
only to wood, but many other products. In this direction, the development of
multiscale models is very promising (Perré 2007).
11 Stress development
Patrick Perré and Joëlle Passard
LERMAB
Laboratoire d´Etude et de Recherche sur le Matériau Bois
UMR1093 INRA, ENGREF, UHP,
ENGREF 14, rue Girardet 54 042 Nancy Cedex
France
perre@nancy-engref.inra.fr
Joelle.Passard@iutnb.uhp-nancy.fr

ABSTRACT
Stress development during wood drying is the main reason for mechanical degrade. In this
chapter, the mechanical aspects of wood drying are briefly presented and used to explain the
trends in stress development during the process, including case hardening.

Then, a relevant mathematical formulation of these mechanisms is proposed and used in a


computational model. Several simulation examples are depicted, which clearly corroborate
the rules enacted in chapter 1 to obtain a fast and good drying process.

KEYWORDS: shrinkage, stress, strain, viscoelastic, mechanosorptive, case hardening,


thermo-activation, hygro-activation, computational simulation.

Fundamentals of wood drying, Patrick Perré editor, pages 243 to 271.


244 Fundamentals of wood drying

11.1 Introduction

Industrial wood drying consists of not only removing moisture from green-wood,
but also ensuring that its quality (to be defined for its purpose) is adequate. As a
result of wood undergoing shrinkage during drying, deformations and stresses
develop that can lead to unusable end products. The understanding of these aspects
must account for the complex mechanical behaviour of wood, including its memory
effect.

This chapter is devoted to understanding stress development during drying and


includes a brief description of the mechanical phenomena responsible for drying
stress. This is just an illustrated summary of the detailed information given in
chapters 5 and 8. From the basic concepts developed in those chapters, the time
evolution of the stress field is analysed across the board section throughout the
drying process, including casehardening. Keeping in mind that it is not possible to
measure stress directly, but just to observe and evaluate the effects of the stress
level, some experimental configurations have been selected for which analysis is
possible.

Finally, using a comprehensive formulation and an associated computational code


known as TransPore, several simulations are proposed to confirm the effect of
drying conditions on the stress development.

11.2 A brief overview of mechanical aspects

Green
FSP Ovendried Length
wood

0 FSP Moisture content

Figure 11.1. Wood shrinkage. The fibre saturation point (FSP), often close to 30%
in moisture content, depends on species and temperature
Shrinkage is the “driving” force for drying stress: without shrinkage, no drying
stresses would develop. Figure 11.1 exhibits the dimensional variation of an
unloaded sample with moisture content (the latter is assumed to be uniform). Under
normal conditions, the dimensions do not change until the moisture content attains a
value close to the acknowledged fibre-saturation point (see Chapter 4). At this point,
the dimension variations are almost proportional to the change of moisture-content.
Stress development 245

This strain field is called free shrinkage/swelling (sometimes known as stress-free


hygro-expansion)

Constant Time t=0 + Time t After cycles of


moisture moisture content
content

LOAD LOAD
LOAD

Figure 11.2. Viscoelastic and mechanosorptive behaviour of wood: a typical


experimental protocol

Moisture
content

Time
Load

Time
mechanosorptive
Length
variation viscoelastic
elastic

Time

Figure 11.3. Viscoelastic and mechanosorptive behaviour of wood: schematic shape


of the sample’s response to the previous protocol
246 Fundamentals of wood drying

A sample submitted to tensile or compressive stress exhibits at first an instantaneous


deformation (elastic part), which then increases in time (viscoelastic creep). After
cycling the moisture content to and from a higher level, the creep has been
significantly greater due the so-called mechanosorptive action (Fig. 11.2 and 11.3).

Thus, a sample subjected to a compressive stress during drying (case 1, in


Figure. 11.4) exhibits a smaller length at the end of the process than an unloaded
specimen (case 2), which itself has a smaller length than the sample subjected to a
tensile stress (case 3). In this experiment, the viscoelastic behaviour results because
of time, and the mechanosorptive behaviour results because of the removal of water
molecules due to drying.
+ Tim e t
Tim e t=0
High m oisture content Low m oisture content

LO AD

1 2 3 Drying 1 2 3

LO AD

Figure 11.4. Effect of the load level on the dimension changes of a specimen during
drying (after Perré 1996)

11.3 Stress development in a board during drying

The ideas presented previously can now be applied to the drying of lumber boards,
assumed to be initially stress-free. At the beginning of drying (constant drying-rate
period), sap throughout the entire board remains free. No shrinkage occurs, hence
stress build up is absent.

At the beginning of the second drying period, shrinkage exists close to the exposed
surfaces (Fig. 11.5a). At this moment, if the section were cut into slices, the outer
slices would have a shorter length than the inner ones (Fig. 11.5b). This
displacement field is not compatible and induces, in the actual section, a tensile
stress in the surface layers and (because of equilibrium conditions) a counteracting
Stress development 247

compressive stress in the core layers (Fig. 11.5c). During this period, surface
checking is possible. From this point onwards, the wood layers dry under load.

As the drying proceeds, thermally activated viscoelastic creep develops, together


with mechano-sorptive creep. The outer slices appear similar in configuration to that
exhibited for slice #3 in Figure 11.4, while the internal slices resemble slice #1.
Consequently, in spite of the flat moisture-content profile, slicing the section at the
end of the drying would give picture b in Fig. 11.6: the core slices, dried under
compression, are smaller than the outer ones, dried under tension. In the actual
section, tensile stress exists in the inner part (Fig. 11.6c).

Second drying period


a b c

tensile
stress
FSP
compressive
stress

Figure 11.5. Appearance of drying stresses following upon shrinkage

End of drying
a b c

tensile
stress
FSP
compressive
Xeq stress

Figure 11.6. Stress reversal due to the memory effect of wood

This phenomenon is known as stress reversal or casehardening. The residual stress


level depends on many parameters (growth history, sawing pattern, drying
conditions, species, thickness…), which provide most of the problems for drying
optimisation. In addition, one must keep in mind that the gradients of moisture
248 Fundamentals of wood drying

content, strain and stress exist along the thickness. This explains the curvature of the
slices observed in a prong test or during the cup method commonly used in industry
to assess stress levels (Fig. 11.7). An interesting simulation of this test can be found
in Dahlblom et al. (1996).

Second drying period End of drying

Prong
test

Cup
method

Figure 11.7. Two experimental methods used to assess drying stresses

11.4 Experimental evidence of drying stress

Shrinkage induced drying stress has been introduced as one of the main reasons for
the poor quality of dried products. Recall that stress cannot be observed, nor
measured. In fact all what can be assessed is the effect of the stress fields, namely
the strain fields due to stress relaxation. This fact is well-known in the field of
continuum mechanics, but is worth a reminder. For example, a load cell just
indicates the deformation, using strain gauges, of a calibrated structure.

Drying stress follows this rule and different experimental methods allow the stress
field to be converted into a deformation field:
a) In the case of a non-uniform shrinkage field in a piece of wood (reaction wood,
fibre angle, material anisotropy), the induced stress field can be almost entirely
relaxed as a result of the global deformation (Fig. 11.8). This kind of
deformation due to the properties of the raw material (see Chapter 5) is usually
very difficult to avoid.
Stress development 249

b) When the stress level exceeds the modulus of rupture, checks appear in the
section as surface checking during the second drying period and internal
checking at the end of the process (Fig. 11.9). This kind of default is usually
strongly tied to the drying schedule or to problems in kiln regulation. The main
concern of wood drying is to avoid these defaults by suitable, improved or
innovative procedures.
c) Finally, because stress cannot be observed, all techniques imagined in the
industry or in laboratories to assess the stress level consists in relaxing one part
of the stress field by splitting the piece into smaller parts (Fig. 11.7).

Figure 11.8. An example of board deformation after drying: this default is very
easy to assess (Courtesy of Osmar Aguiar, EMBRAPA, Brazil)

Figure 11.9. Example of severe internal checking (Oak dried at 120°C). This
default, which results from a too high level of stress reversal, can be observed only
once the board is cut
250 Fundamentals of wood drying

11.5 Observation on micro-samples

11.5.1 Shrinkage: the driving force for stress

Laboratory experiments carried out on small samples are very informative on the
mechanisms involved in mechanical degrade during drying, namely the interaction
between the minute structure of wood and its behaviour.

A new apparatus was designed and built to determine the drying behaviour of small
samples (Perré 2003). The unit incorporates a highly sensitive electronic
microbalance, a high-speed laser scan micrometer and an infrared pyrometer for
non-contact determination of width and surface temperature. The micro-samples
(typically 20 x 10 x 4 mm3 in L, T and R directions respectively) are carefully
prepared with a diamond wire sawn. For such small samples, the stress level is not
high enough to produce checks: the sample shrinkage measured by the micrometer is
just an average value of the shrinkage through the thickness.

The first test (Fig. 11.10) depicts the results obtained for a sample of beech,
sapwood dried in the radial direction. The second plot corresponds to a sample of
spruce, sapwood, dried in the same direction (Fig. 11.11). The drying conditions
used for these tests are gentle: 30°C for the dry bulb temperature and 25°C for the
wet bulb temperature.

30 9

Dry bulb temperature


Surface temperature
Tangential shrinkage (%)
Temperature (°C)

28 6
Shrinkage

26 3
Wet bulb temperature

24 0
0 5 10 15 20
Time (hours)

Figure 11.10. Drying test on micro-samples: surface temperature and sample width
versus time (beech, transfer in the radial direction)

In both tests, the surface temperature progressively increases from the wet bulb
temperature towards the dry bulb temperature. However, the shape of the curve
strongly differs between the two species. In the case of beech, the existence of a real
constant drying rate period becomes evident through the beautiful plateau observed
Stress development 251

at the wet bulb temperature during more than three hours. The absence of shrinkage
confirms the presence of liquid water on the exchange surface during this period. In
the case of spruce, the temperature slowly increases from the commencement of the
drying: no real plateau can be observed. Consequently, this first phase of drying is
not a constant drying rate period. This trend is confirmed by the shrinkage
behaviour: contrary to the test on beech, this spruce sample starts to shrink from the
first instants of drying. This observation means that the stress field due to drying,
especially tension stress close to the exchange surfaces, does exist from the
commencement of drying.

30 6

Dry bulb temperature

Tangential shrinkage (%)


Surface temperature
Temperature (°C)

28 4
Shrinkage

26 2
Wet bulb temperature

24 0
0 5 10 15 20
Time (hours)

Figure 11.11. Drying test on micro-samples: surface temperature and sample width
versus time (Spruce, transfer in the radial direction)

11.5.2 Checking: the end result of stress

Simple observation of the exchange surface during drying is very informative. For
this purpose, the experimental set-up uses an optical microscope with a digital
camera (Perré 2003). The control of the drying conditions is very simple: the device
is just at room temperature, an electric fan ensures an airflow at about 2 m/s around
the sample and a halogen lamp heats and lights the sample surface. Therefore, the air
temperature is low (≈ 25°C), however due to the heating flux, the drying conditions
are rather severe. The micro-samples (typically 5 x 20 x 10 mm3 in L, R and T
directions respectively) are carefully prepared with a diamond wire sawn. The
exchange surface is prepared with a sledge microtome to ensure a good quality of
the image. A thin thermocouple is inserted into the sample close to the exchange
surface and all faces but one are insulated to guarantee one-directional transfers.
Two typical tests are depicted here: the first one was on a cross-section of Douglas
fir and the second one on a cross-section of beech, both coming from the sapwood
part of freshly cut trees.
252 Fundamentals of wood drying

In the case of softwood, the fist image, taken at 40 minutes, shows a large part of the
earlywood region with empty tracheids. At this instant, the surface temperature is
still close to the wet bulb temperature. When the drying progresses, the dry zone
continues to develop near the surface. The temperature increases and the dry zone
becomes thick enough for checks to develop. Note that checks are initiated in the
latewood zone, where the shrinkage has the highest values (Fig. 11.12, 1.5 hour) but
they finally spread in the earlywood zone (Fig. 11.12, 3 hours). At the end of drying,
these checks partly disappear because of stress reversal. Nevertheless, due to the
much higher shrinkage value of latewood compared to earlywood, small checks
persist in latewood even at the end of drying (Fig. 11.12, 10 hours).

40
Temperature (°C)

35

30

25
0 2 4 6 8 10
Time (hours)

Figure 11.12. Microscopic observation of the exchange surface during drying


(Douglas fir)
Stress development 253

In the case of beech, the vessels, with internal diameters of about 60 µm have much
lower capillary pressure than in the surrounding fibres: vessels empty at first, before
the end of the first drying period. When the surface leaves the wet bulb temperature,
shrinkage develops at the surface and initiates checks in small rays (never in the
large ones). Within earlywood they sometimes pass through vessels: in this case, the
checks have a jagged shape (although beech is a diffuse porous species, its density is
smaller in earlywood with larger and more abundant vessels). In spite of this slight
difference, this shrinkage is homogeneous enough for the checks to disappear
completely at the end of the test (Fig. 11.13, 8 hours).

28
Temperature (°C)

26

24

22

20
0 2 4 6 8
Time (hours)

Figure 11.13. – Microscopic observation of the exchange surface during drying


(Beech)
254 Fundamentals of wood drying

11.6 The “Flying-wood” test

Surface checking can be observed, and even measured (Canteri et al. 1997).
Eventually residual stresses, including casehardening, can be assessed at the end of
the process using prong or cup tests. Nevertheless, as already stated, it is very
difficult to gather information on the stress field within boards during drying.

No shrinkage Shrinkage throughout the whole


Shrinkage and
No stress thickness
tensile stress
Compressive stress near the surface,
near the surface tensile in the inner part
MC

EMC

Falling drying rate period

Domain of free w ater


Hygroscopic range
Typical board
Tension

Compression

One exchange surface

Curvature

Time

Figure 11.14. Principle of the “Flying wood” test

The “Flying Wood” test has been conceived and elaborated in our laboratory at
ENGREF/Nancy (Brandao and Perré 1996, Aguiar and Perré 2000a). The
observation, in laboratory experiments, of the shape obtained after contact drying is
at the origin of this work. It contributes to the need for an easy and fast test to
characterise the drying behaviour of wood (Terazawa 1965, Brandão and Jankowsky
Stress development 255

1990). Note that some similar works can be found in the literature (Nisho 1976,
Lamvik 1982, Wang and Choong 1994).

The test consists in non-symmetrical drying of small boards (0.5 x 10 x 20 cm3).


Five faces of the sample are coated, so that the moisture migration occurs only in
one direction. This strategy transforms a part of the drying stresses encountered
during conventional drying into a global deformation of the sample (Fig. 11.14). The
latter can subsequently be measured using any non-destructive method (digital
image or displacement sensor with light contact).

Note the three successive stages: the first drying period with liquid water at the
surface with no shrinkage and hence no deformation; the second period with
shrinkage at the surface and positive curvature; and the final stage, which starts
when shrinkage occurs close to the impervious face. Due to the memory effect of
wood, which is responsible for stress reversal in normal boards, the curvature
becomes negative in the case of non-symmetrical drying. Even when the external
conditions are the same, the response of each specimen to this test strongly depends
on the species, the sawing pattern and the position in the tree (sapwood/heartwood).

The sapwood part of beech exhibits a very long constant drying period (from 80%
down to 40% of MC) during which the drying rate is almost constant and the section
remains flat because no shrinkage occurs at the surface (Fig. 11.15). Then, because
beech has a large value of tangential shrinkage, the curvature increases rapidly to a
large maximum curvature. Moisture transfer in beech in tangential direction is not
easy in the hygroscopic range: a proof of this lies in the dramatic decrease in drying
rate after the first drying period. Consequently, the shrinkage and stress histories
differ significantly between the surface and impervious faces. This explains the large
absolute value of the final, negative, curvature.

1.2 0.15
Drying rate
Curvature 0.1
1.0

0.05
Curvature (cm-1)
Drying rate (g/mn)

0.8
0

0.6 - 0.05

- 0.1
0.4 Fitted curve
- 0.15
0.2
- 0.2

0 - 0.25
0 10 20 30 40 50 60 70 80

Moisture content (%)


Figure 11.15. Drying rate curve and curvature obtained for the “Flying wood” test
on a specimen of beech (sapwood – flatsawn)
256 Fundamentals of wood drying

Compared to beech and in agreement with Figure 11.11, the constant drying rate
period almost disappears for the sapwood specimen of Scots pine (Fig. 11.16).
Consequently, a positive curvature develops very soon after the commencement of
drying. The curvature increases slowly up to its maximum. In addition, the drying of
pine is quite easy (high value of the mass diffusion coefficient). This explains why
the final section is almost flat, which reveals a low level of memory effect.

1.2 0.15

Curvature 0.1
1.0
0.05

Curvature (cm- 1)
Drying rate (g/mn)

0.8
0

0.6 Fitted curve - 0.05

- 0.1
0.4 Drying rate
- 0.15

0.2
- 0.2

0 - 0.25
0 10 20 30 40 50 60 70 80

Moisture content (%)

Figure 11.16. Drying rate curve and curvature obtained for the “Flying wood” test
on a specimen of Scots pine (sapwood – flatsawn)

1.2 0.15

Curvature
0.1
1.0

0.05
Drying rate (g/mn)

Curvature (cm- 1)

0.8
0

0.6 Fitted curve - 0.05

Drying rate - 0.1


0.4

- 0.15
0.2
- 0.2

0 - 0.25
0 10 20 30 40 50 60 70 80

Moisture content (%)

Figure 11.17. Drying rate curve and curvature obtained for the “Flying wood” test
on a specimen of oak (heartwood – flatsawn)
Stress development 257

The third example is a specimen of oak, flatsawn in the heartwood part (Fig. 11.17).
Oak is known to be one of the most difficult species regarding drying. This flying
wood test confirms this status: no first drying period, quick decrease of the drying
rate, high maximum curvature and very large absolute value of the final curvature.

The final shapes obtained, for the same drying conditions, of samples taken from
different species and different parts of the log exhibit the great variability of the
wood material (Fig. 11.18). Compared to flatsawn boards, the drying rate is smaller
for quartersawn samples, due to the low moisture migration in the tangential
direction. Their curvature is also smaller, because the radial shrinkage, rather than
tangential, is involved in this case.

Sapwood

Heartwood – rift sawn

Sapwood
Sapwood

Heartwood – plain sawn


Heartwood – rift sawn

Heartwood – quarter sawn Heartwood – quarter sawn

Figure 11.18. Final section shapes obtained for Scots pine (above), oak (left) and
beech (right) as a function of the sawing pattern (same drying conditions : dry air at
95°C)

11.7 Drying stress formulation

During drying, shrinkage appears in all parts of the board for which the moisture
content, X, is within the hygroscopic range. The shrinkage strain is proportional to
the difference between the local moisture content and the local value of the moisture
content at fibre saturation at the same temperature. A deformation field tensor
sh
denoted ε is defined in the material’s axes by Equation (11.1).
258 Fundamentals of wood drying

A 0 0
sh  
ε = H ( x)  0 B 0 
 0 0 C
 
 0 if X ( x ) ≥ X fsp
with H ( x ) =  (11.1)
 X ( x ) − X fsp if X ( x ) ≤ X fsp
A, B and C are real positive numbers
If this deformation field does not fulfil the geometrical compatibility, a strain tensor
mec
ε related to stress is generated. The constitutive equation, which represents the
mec
mechanical behaviour of the material, relates this strain tensor ε and the stress
mec
tensor. Due to the memory effect of wood, this tensor ε has to be divided into
elas
two parts: an elastic strain ε , connected to the actual stress tensor and a memory
mem
strain ε which includes all the strain due to the history of that point.
mem
ε enables the plasticity, viscoelasticity and mechano-sorption effects to be
captured within the model.

mec elas mem


ε =ε +ε (11.2)
tot
The geometrical compatibility applies to the total strain field ε . When solving the
mechanical problem in terms of displacement, the total strain tensor is deduced from
the displacement field and this geometrical condition is automatically fulfilled
within the domain. The stress field must satisfy the local mechanical equilibrium
and the boundary conditions. Finally, the complete formulation of the stress problem
is given by the following set of equations.

 tot 1 T
ε = 2 (∇u + ∇ u) over Ω

σ + ρ f = 0 over Ω

σ = a (ε tot − ε 0 ) over Ω 0 sh mem (11.3)
 ε =ε +ε
( )
 σ.n = T on Γ
 i
i Ti
 ∀i, Γ Di ⊕ Γ Ti = Γ
ui = Di = 0 on Γ Di

This static formulation requires that boundary and volumetric forces satisfy the
global equilibrium. Γ is the surface surrounding the domain Ω. Γ Di refers to the
sub-domain of Γ where the ith component of the displacement is known and Γ Ti
refers to the sub-domain of Γ where the ith component of the traction force is
known. In order to ensure the uniqueness of the solution, additional conditions are
Stress development 259

required on the boundary: ∀i, mes (Γ Di ) > 0 . Otherwise, the solution is defined
within a rigid body motion. Finally, wood being orthotropic, each behaviour law
involves nine independent terms. In fact, it is more common to define the inverse of
ai j k l which, for the case of linear elasticity, leads to the so-called generalised
Hooke’s law:

 1 ν RL ν TL 
 E − − 0 0 0 
 L ER ET 
 ν 1 ν TR 
 elas
ε LL   − LR − 0 0 0 
   EL ER ET   σ LL 
   σ 
elas
ε RR
   − ν LT −
ν RT 1
0 0 0   RR 
   EL   σ TT 
elas
εTT ER ET (11.4)
 elas = σ 
 2ε LR = γ LR   0 0   LR 
1
0 0 0
 elas   GLR   σ LT 
 2ε LT = γ LT     σ 
 elas   0 0 0 0
1
0   RT 
 2ε RT = γ RT   GLT 
 1 
 0 0 0 0 0
 GRT 

Memory effect
mem
In describing the strain field ε lies the entire problem of developing a
constitutive model for wood, which requires both theoretical and numerical work.
Comprehensive formulations are also very difficult to characterise (Ranta-Maunus
1975). The problem lies in the fact that the memory effect of wood depends not only
on the temperature and moisture content values, but also on their variations in time
and on the history of their variations in time. This subject remains a matter of some
scientific debate (Keey et al. 2000).

Nevertheless, in the case of drying, the moisture content only decreases and some
simplifications apply. Here, only the most common way to express viscoelastic and
mechanosorptive creep will be presented. The general formulation of the time
dependency of the viscoelastic property involves the whole stress history:
t

∫ J (t − t ') dt ' dt '
visc
ε = (11.5)
−∞

where J (t ) is the creep compliance tensor and t the actual time. The experimental
creep function is often analysed as the response of Kelvin elements in series (Fig.
11.19), each having the property of a spring and dashpot in parallel (Martensson
1992, Mohager and Toratti 1993, Dahlblom et al. 1996, Hanhijärvi 1999). In the
case of uniaxial load, this leads to:
260 Fundamentals of wood drying

 N
  −t   
J (t ) = J 0 1 +∑ an 1 − exp    
 n =1   τ n   
(11.6)

Element 1 Element 2 Element N


2
Element 0

Figure 11.19. Modelling the viscoelastic behaviour of wood with the number of
Kelvin elements in series

The temperature and moisture dependency of that function can be expressed using a
material time, or by changing the characteristic time n. The thermal activation, for
example, is often expressed with the aid of an Arrhenius law:

∆Wn
τ n = τ n∞ exp(+ ) (11.7)
RT
where Wn is the activation energy associated to element n.

In order for the computational results to capture reality and be predictive, the thermo
and hygro-activated viscoelastic behaviour of wood have to be characterised very
carefully. In particular, it has to remain valid over a suitable range of temperature
and moisture content values. In the following simulation session, four thermo-
activated Kelvin’s elements are used. Their values have been fitted from different
creep tests performed on Spruce at an increasing temperature followed by a plateau
at 65°C, 85°C, 105°C and 120°C respectively (Passard and Perré 2001, Perré and
Passard 2004). In addition, the effect of moisture content on the viscoelastic
activation is very important to understand and to model the drying of wood. A
corrective factor has been adapted from literature results (Irvine 1984) in order to
account for the effect of moisture content, which becomes significant below 18% in
moisture content. The Cole-Cole plot built from moist wood (MC above FSP) and
dry wood (MC equal to 10%) summarises this thermo and hygro-activation (Fig.
11.20).
Stress development 261

200
Moist wood (MC > FSP)

Thermal softnening
150
Loss modulus (MPa)

60°C
100
40°C
80°C
50

0
0 100 200 300 400

Storage modulus (MPa)

200
Dry wood (MC = 8 %)

150
Loss modulus (MPa)

100
60°C

40°C
50
80°C

0
0 100 200 300 400

Storage modulus (MPa)

Figure 11.20. The Cole-Cole plot obtained with the Kelvin’s elements for moist and
dry wood at different temperature levels (40°C, 60°C and 80°C). The frequency
ranges from 3·10-5 to 1.5·10-3 Hz. This range corresponds to a time period stretching
from 10 minutes to 10 hours, which is representative of the drying process (after
Perré and Passard 2004)

The mechanosorptive effect occurs as soon as the moisture content changes during
load. The simpler way to express this effect consists in assuming that the strain rate
tensor depends linearly on both the stress field tensor and the time derivative of the
moisture content:

ms
dε dX
=mσ (11.8)
dt dt
The mechanosorptive strain rate is always in the direction of the stress field, hence
the absolute value of the time derivative of the moisture content (11.8). One should
be aware that the simple formulation of the mechanosorptive creep as described in
262 Fundamentals of wood drying

Equation (11.8) ignores effects pointed out in some works, namely the
unrecoverable part of the mechanosorptive creep and a possible interaction between
viscoelastic and mechanosorptive effects (Hanhijärvi and Hunt 1998, Swensson and
Toratti 2002, Navi et al. 2002).

A tri-dimensional resolution is very costly in terms of calculation time and computer


memory space. The need for such a cost is justified whenever the objective of the
calculation lies in evaluating the overall deformation of the board. In this way, the
effect of reaction wood, fibre angle, and property variations can be analysed. Good
examples of these possibilities can be found in the literature (Dahlblom et al. 1996,
Ormarsson et al. 2003). Nevertheless, to study the stress development within a
section far from the ends of the board, a two-dimensional simulation is sufficient. A
“planar displacement” formulation has to be used in this case, assuming small
displacements (Perré and Passard 1995, Chen et al. 1997) or large displacements
(Mauget and Perré 1999).

11.8 Some simulation examples

All examples presented in this session, except the non-symmetric case, refer to a
flatsawn board of heartwood, 20 mm thick and 40 mm wide. They have been
computed using the numerical code TransPore (Perré and Turner 1996, 1999,
Mauget and Perré 1999, Perré and Passard 2002). Figure 11.21 depicts the moisture
and stress fields calculated at different drying stages for quite severe conditions at a
medium temperature level (Td = 80°C, Tw = 60°C). After 6 hours of drying, the
external part of the section is within the hygroscopic range, which gives rise to a
tensile stress due to shrinkage in the zones close to the exchange surfaces. As a
consequence of the mechanical equilibrium, internal zones undergo a compressive
stress. A negative shear stress exists close to the edge (the right angles have
decreased). At the end of the drying process (60 hours) the moisture content is
almost equal to the equilibrium moisture content (7%) throughout the section. The
stress reversal due to the memory effect of wood is clearly exhibited. Note that the
shear stress also changes in sign.

When a face of the board is insulated, the drying conditions are no longer
symmetrical and significant section deformations can be observed experimentally
(Brandão and Perré 1996). Figure 11.22 shows an example of non-symmetrical
drying. In order to increase the section deformation, a thin quartersawn board has
been simulated, here 5 mm by 80 mm. The large displacement formulation is
essential in this case. In this configuration, one part of the drying stress is
transformed into section deformation, hence the stress-reversal phenomenon induces
a negative final curvature.
Stress development 263

Exchange faces

Planes of symmetry

Moisture Content Sigma xx


a) 1.0 1.0

0.8 0.65 0.8 1.54


0.58 1.24
Thickness (cm)

Thickness (cm)
0.51 0.93
0.6 0.44 0.6 0.63
0.37 0.32
0.29 0.01
0.4 0.4
0.22 -0.29
0.15 -0.60
0.2 0.2

0.0 0.0
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
Width (cm) Width (cm)

Sigma xy Sigma yy
1.0 1.0

0.8 0.15 0.8 1.29


0.09 1.04
Thickness (cm)

Thickness (cm)

0.03 0.79
0.6 -0.03 0.6 0.53
-0.09 0.28
-0.15 0.03
0.4 0.4
-0.21 -0.22
-0.28 -0.47
0.2 0.2

0.0 0.0
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
Width (cm) Width (cm)

Moisture Content Sigma xx


b) 1.0 1.0

0.8 0.08 0.8 0.12


0.08 -0.29
Thickness (cm)

Thickness (cm)

0.07 -0.70
0.6 0.07 0.6 -1.11
0.07 -1.52
0.07 -1.93
0.4 0.4
0.07 -2.34
0.07 -2.75
0.2 0.2

0.0 0.0
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
Width (cm) Width (cm)

Sigma xy Sigma yy
1.0 1.0

0.8 0.24 0.8 -0.10


0.21 -0.35
Thickness (cm)

Thickness (cm)

0.18 -0.60
0.6 0.15 0.6 -0.86
0.12 -1.11
0.09 -1.36
0.4 0.4
0.06 -1.62
0.03 -1.87
0.2 0.2

0.0 0.0
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
Width (cm) Width (cm)

Figure 11.21. Moisture content and stress fields after a) 6 hours and b) 60 hours
(Td = 80°C, Tw = 60°C)
264 Fundamentals of wood drying

7 hours

Moisture: 0.05 0.09 0.14 0.18 0.22 0.26 0.31 0.35

30 hours

Figure 11.22. Example of non-symmetrical drying: moisture content and section


deformation (80°C/60°C)

As a result of the large range of validity of the material parameters input to the
computer code, the model has good predictive capabilities. In particular, the model
is able to illustrate the three rules stated in chapter 1 to obtain a fast and good drying
process (Table 11.1)

Table 11.1. Recipes to obtain both a fast drying operation as well as a good
product quality
Parameters Value Reasons why Problems
Unique way to obtain a low Could be difficult to ensure
Drying potential MC gradient. Imposes a high for certain kilns. Additional
Low
(dry bulb – wet bulb) EMC (low shrinkage, high problems: discoloration,
creep) mould growth
Easy internal transport and Possible increase of
Temperature High
high viscoelastic effect collapse
Good homogeneity of MC Increased electricity
Air velocity High
throughout the stack consumption

The next example illustrates these rules using three different constant drying
conditions (Td stands for dry bulb temperature and Tw for wet bulb temperature):
a) Td = 40°C, Tw = 35°C; mild conditions at low temperature (the equilibrium
moisture-content, EMC, is equal to 15%);
b) Td = 80°C, Tw = 60°C; rather severe conditions at medium temperature (EMC =
7 %);
c) Td = 80°C, Tw = 76°C; mild conditions at medium temperature (EMC = 14%).
Stress development 265

Figure 11.23 depicts the variation of the average moisture content and stress levels
(direction parallel to the exchange surface) at different positions versus time. All
tests show a first drying period, without drying stress, then a stage with tensile stress
in the peripheral zones and finally the last drying stage, which exhibits the stress
reversal. However, the duration of each stage and the stress level depend strongly on
the drying conditions.

For the mild drying conditions at low temperature (Td = 40°C, Tw = 35°C), the
drying time is rather important. The first drying period lasts for around 10 hours and
the stress level is high, both for the second drying stage and the final drying stage.
The drying conditions are mild concerning heat and mass transfer, while the
temperature level is not high enough for the creep field to relax the stress field: the
final stress level reveals the importance of the memory effect.

As a consequence of the low relative humidity of the air, the second test (Td = 80°C,
Tw = 60°C) is very fast. However, both the maximum tensile stress level and the
final stress reversal are important: the rapid external transfer imposes a high
moisture-content gradient within the board. In addition, the viscoelastic creep is not
sufficient to cancel the memory effect. At the beginning of drying, the board
temperature is close to the wet-bulb temperature, which is below the glass-transition
zone. At the end of drying, the temperature level is sufficient for greenwood, but not
for the dry part of the board, to active the viscoelastic behaviour. Consequently, the
outer parts, which are close to the equilibrium moisture content, are below the
softening zone.

In the third test (Td = 80°C, Tw = 76°C), the difference of moisture content between
the surface and core remains low. The first drying period lasts for a majority of the
total drying time. Due to the high value of the equilibrium moisture content, the
board temperature is always above the softening zone: consequently all stress levels
remain very low. These conditions allow wood of good quality to be obtained
relatively free of stress reversal with a moderate drying time (less than 150 hours
against 400 hours for the low temperature test).

To obtain a better understanding of the effect of drying conditions on the stress


evolution, Figure 11.24 depicts the moisture content profiles. On each sub-graph,
profiles 2 to 6 are plotted for an average moisture content equal to 70%, 55%, 40%,
30% and 20% respectively. During the initial transient period, some condensation
might appear which depends on the gap between the initial sample temperature and
the wet bulb temperature of the surrounding air. This is why condensation does not
exist for the first run and is so important for the last one (see the first profile of each
sub-graph). These profiles also show the difference in drying time and in MC
gradient as a function of the external conditions. The gradient within the
hygroscopic range has to be particularly noted. The change in board thickness can be
observed on the last profiles, especially when the equilibrium moisture content is
low (150 hours in Figure 11.24b).
266 Fundamentals of wood drying

90 1.5
a) MC
Surface
75 5 mm 1.0
10 mm
Center
Averaged MC (%)

60 0.5

σ (MPa)
45 0

xx
30 -0.5

15 -1.0

0 -1.5
0 100 200 300 400

Time (hours)

90 1.5
b) MC
Surface
75 5 mm 1.0
10 mm
Center
Averaged MC (%)

60 0.5

σ (MPa)
45 0

xx
30 -0.5

15 -1.0

0 -1.5
0 50 100 150

Time (hours)

90 1.5
c) MC
Surface
75 5 mm 1.0
10 mm
Center
Averaged MC (%)

60 0.5
σ (MPa)

45 0
xx

30 -0.5

15 -1.0

0 -1.5
0 50 100 150 200

Time (hours)

Figure 11.23. Averaged MC and σxx versus time: a) Td = 40°C, Tw = 35°C, b) Td =


80°C, Tw = 60°C and c) Td = 80°C, Tw = 76°C
Stress development 267

a) 1.0
Moisture content (kg/kg)
0.8 1H

8H
38 H
0.6
100 H

162 H
0.4

242 H
0.2
400 H

0
0 0.5 1.0 1.5

Position (cm)

b) 1.0
45'
Moisture content (kg/kg)

0.8
12 H 3H
30 H
0.6
48 H

0.4
68 H

0.2
150 H

0
0 0.5 1.0 1.5

Position (cm)

c)
1.0
2H
Moisture content (kg/kg)

0.8
13 H
30 H
0.6
56 H
82 H
0.4

110 H
0.2
200 H

0
0 0.5 1.0 1.5

Position (cm)

Figure 11.24. MC profiles at different elasped times: a) Td = 40°C, Tw = 35°C,


b) Td = 80°C, Tw = 60°C and c) Td = 80°C, Tw = 76°C
268 Fundamentals of wood drying

These simulations are in good agreement with non-symmetrical drying experiments


performed on Oak (Quercus rubra) boards using the same drying conditions
(Fig. 11.25). However, a closely similar schedule for another hardwood (Nothafagus
truncata), resulted in gross deformations and thermal degradation due to the high
extractives content of the wood (Grace 1996).

Figure 11.25. Final section curvature of oak samples dried on their upper face with
conditions n°2 (Td = 80°C,Tw = 60°C) and n°3 (Td =80°C, Tw = 76°C) respectively
(after Perré 2001b)

Based on this reasoning, new drying procedures have been devised and tested on
different tropical species, including numerous tests in an industrial kiln having 100
m3 capacity (Aguiar and Perré 2000b). The product quality was always very good,
often with very little checking and rather less deformation than with conventional
drying. Most importantly, this method needs only one-half to one-third of the time
required for drying according to conventional schedules.

The code can also be used to test different drying schedules (Figs 11.26 and 11.27).
The first one (Schedule A) is recommended for softwoods while the second one
(Schedule B) is recommended for hardwoods. Schedule B proposes lower
temperature levels and higher relative humidity values. In this drying schedule, the
equilibrium moisture-content (EMC) decreases significantly only at the end of the
process, when the board is supposed to be dry with a low moisture-content gradient.
As a first consequence of these drying conditions, one observes a much longer
drying time for schedule B (130 hours against 50 hours). The first drying period
lasts also for a longer time for the second procedure (30 hours instead of 10). Note
however, that the first drying-period duration represents about the same percentage
Stress development 269

of the total drying time for both schedules. This remark still stands for the stress
level. One can easily consider a relative time (current time over total drying time) at
which all curves have the same shape and the same stress magnitude.

Schedule A
Average MC Dry bulb T Wet bulb T RH (%)
(%) (°C) (°C)
Green 71 66 80
50 76.5 68.5 70
30 82 70.5 60
20 88 67.5 40

100 70

80 60
Moisture Content (%)

Temperature (°C)
Temperature Edge
60 Temperature Center 50
MC Edge
MC Center
MC Surface
40 MC Average 40

20 30

0 20
0 50 100 150

Time (hours)
2

Surface
5 mm
10 mm
1 Center
σ xx (MPa)

-1

-2
0 25 50 75 100 125 150

Time (hours)
Figure 11.26. Simulation of drying for schedule A: moisture content, temperature
and stress level at different positions versus time
270 Fundamentals of wood drying

Schedule B
Average MC Dry bulb T Wet bulb T RH (%)
(%) (°C) (°C)
Green 40.5 38 85
60 40.5 37 80
40 43.5 39 75
35 43.5 38 70
30 46 39.5 65
25 51.5 43 60
20 60 47.5 50
15 65.5 49 40

120 90

100 80
Moisture Content (%)

Temperature Edge

Temperature (°C)
80 Temperature Center 70
MC Edge
MC Center
MC Surface
60 MC Average 60

40 50

20 40

0 30
0 20 40 60

Time (hours)
2
Surface
5 mm
10 mm
1 Center
σ xx (MPa)

-1

-2
0 20 40 60

Time (hours)
Figure 11.27. Simulation of drying for schedule B: moisture content, temperature
and stress level at different positions versus time
Stress development 271

In the first example, the advantages to be gained from using high relative humidity
levels (low moisture-content gradient, high hygro-activation of the viscoelastic
behaviour) hardly offsets the negative effect of the low temperature levels (slow
moisture migration and low thermo-activation of the viscoelastic behaviour). A
careful analysis of these approaches is very promising. New rules can be derived to
improve existing drying schedules or to devise innovative drying procedures.

Finally, the reader must be aware that this field of drying procedure improvement,
which also recognises the advancements made with computational models, is a very
active field (Salin 1999, Carlsson and Arfvidsson 1999). Additional information
could be found in Keey et al. (2000).

11.9 Conclusion

Stress development during wood drying is the main reason for mechanical degrade.
In this chapter, the mechanical aspects of wood drying were briefly presented and
used to explain the trends in stress development during the process, including case
hardening.

The experimental evidence of the effect of drying stress (deformation, checking)


reported throughout this chapter provided a fundamental understanding of this
phenomenon, both at the anatomical and at the macroscopic levels.

A mathematical formulation was derived to predict these mechanisms, which was


then implemented as a computational model. Several simulation examples are
presented to clearly corroborate the rules enacted in chapter 1 to obtain a fast and
good drying process.

11.10 Acknowledgements

The authors wish to acknowledge the contribution of Prof. Ian Turner for his careful
review of this chapter.
12 Discoloration of wood during drying
Gerald Koch1 and Michalis Skarvelis2
1
Institute of Wood Biology and Wood Protection
Federal Research Centre of Forestry and Forest Products
Leuschnerstr. 91, D-21031 Hamburg
Germany
gkoch@holz.uni-hamburg.de
2
N.AG.RE.F./Forest Research Institute of Athens
Terma Alkmanos, 115 28 Athens
Greece
Skmi@fria.gr

ABSTRACT

The various kinds of discolorations in freshly felled and kiln-dried wood described in the
literature are reviewed and classified according to the different physiological, biochemical,
and chemical reactions involved. In freshly felled and stored round wood discolorations are
initiated predominantly through physiological reactions of living parenchyma cells. Typical
reactions are the formation of phenolic compounds and tyloses triggered by oxygen
penetrating the tissues. Discolorations can also be caused by microorganisms, for instance
blue stain fungi, mould fungi, and bacteria which affect the wood surface of inadequately
stored and kiln-dried wood. The discolorations during kiln-drying and steaming are based
mainly on chemical reactions of the accessory compounds and cell wall components in the
woody tissue. The reaction mechanisms are dependent on the process parameters and
chemical composition of the wood which have to be determined in individual cases. Measures
to be taken for facing these phenomena are also reviewed, depending upon the cause and
wood species.

KEYWORDS Discoloration, extractives, kiln-drying, reaction mechanisms, remedies,


steaming, sticker staining

Fundamentals of wood drying, Patrick Perré editor, pages 273 to 289.


274 Fundamentals of wood drying

Part I: Description of different types and causes of discolorations by G. Koch

12.1 Introduction

The discoloration of stored round wood after felling and in sawn timber during
steaming and drying is a considerable economic problem due to the increased
demand of light and uniformly coloured wood in Europe. In order to provide
practical procedures and techniques for its prevention, it is necessary to identify
these discolorations as either physiological processes already initiated in the living
or freshly felled tree or as secondary chemical reactions during steaming or drying.
Based on the literature (e.g. Zimmermann 1974, Ward and Pong 1980, Bauch 1984,
1986, Chang et al. 1999) discolorations related to wood during drying can be
classified into five groups (Table 12.1).

Table 12.1 Classification of discolorations on the basis of the different types of


reactions (according to Bauch 1986)
Cause Example and Description
(1) Fungi Blue-stain as in sapwood of Pinus due to the chromophoric
pigments in the hyphae
(2) Physiological reaction in Formation of tyloses and accessory compounds as in Fagus
living cells due to the reaction of living parenchyma cells
(3) Biochemical reactions Orange discoloration as in Alnus due to the enzymatic
reaction of polyphenoloxidases inducing the production of
oregonin
(4) Chemical reactions Metal-tannin reactions as in Quercus

(5) Combination of Yellow discoloration of Quercus and Castanea due to a


reactions fungal infection and reaction of hydrolysable tannins

In this study, various kinds of discolorations in freshly felled and kiln-dried wood
described in the literature are reviewed and additional information on some
individual case studies are presented. On the basis of this review, it may be possible
to better understand basic patterns of discolorations and the different reaction
mechanisms involved.

12.2 Discolorations in freshly felled and stored round wood before kiln-drying

In freshly felled and stored round wood discolorations are initiated predominantly
through physiological reactions of living parenchyma cells for at least for several
weeks after felling. Typical reactions are the formation of phenolic compounds and
tyloses, e.g. in beechwood and maple (Figure 12.1), triggered by oxygen penetrating
the tissues (e.g. Dietrichs 1964, Ziegler 1968, Murmanis 1975, Magel and Höll
1993).

Furthermore, biochemical reactions are also responsible for the discoloration of


round wood during storage. For instance the brown stain of lumber of pine species
Discoloration of wood during drying 275

(Oldham and Wilcox 1981) and Douglas fir (Miller et al. 1983) is a serious problem
that begins during storage in moist conditions. The stain is a result of an enzymatic
reaction involving peroxidases on tannins and phlobaphenes. The orange
discoloration of Alnus is an additional example for the enzymatic reaction of
polyphenoloxidases inducing the production of oregonin (Hrutfiord and Luthi 1981).
The intensity of the stain during storage of round wood is dependent on the moisture
content of the wood, the length of time the wood is exposed to air, and the
surrounding temperature (Millett 1952, Zycha 1952, Liese 1958).

Discolorations can also be caused by microorganisms, for instance blue stain fungi,
mould fungi, and bacteria which infect the wood surface of inadequately stored and
kiln-dried wood (Figure 12.2). While blue stain is based on the chromophoric
pigments in the hyphae (e.g. v. Aufseß 1986), mould fungi and bacteria can
decompose the soluble carbohydrates and storage substances in the tissue. The
staining is attributed to the metabolism of the microorganisms accompanied by a
biochemical reaction of the extractives in the tissue (e.g. Bauch et al. 1985, Yazaki
et al. 1985). Furthermore, bacterial wetwood, e.g., in sugar pine and fir, can
predispose the wood to oxidative brown stain during drying. This reaction depends
on the composition of the microbial population (Ward and Pong 1980).

The described discolorations during storage can be prevented, if the changes in


moisture content and temperature that favour growth of microorganisms and the
initiation of physiological and biochemical reactions are restricted. Quick
harvesting, transport, and processing of round wood, especially during the growing
season, is an important contribution to the conservation of wood quality (e.g. Koch
et al. 2000). The influence of the cutting season has been discussed, but not well
examined yet. According to Asikainen 2002, the discoloration of birch wood (Betula
pendula) is mainly influenced by the felling season. On the other hand, no
significant differences were found in colour between winter and spring processed
scotch-pine (Pinus sylvestris) timber (Terziev and Boutelje 1998).

Figure 12.1 Staining of inadequately stored maple round wood. The discolorations
are caused by oxidative reactions of extractives (!) in the wood rays
276 Fundamentals of wood drying

12.3 Discolorations during kiln-drying

The discolorations during kiln-drying are based essentially on chemical reactions of


the accessory compounds and cell wall components (lignin and hemicelluloses) in
the woody tissue (Fig. 12.2). It has been suggested that the main factors involved in
chemical reactions are oxidation and condensation of phenolic compounds (Wengert
1990, Taylor 1997), although the reasons in many cases are unknown. For example,
the phenols, including condensed tannins, are complex compounds, which can be
converted in unsoluble reddish compounds by oxidation and polymerisation in hot
and acidic conditions (Hillis 1985). The discolorations can be also caused by
hydrolysis of hemicelluloses. The hemicelluloses are degraded gradually to
monosaccharides which then react with nitrogen compounds in the wood.

These chemical reactions occur at temperatures above 40°C and a moisture content
of 30% to 60%. A decrease in pH value during drying intensifies the hydrolysis (Ifju
1973). The reaction mechanisms are dependent on the process parameters and the
chemical composition of wood (Koch and Bauch 2000).

Figure 12.2 Discolorations of kiln-dried (left part figure) and fungal infected (right
part figure) beechwood. The colour alterations in the kiln-dried wood are caused by
condensed reactions of phenolic compounds (!) in the parenchyma cells

12.3.1 Mechanisms of colour reactions during kiln-drying of hardwoods

One of the most important problems in conventional drying of hardwoods are


discolorations of oak, which cause major financial losses for the industrial dryers.
According to Wegener and Fengel (1987) the discolorations can be classified into
two groups: brown and yellow.

The brown discoloration is induced by hydrolysis and oxidative transformation of


ellagtannins. Particularly, vescalagin and castalagin are involved in the hydrolytic
Discoloration of wood during drying 277

mechanisms influenced by temperature and oxygen during conventional kiln-drying


(Charrier et al. 1995). Wassipaul and Fellner (1987), Fortuin et al. (1988a, 1988b),
and v. Hundt (1990) have characterised the specific conditions favouring such
discolorations, e.g., wood moisture content between 30 and 60% combined with a
kiln-temperature above 30°C and a relative humidity at about 70%. The same
authors showed that the brown discoloration penetrates from the surface to the
center of the wood (Figure 12.3). Stich and Eberman (1984) found enzymatic
activities by peroxidases and polyphenoloxidases in oak heartwood, which could be
also responsible for the discoloration in moist conditions.

The discoloration is probably independent of wood origin and can be prevented by


quick kiln-drying without oxygen (Charrier et al. 1992). According to Wassipaul
and Fellner (1992) and Fortuin et al. (1988a) air drying or low temperature
predrying from the green to approximately 25 percent moisture content before kiln-
drying are the most efficient ways of protecting the wood from discolorations. Other
techniques like drying in a nitrogen atmosphere would be a good preventive method
(Wassipaul and Fellner 1992), but such laboratory scale systems preventing
oxidative reactions have not yet found an industrial application. In recent years, kiln-
drying under vacuum and super heated steam vacuum drying have been proven as
effective methods preventing the brown discoloration (Welling and Wöstheinrich
1995, Brunner 1999). These methods combine two advantages: quick drying and a
low oxygen concentration inside the kiln (Welling and Wöstheinrich 1995).

Figure 12.3 Brown staining of kiln-dried oak. The discolorations are dependent on
increasing kiln- temperature and relative humidity

The yellow discoloration of oak, which frequently occurred as sticker stain or as


longitudinal yellow streaks, were identified to be caused by an infection with the
mould fungus Paecilomyces variotii (Bauch et al. 1991). The fungus colonises the
vessels of the wood reacting with the hydrolysable tannins (lactone derivatives)
278 Fundamentals of wood drying

likely to be responsible for the yellow discoloration. These specific heartwood com-
pounds were found in Quercus species of the Section Robur (white oak) and also in
Castanea sativa, but not in Quercus species of the Section Rubrae (red oak). Given
the thermophilic adaptation of the fungus –tolerating an acidic medium of pH-3 as
well as a temperature of up to 50°C (Husain and Zamir 1968) – the yellow discol-
orations can arise during the first stage of kiln drying.

The prevention of mould growth can be achieved by good ventilation of sheltered


stacks, or by faster drying schedules. In situations where these prophylaxes are not
practicable, a pre-treatment of green lumber with 5% to 10% propionic acid can
prevent mould growth during the critical phase of drying (Bauch et al. 1991).

Another severe problem during kiln-drying are yellow or grey discolorations of light
coloured timbers like birch or maple, which normally occur in the inner parts of the
boards, while the outer layer to a depth of 1-5 mm from the surface remains
unaffected (Trübswetter 1995, Luostarinen and Luostarinen 2001). In this case,
visual sorting by colour is impossible before the boards are processed. The reason of
the discoloration is described as oxidative reaction of water soluble low molecular
compounds including carbohydrates, amino acids, and phenolics. The intensity of
the discoloration is mainly a function of the drying temperature, more severely
under so warmer conditions. The relative humidity during the drying process has
also been found to affect the final colour of wood. In sugar maple, which behaves
similar to birch during kiln-drying, the final colour is lighter if the relative humidity
is lower (Mc Millen 1976). Contrary to e.g. oak, a low relative humidity could be
used for drying of birch and maple timber because they do not check easily.

To reduce the discoloration, it has been recommended that timber be dried rapidly
(low relative air humidity, high air velocity) at a low temperature (Taylor 1997).
This recommendation is based on the experimental comparison of drying schedules.
It has also been suggested that other factors may affect discolorations in wood, e.g.,
felling date, length and season of storage period, and climatic conditions before
felling (Mc Millen 1976, Kreber and Byrne 1994, Luostarinen and Luostarinen
2001, Luostarinen et al. 2002).

12.3.2 Mechanisms of colour reactions during kiln-drying of softwoods

The development of brown discolorations during kiln-drying of many softwoods –


commonly called kiln brown stain – was described in literature as a serious problem
(e.g. Kreber et al. 1999). Especially in radiata pine sapwood brown stain develops
during kiln-drying and occurs just underneath the wood surface. This type of stain
differs from the brown surface stain that can develop during storage of unseasoned
radiata pine logs and lumber, or the brown stain reported for sugar pine (Pinus
lambertiana) which can extend throughout the total thickness of the board, and may
be due to oxidation of extractives.
Discoloration of wood during drying 279

Mc Donald et al. (1997) and Kreber et al. (1998) showed that water-soluble
constituents were likely to be responsible for the occurrence of kiln brown stain. The
authors reported that the predominant chemical compounds found among the water
soluble compounds included carbohydrates, amino acids, and phenolics. These
compounds, present in the sap of woody tissue, migrate and accumulate close to the
surface, and react (auto-hydrolysis and depolymerisation) during the drying process
(Kreber et al. 1998). In addition, solubilized lignin derivatives may also be a
contributing factor in the development of kiln brown stain.

The intensity of the stain was shown to be dependent on rising kiln temperatures
(McDonald et al. 2000). In order to prevent the discoloration, different kiln
schedules with lower temperatures were tested. Although the levels of staining
observed with kilning at 45°C/35°C (dry-bulb/wet-bulb) may be acceptable from an
industrial perspective, drying times of at least four times what is currently achieved
in commercial 90°C/60°C drying are unlikely to be acceptable from a commercial
point of view (Kreber et al. 1999). Furthermore, the effect of stepped kiln schedules
was assessed. These schedules utilised low-temperature regimes (45°C/35°C) from
green to 60 percent moisture content. The assumption behind this approach was that
pit aspiration and reduced mass flow of moisture (including kiln brown stain
precursors) would reduce the stain intensity. Unsatisfactory results of this drying
schedule indicate that the development of kiln brown stain is a continuous and
perhaps cumulative process, mainly influenced by the kiln temperature in the final
stage of drying. Compared to standard convection drying using 52°C/42°C,
70°C/60°C, and 90°C/60°C kiln schedules, vacuum drying at the same kiln
temperatures consistently reduced the frequency and intensity of kiln brown stain,
but not to satisfactory levels (Kreber et al. 1999).

Kreber and Byrne (1994) reviewed the mechanisms of hemlock brown stain and
reported several types of compounds which may induce staining, namely flavonoids
(e.g. catechin), phenolics, and also anorganic compounds like manganese and copper
which may catalyse stain formation. The special reaction mechanisms were already
described by Barton and Gardner (1966) who emphasised that hemlock sapwood
brown stain is an innocuous tannin-like surface stain and is in no way connected
with decay, sapstain, or mould. It is postulated that catechin is transported with
moisture to the surface of the lumber during drying and, accumulating there with
evaporation of the water, reacts with air in the presence of an enzyme to form a
brown polymerised pigment (Kreber and Byrne 1994). Also, since the stain can be
removed by planing provided the lumber is dry, it is not detrimental to the end
utilisation of the species except where light colour on rough lumber is a necessity
(Barton and Gardner 1966).

In Scots pine where a yellow surface discoloration can be observed after kiln-drying,
an accumulation of nitrogen containing compounds was recorded at the drying front
(Theander et al. 1993). The authors assume that the yellow stain is caused by a reac-
tion of low molecular weight sugars.
280 Fundamentals of wood drying

12.4 Discolorations during pre-steaming and steaming

Hydrothermal treatment of wood in form of pre-steaming, steaming or boiling in


water is commonly used in the wood processing industries. It is applied in the
production of veneer, and also for the conditioning of lumber before or after kiln-
drying (Kubinsky and Ifju 1973). In Europe beech is the most important lumber,
which is frequently subjected to steaming immediately after sawing to achieve
desirable colour (Koch et al. 2002).

The colour modifications during the pre-steaming and steaming process are based
mainly on chemical reactions of the cell wall components (lignin and
hemicelluloses) and extractives in the wood (e.g. Sarni et al. 1990, Burtin et al.
2000). The separation of the acetone and methanol extracts of discolored tissue from
pre-steamed and steamed beechwood fractions shows different polymerised
proanthocyanidins such as catechin and 2,6 dimethoxy-benzochinon (Koch et al.
2003). During the steaming process these low molecular phenols are converted into
high condensed compounds with reduced chromatographic mobility and an
increased molecular weight. Steaming may also initiate a cleavage of lignin-
polysaccharide complexes by organic acids released from hemicelluloses (Puls
1993). On the other hand, it is assumed that some interactions of wood cell wall
components lead to the formation of a secondary lignin-carbohydrate linkage which,
eventually, causes discolorations (Košíková et al. 1999).

According to Kollmann et al. (1951) conjugated (chromophoric) double bonds are


formed in the lignin molecule under steaming conditions. In addition, it is assumed
that some interactions of wood cell wall components lead to the formation of
secondary lignin-carbohydrate linkages (Košíková et al. 1999). The steam treatment
also mobilises water-soluble compounds (e.g. soluble carbohydrates and starch) or
hydrolysis products which migrate from the vacuoles of parenchyma cells into the
lumina of vessels and fibers and can initiate discolorations through oxidation and
condensation (Plath and Plath 1955, Korte et al. 1991, Choong et al. 1999, Yilgor et
al. 2001). The intensity of the colour changes is significantly influenced by the time
of exposure, temperature, and wood moisture content (Brauner and Conway 1964,
Chen and Workman 1980). For instance, in black walnut (Juglans nigra), sapwood
darkening was accelerated remarkably with increased steaming time from 0 to 32h
and temperature from 100 to 120°C (Brauner and Conway 1964). Furthermore, the
authors showed that green black walnut wood exhibited greater and more rapid
darkening than air dried wood. No significant differences occurred in the degree of
darkening between saturated samples of black walnut treated at 75°C and 100°C.
However, a treatment at 125°C exhibited a significant degree of darkening (Brauner
and Loos 1968). Steaming at 100°C for 16h was found to be an efficient treatment in
order to bring sapwood colour closer to the natural colour of heartwood in a walnut
hybrid (Juglans nigra x Juglans regia Burtin et al. 2000). According to the authors,
hydrojuglone glucoside, gallic and ellagic acid derivatives can be regarded as major
precursors of artificial wood colouring, providing chromophores through a
degradation process.
Discoloration of wood during drying 281

In order to provide a better understanding of discoloration during kiln-drying and


steaming detailed studies on the physiological, biochemical, and chemical reactions
of the extractives and wall components have to be carried out in individual cases. As
soon as the reaction mechanisms and chromophoric compounds are identified,
information for establishing comprehensive methods for preventing this
economically important loss of wood quality can be provided.

Acknowledgements

The author thanks Prof. Dr. J. Bauch, Prof. Dr. J. Ressel, Dr. H.G. Richter, and Dr.
J. Welling, BFH Hamburg for critical reading the manuscript.
282 Fundamentals of wood drying

Part II: Description of remedies preventing discoloration by M. Skarvelis

12.5 Remedies

In order to face the problem of undesired discolorations during wood processing,


there are some suggestions sited below, in different phases of wood conversion.

12.5.1 Before sawing

Under favourable weather conditions, fungal discoloration can occur within hours
after cutting the tree. When temperatures are below 10°C, no serious problems are
reported, so in northern areas attention should be given only in summertime. In all
other areas, especially in the south, almost all over the year this factor has to be
faced. Fungi attack can be achieved by eliminating one of the four prerequisite
elements they need for growing: food, water, warm temperature, and oxygen
(Tsoumis 1983). In warm weather only oxygen could be eliminated, usually done by
spraying water or dipping the logs in water tanks. Practice has shown that logs can
be stored under water from 6 to 18 months, without substantial damage
(Cassens 1991).

Chemical fungicides have been also used to prevent log staining, making another
factor (food) unsuitable for fungi. Pentachlorophenol dissolved in oil or its salt and
sodium pentachlorophenate (PCP) mixed with borax have proven to be effective, but
their use has now been restricted due to environmental and hygienic reasons. Less
harmful formulations are in use now. If treating logs, it should take place within 24
hours of cutting or less during summer (Amburgey et al. 2000). Not all wood species
do respond to the treatment. Good behaviour was shown for oak (Quercus sp.), elm
(Ulmus sp.), maple (Acer sp.), yellow-poplar (Liriodendron tulipifera), sweetgum
(Liquidambar styraciflua), robinia (Robinia pseudoacacia) magnolia (Magnolia sp.)
hickory (Carya sp.), sycamore (Platanus occidentalis), white and yellow birch
(Betula sp.). Black oak (Quercus velutina), beech (Fagus sp.) and some pine species
did not respond (Cassens 1991).

It is not correct that only the end surfaces of the log need treatment. The bark offers
a protection, but openings in the bark due to partial debarking or cut branches permit
the infection of the log. In applying chemicals, all prophylactic care should be given
to avoid contact with the skin or inhaling chemicals.

Discolorations during storage can be prevented, if changes in moisture content and


temperature that favour growth of microorganisms and the initiation of physiological
reactions are restricted. Otherwise these reactions may affect the colour and this can
be obvious immediately or much later, when lumber is dried. Sticker stain, interior
greying, brown stain, interior pinking or browning of light-coloured species can be
the result of these reactions. So, quick harvesting, transport, and processing of round
wood, especially during the growing season, is an important contribution to the
Discoloration of wood during drying 283

conservation of wood quality (e.g. Koch et al. 2000). In any case, logs should not
stay for a long period in the yard. First arrivals should be sawn first. This prevents
logs staying in the yard more than required, infections and reactions are avoided
and, the costs of inventory are limited. A common practice also is to saw first
species with great proportion of sapwood.

12.5.2 After sawing

As it happens with round wood, lumber is affected by the same factors and
discoloration may develop under appropriate conditions. Especially under warm and
humid conditions (very common in the south and during first steps of kiln-drying)
the circumstances are quite suitable for all the above described kinds of
discoloration.

As a general rule to face all kind of reasons, lumber has to be stacked as soon as
possible. Especially under warm weather it has to be done within 12 hours after
sawing and the stacks have to be driven soon to kiln dryers. If this is not the case,
stacks should be stored under shelters with adequate natural or forced air circulation.
Special care must be given to species prone to checking.

a) Fungal discoloration

A lot of fungus species cause discoloration to sapwood, known as sapstain. Most


common is blue stain in softwoods (blue, grey or black colour). Blue stain does not
necessarily mean that wood is also decayed or has lost its mechanical properties, but
in any case lumber prices are falling dramatically. Another type of fungi that
discolours wood but usually does not penetrate as deep as sapstain is mould
(Cassens 1991). However, if sapstain is appeared, conditions are favourable for
mould or decay to develop later.

Most fungi growth stops at temperatures below 5°C and above 40°C and moisture
content lower than 20%. It doesn’t mean that lumber can not be further infected if
rewetted.

Dip or spray treatments with anti-sapstain formulations are used to avoid lumber
discoloration. In warm and moist environments lumber has to be chemically treated
with fungicides within 36 hours after sawing and should be stacked immediately
after that. With classical methods chemical can not penetrate lumber more than 2 or
3 mm and it is possible for fungi to develop in lumber’s interior. So, when kiln-
dried, it is important to set the initial dry bulb temperatures above 55°C, as soon as
possible, in order to avoid internal staining of lumber. Sapstain affects during drying
the surfaces of untreated boards, when drying at low temperatures and moisture
content is more than 20%. This has occurred in dehumidification drying and early
stages of solar drying (Skarvelis 1996).
284 Fundamentals of wood drying

Only one species of fungus (Paecilomyces variotii) is found to attack heartwood of


species such as oak. The prevention of mould growth can be achieved by good
ventilation of sheltered stacks, or by faster drying schedules. In situations where
these prophylaxes are not practicable, a pre-treatment of green lumber with 5% to
10% propionic acid can prevent mould growth during the critical phase of drying
(Bauch et al. 1991).

b) Bacterial discoloration

Bacterial activities in freshly cut lumber and yeasts change the pH value of wood
into alkaline values (> 7.5) and trigger the reactions of some accessory compounds
during kiln drying (due to higher temperatures) (Bauch et al. 1985). This can be
faced by spraying green lumber with aqueous solutions of weak organic acids, such
as propionic acid (a substance used also for conserving human food), which
neutralizes alkaline environment (Simpson 1991).

c) Chemical discoloration

As it is mentioned above chemical discolorations occur due to oxidative and


enzymatic reactions with chemical constituents of the wood. Species without
obvious heartwood (such as beech, ash), are the most susceptible to stain.
Hardwoods are also more prone to chemical discoloration than softwoods. In some
hardwood species (e.g. alder, dogwood) intense discoloration could appear within an
hour after exposing freshly cut surfaces to the air (Simpson 1991). In other species
(e.g. walnut and teak) a discolored core may appear after drying, but this is
temporary and will disappear with light and time (Wengert 1992). Stain could be
eliminated in the surface and easily removed by planing, but it also could go deeply.
Chemical stains can be removed by bleaching them with reagents, while fungal
stains can not. A simple application of oxalic acid will indicate the presence of a
chemical stain (Bois 1970).

The degree of staining depends mainly upon the chemical constituents of sapwood
and the drying temperature, especially above fiber saturation point. In oak drying
brown or brown-greying discoloration occurs during air drying above 25°C and
moisture content between 30% – 60% and also during kiln drying at mild or severe
temperatures (Resch et al. 2000). Above 60°C the whole mass of sapwood may turn
in brown, yellowish or pinkish in species as beech, birch, alder, maple, hickory and
ash. For producing light finishes schedules which start at temperatures less than
43°C with a wet bulb depression of 6°C are recommended. Drying temperatures
should be kept below 55°C until average moisture content reaches 15%
(Simpson 1991).

The opposite could also appear: deep greyish-brown discolorations in sapwood after
long exposing period in air-drying yards or pre-dryers (low temperatures). It is
mentioned as an important problem to a lot of species, mainly oak, ash, maple, birch
and Douglas-fir. To prevent this, green lumber should be stacked with stickers as
Discoloration of wood during drying 285

soon possible after sawing and ensure good air circulation. Drying temperatures
should be above 20°C in the beginning. Heating or steaming green lumber at 100°C
has been tried to inactivate the enzymes that activate the discoloration chain, but
with limited success (Simpson 1991).

It is not only chemical constituents and drying temperatures that affect the colour of
lumber, but also relative humidity does. Drying at relative humidity above 60% was
found to promote kiln brown stain in radiata pine (Kreber and Haslett 1997). Also
brown stain development was found in kiln drying of western hemlock at 80% RH,
instead of drying at 40% RH (Avramidis et al. 1993). The opposite found in radiata
pine kiln-drying, where brown stain was more intense after drying at low humidity
(Kreber and Haslett 1997).

Drying oak in a R/F-Vacuum kiln resulted in shorter drying times and more light
colour of oak and beech lumber, perhaps due to restricted levels of oxygen inside the
kiln (Charrier et al. 1992, Resch and Gautsch 2000). No extractive staining was also
experienced in R/F-Vacuum kiln drying of some softwoods (Avramidis and
Wick 1996.)

As an oxidative reaction assisted by the action of enzymes, there have been efforts
in the past to face the problem by the use of chemicals inhibiting oxidation and
enzymes (e.g. Dipping for 5 minutes in 5% sodium bisulfite (wt/wt basis) and a
compatible biocide, followed by 14 days storage for diffusion (Cassens 1991)).
Similar tests also have been made with borate solutions. It seems effective on the
surface layers but the problem is that the penetration is not deep enough to protect
the whole cross section. Other possible problems are that the sodium bisulfite may
cause corrosion to steel vats and (in oak) bisulfite reacts with iron oxide and lumber
will turn to a blackish-blue colour.

Logs or lumber fumigation under tarps with methyl bromide, prior to drying,
showed to be lethal to living parenchyma cells of some species (mainly hardwoods),
thus preventing further discoloration (Kreber et al. 1994), killing simultaneously
other harmful pests living in lumber (insects, nematodes). Method is economically
advantageous compared to some current industrial practices (e.g. pre-steaming is 35
times more costly (Schmidt 1997)).

Comparative tests for preventing brown stain in pine lumber (Pinus strobus) showed
that fumigation with methyl bromide was 45% less effective in avoiding
discoloration than dipping in a fungicide formulation based upon sodium thiosulfate
(Schmidt et al. 1995).

About 90% of the methyl bromide enters the atmosphere after treatment. It is
poisonous to living organisms and also contributes significantly to the destruction of
earth's stratospheric ozone layer. So its use has to be reduced. As substitutes other
fumigants, such as sulfuryl fluoride and methyl iodide, are reported (Schmidt 1997,
Schmidt et al. 1998).
286 Fundamentals of wood drying

d) Sticker staining

Sticker staining is one of the most common cases of chemical discoloration. The
stained area locates across the width of the lumber, exactly in the places where
stickers were located during drying. It may be darker or lighter than the rest area of
the board and often it is not visible on rough wood surfaces. It may discolour only
the surface and in this case can be easily removed by planing but, it also can develop
to a depth of several mm. Although this problem is well known since a lot of years,
major concern has been given to it only in recent years. Explanation for this might
be the use of fashion and availability driven light-coloured species and light finishes,
and elimination of bleaching in finishes (Wengert 1992).

Sticker staining is NOT a fungal stain, but if other conditions leading to fungal stain
are favourable it may be combined. Sticker stain may occur even in cases when
lumber has been dipped in a fungicide (Mc Millen 1975). Also stickers’ species,
moisture content, width and possible fungal infection do NOT cause the problem,
but may slightly worsen it. Initial slow drying of lumber (possibly less than 5% MC
loss per day) at warm temperatures (10°C to 54°C) is the basic reason for staining.
High temperatures (> 65°C) at the end steps of drying schedules contribute in
worsening the phenomenon and final discoloration may occur below 15% MC.

For preventing sticker staining the following instructions are recommended (Denig
et al. 2000). Some of them are assumed to be “not really necessary” for fresh boards
which are placed in kiln immediately after stacking and when required kiln
conditions are achieved within 6 hours, but in general they are inexpensive “just in
case” insurance:

Before sawing
• Use of fresh logs that have been stored less than 2 weeks during warm
weather
• In warm weather, stack the lumber within 12 h after sawing

Stacking and Handling


• Use stickers with 8% to 10% MC - Check MC with a moisture meter – Ob-
tain stickers from the unstacker rather than from storage
• Use stickers 3.2 cm wide, 1.9 or 2.2 cm thick
• Use grooved stickers
• Protect stacked lumber from rain, especially lumber with high MC and dur-
ing warm weather
• If stacked lumber is not loaded directly into the kiln, place it in fast-drying
locations. Fan sheds (or blow boxes) are ideal
• After lumber has dried for several days, disassemble the stack and re-stack
the lumber with dry stickers positioned (on the lumber) a short distance
(several centimetres) away from original sticker position
Discoloration of wood during drying 287

Kiln Equipment and Procedures


• Use a kiln load narrower than 5 m. For white woods, use velocity over
2.54 m/sec
• Do not use a “snow melting” or “thawing” kiln procedure
• For white woods, within the first 6 h of drying develop kiln RH levels that
are at least equivalent to 6°C depression. Greater depression may be re-
quired, if recommended in the schedule. In humid weather, kiln tempera-
tures may have to be raised slightly to achieve the low humidity required
• Load the kiln with only the two rows of lumber for the first 12 to 36 h to
help achieve the required humidity immediately. After RH has been main-
tained for ≥ 12 h, load the remaining lumber. The required RH should be
more easily achieved in this manner at all times
• Use low kiln temperatures (43°C to 49°C) initially, as recommended in the
schedule. Exceed 71°C only during equalizing and conditioning
• Reverse fans every 2 h
• Use correct kiln sampling procedures
• If lumber has been air-dried or possibly mishandled, use a special kiln
schedule that operates at very low temperatures

The use of the above described techniques (or the majority of them) is reported to
reduce the problem of sticker staining during iroko lumber kiln drying, but not quite
cure it (Zigrikas 2000). Their use is also recommended as suitable for preventing
any kind of chemical stain. But as chemical discolorations from old logs are
assumed to be almost inevitable (Denig et al. 2000), care should be given also to
other techniques that might eliminate or “hide” their presence. Segregation of
individual pieces of lumber and veneer is possible according to their colour.
Recently it is also efficient and more accurate using electronic colour measurement
(Kline et al. 1999, Resch et al. 2000).

Redistribution of the soluble sugars and nitrogen during air and kiln drying affected
significantly the colour of white painted pine boards, but weren’t visible with naked
eye. Differences could be realized only by the use of colorimeter (Terziev and
Ekstedt 1997).

e) Mineral discoloration

Mineral discoloration could appear in wood with high percentage of tannins (mainly
oak, walnut, chestnut) and needs also the presence of water and iron. So, during kiln
drying or when wood in use is wetted, mineral stains can develop in places where
nails or screws are present. During kiln drying rusty metal surfaces (roof vents, fan
floors or steam pipes) could be the iron sources. So spray lines should face upwards
and a proper water drainage system is needed, so that the water from condensation
does not drop on the wood stack. As a source of iron, metal chains or forks used in
handling the lumber are reported (Wengert 1992), but in this case mineral stain does
not penetrate lumber’s surface more 0.5 mm. Steel straps used to tie packages of
lumber maybe also the source for mineral stains.
288 Fundamentals of wood drying

Iron stain inhibitors can be added to anti-sap stain formulations, when lumber is
treated with the latter. But in any case, unnecessary contact between ferrous metal
and treating solutions should be avoided and the equipment for dipping or spraying
lumber is preferable to be constructed from stainless steel or lined with stainless
steel, fibreglass, or other non-ferrous materials (Amburgey et al. 2000). Superficial
mineral discoloration can easily be removed using bleaches.

f) Alkaline discoloration

Alkaline stains can develop when wood extractives react with calcium or potassium
hydroxides. The presence of such hydroxides could be possible in cases of concrete
or brick kiln constructions (Simpson 1991), so a remedy for this is to keep kiln
chamber in good repair.

12.6 Drying schedules for light-coloured woods

From the USDA Forest Service’s Forest Products Laboratory special schedules have
been designed for white-coloured species, such as ash, lime, hackberry and maple
and been in use with success for more than 40 years (Wengert 1992). Additionally to
this Wengert (1992) suggests the following six-step procedure to use any kiln
schedule and achieve the colour control desired:
1. Never exceed 70°C, except during equalizing and conditioning. (Dry-bulb
temperatures in the schedules are the maximum temperatures).
2. Use dry-bulb temperatures below 55°C when above 25% MC (or above
20% MC for better protection) if the lumber has been previously mis-
treated.
3. Never exceed a 25°C depression.
4. Never use less than 6°C depression, especially when first starting the
schedule. Remember, it is “what you get and not what you set”, that is im-
portant.
5. Use correct sampling.
6. Air velocity must be brisk throughout the load of lumber.

Although resin exudation during drying in softwoods, as a kind of discoloration, is a


problem that can easily be solved by planing, it may cause troubles with tools,
gluing, finishing, etc. Resin set can be achieved by using high temperature at the
start of drying, but - if brown stain is a problem - the best compromise is to use an
anti-brownstain schedule at the start of drying and finish up with 70°C or higher, for
material about 25 mm thick. For thicker stock temperatures over 75°C are required
(Simpson 1991).
Discoloration of wood during drying 289

12.7 During storage (after drying)

After sawn lumber is dried, the danger for discoloration has diminished but not
disappeared. Chemical stains are not in front stage now, but sapstain may develop if
lumber’s MC exceed 20% (e.g. in partially dried lumber or when the stacks remain
in open areas, where rewetting may occur). High ambient humidity can cause
serious MC upraising, but this only happens quite rapidly when lumber is directly
exposed to rain. So, the need of storage under shelters with enough air circulation is
imposed. Stacking without stickers after drying, not only reduces the storage volume
needed, but also prevents fast rewetting.

Care should be given also during transportation. Bluestain development is referred


during softwood transportation by the sea in closed storerooms, when in lumber a
high MC was existing (Skarvelis and Karaoulanis 2001).

Plastic or fabric warps used either during transportation or storage, should be


considered temporary, as they trap moisture and this can lead to stain
(Wengert 1998).

A long term exposure to light (especially to direct sun-light) leads in discoloration of


lumber, due to ultra-violet light reaction with lignin. It is possible during air drying
or storage in open areas and also when the stacks are placed under a shelter but near
to its end point. The exposed surfaces (not the whole stack) usually become darker
or turn to grey. In some species as iroko, this may happen in a very short time. This
is mainly a surface phenomenon, as light does not penetrate wood more than
0.200 µm (Feist and Hon 1984). It can easily be removed by planing.

12.8 Conclusions

Handling round or sawn wood with care is a prerequisite to avoid undesirable


alteration of its colour during drying. Kind of measures to be taken depends on
wood species and whether it happens due to fungal causes, physiological, chemical
or biochemical reactions. Although a lot of experience is concentrated and applied in
most of the cases, there are still serious problems.

Acknowledgements

The author thanks Dr. P.K. Kavvouras, FRI Athens for critical reading the
manuscript.
13 Airflow within kilns
Stefan F. Ledig1, Bart Paarhuis2 and Michel Riepen2
1Technische Universität Dresden
Institut für Verfahrenstechnik und Umwelttechnik
Mommsenstrasse 13
D-01069 Dresden, Germany
Stefan.Ledig@mailbox.tu-dresden.de

2TNO TPD
Models & Processes Division
P.O. Box 155
2600 AD Delft, The Netherlands
Paarhuis@tpd.tno.nl, Riepen@tpd.tno.nl

ABSTRACT

The airflow within a timber kiln enables the transfer of heat to the wet wood and the removal
of the air containing evaporated moisture. The temperature, relative humidity and velocity of
the airflow determine the external drying conditions for the wood.
In the ideal case, the same external conditions would exist at every location in the timber
stack leading to the same drying rates. However, in a real drying kiln, neither the flow
velocity nor climate conditions are uniform throughout the stack. The airflow in the kiln is
influenced by various components causing pressure losses and flow non-uniformities, while
the confined space allows just a partial redistribution before the air enters the stack.
Differently sized channels between the boards and packs create other sources of flow
maldistribution. Due to the direct impact of external conditions on first period drying rates, a
non-uniform velocity distribution can result in a considerable variation in drying rates across
the stack and intolerable timber degrade. Moreover, in the effort of reducing drying times as
much as possible, equalising periods are often shortened, which may preserve large moisture
content variations until the end of the process. For these reasons, a uniform airflow
distribution is an important prerequisite for an efficient kiln operation and high product
quality.
This chapter will therefore investigate the flow pattern inside and outside the timber stack
focusing on the causes and effects of flow maldistribution as well as suggesting ways to
secure a sufficiently uniform flow field in the kiln. In the course of the discussion, both
theoretical and experimental aspects will be examined.

KEYWORDS: airflow, fluid dynamics, pressure drop, timber kiln, velocity distribution

Fundamentals of wood drying, pages 291 to 332.


292 Fundamentals of wood drying

13.1 Introduction

Recently, research efforts have focused on various novel timber-drying methods


such as vacuum and radio-frequency/microwave drying. Even so, most timber in
industry is still dried in forced-circulation convection kilns at atmospheric pressure.
The general design for this kind of dryer is shown in Figure 13.1. Hot and humid air
sweeps through the sticker space between stacked timber boards. On its way
through, the air transfers heat to the wood surface and picks up evaporated water
while being cooled down and humidified in the process. Fans, which are usually
positioned overhead, generate the desired flow rate by providing the pressure
difference to compensate for pressure losses through the stack and kiln components.
Vents in the ceiling allow the removal of humid air and the supply of fresh dry air.
Heating coils re-establish the necessary gas temperature. Steam and/or water is
injected during specific drying stages to control the relative humidity in the kiln. In
large kilns, the airflow is reversed regularly to reduce moisture variations along the
flow path, which result from the decreasing drying capacity of the air as it is cooled
through the stack.

An efficient operation of a convection kiln aims at drying timber quickly and


uniformly while minimising the energy consumption. This is usually achieved only
to a certain extent, and in practice different drying kilns can exhibit different
efficiencies. One reason for this limitation is a possible airflow maldistribution
across the stack, which causes non-uniform drying with excessive variations of
moisture content and considerable additional energy costs due to stack bypassing
airflow. Airflow maldistribution can be the result of an inadequate kiln geometry
with wrongly sized bends and plenum chambers as well as improperly positioned
fans, baffles, and heating coils. A poor stacking strategy can also contribute to non-
uniform airflow when it brings about large gaps between packs or near the kiln wall

Overhead fans

Interior ceiling

Heating coils

Stack baffles

Filetted timber

Plenum chamber

Figure 13.1. View of a typical box-shaped timber kiln


Airflow within kilns 293

causing bypass flow. All these factors can generate non-uniform flow in all three
dimensions of the kiln.

The airflow in conventional wood dryers has been studied by several researchers, a
few of whom should be mentioned here because of their more general approach.
Other literature resources will be referred to specifically in the appropriate sections.
Early work on flow distribution in kilns was carried out in the 1930s by Greenhill
(1937) who measured velocities using a vane anemometer. At the same time
Schlueter and Fessel (1939) examined the performance of industrial kilns applying
vane and hot-wire anemometry. Later Kroell (1954) investigated the application of
turning vanes in right-angled bends to avoid flow maldistribution in drying kilns.
Werner (1972) carried out experiments on velocity and temperature distribution as
well as mixing processes in direct-heated kilns. Horton and Resch (1976) measured
air velocity profiles down the stack height for various plenum and sticker sizes. A
summary about the influence of different kiln components on the airflow was given
by Krischer and Kroell (1978) His study includes the flow around fans, effects of
baffles, flow guides, and distribution devices (screens).

To accompany the experiments, numerical methods such as CFD (computational


fluid dynamics) have been applied to analyse velocity distributions in kilns. Arnaud
et al. (1991) tested the influence of different stack and kiln geometry parameters on
the flow uniformity. Riepen and Gard (1996) simulated the effect of a baffle in the
plenum chamber. Ledig and Militzer compared measured and simulated flow fields
of kilns with end-mounted and overhead fans at atmospheric pressure and vacuum
conditions (Ledig and Militzer 1999a, b).

Nijdam and Keey developed a hydraulic kiln model to study the influence of plenum
width and different shapes of right-angled bends on the velocity distribution along
the stack height (Nijdam and Keey 1999, 2002). These data, as well as experimental
results from a semi-scale kiln, were successfully described by a one-dimensional
flow model (Ledig et al. 2001, Nijdam and Keey 2000). A round-up of some airflow
aspects in kiln drying of lumber is included in a book by Keey et al. (2000). Hua et
al. (2001) applied a kiln-flow model, which exploits the analogy of momentum and
heat/mass transfer to calculate drying rates of individual boards, to simulate the
influence of plenum and roof design.

Despite the numerous recommendations for improving the uniformity of the airflow
in kilns, there is no general conclusion for the optimum kiln design. Because of the
various geometric parameters and their possible combinations, as well as the
individual demands of the customer, a tailored solution for the specific kiln has to be
found in each case. In the following paragraphs, the reader will be provided with
information about the basic features of airflow in kilns. Other sections will present
experimental and numerical tools to investigate the flow field. The general
description will be illustrated with examples from actual wood kilns, which are also
used to deduce recommendations for kiln design and stacking strategy.
294 Fundamentals of wood drying

13.2 Observed airflow phenomena in a kiln

The box shape of a typical wood drying kiln offers a rather confined path for the
drying air, which results in various flow resistances both outside and inside the piled
timber. In addition to the three-dimensionality and unsteadiness of the airflow, this
leads to a complex flow configuration in the kiln. The flow in a kiln is therefore
inherently non-uniform. The degree of this non-uniformity can be expressed by a
variation coefficient of air velocity:

1 1 n n
v
Vd =
vam
∑ (vl − vam )2 ,
n − 1 l =1
vam = ∑ l .
l =1 n
(13.1)

Regarding the spatial distribution of the flow, two characteristic sections can be
distinguished: the region outside the timber stack and the spaces between the boards
or packs. These are discussed in the following two sections.

13.2.1 Kiln-wide flow phenomena

The fans generate the airflow in the kiln. The flow capacity of the fans depends on
the pressure drop created by the stack and various kiln parts such as bends, baffles,
heating coils, and screens (see also Figure 13.1). The pressure drop of kiln parts is
related to the velocity via the pressure-loss coefficient ζ (13.2).

ρ
∆pL = ζ v2 (13.2)
2
ρ ρ
v12 + gz1 + p1 + ∆p P = v 22 + gz 2 + p 2 + ∆p L (13.3)
2 2
∑v
j
1j A1 j = ∑ v2 j A2 j
j
(13.4)

To calculate the flow through industrial equipment, it is common practice to follow


the path of a streamline along which, for an incompressible steady flow, an extended
form of Bernoulli's equation can be used (13.3). It contains the expressions for
kinetic and potential specific energy, static
pressure, as well as additional terms for pressure
loss (13.2) and pressure rise (e.g. generated by a
fan). With the help of the incompressible steady-
state continuity equation (13.4), the flow in duct
networks can be readily quantified.

In an overhead-fan timber kiln, the airflow is


turned after the fan battery, passing through a 90°
bend. The flow patterns shown in Figure 13.2
demonstrate one cause for pressure losses: flow
Figure 13.2. Flow
separation. This occurs when the flow is
separation in a 90° bend
Airflow within kilns 295

decelerated and the pressure rises in the


direction of the flow. Another reason for
pressure loss is the existence of secondary
flows, as depicted in Figure 13.3.

After the bend and following any additional


coils or baffles, the air stream reaches the
plenum chamber where it is slowed down
and distributed across the fillet spaces. A
typical velocity distribution at the stack
inlet over the plenum height is shown in
Figure 13.4. For identical plenums on both
sides of the stack, inertial effects are Figure 13.3. Secondary flow in
balanced against each other and hence a 90° bend
frictional effects determine the flow
distribution across the stack height. The
narrower the plenum chamber width, the
higher is the pressure drop over its height.
According to (13.3), the total pressure
along a streamline remains constant, and
thus the sticker space velocity decreases
down the height of the plenum chamber.

13.2.2 Stack flow phenomena

The flow down the plenum chamber


resembles the flow in a manifold with the
fillet spaces as its branches. The board
layers form channel walls, which exert a
drag on the fluid. The resulting wall shear Figure 13.4. Characteristic velocity
stress τW for a Newtonian fluid is distribution at the stack inlet across
proportional to the gradient of the velocity the plenum chamber height
normal to this wall:

∂vx
τW = − µ (13.5)
∂y y =0 .

The dimensionless shear stress in internal flow is called friction factor (λ) and can
be derived from a force balance for a rectangular duct (width>>height):

4τ W ∆p
λ= = = 4f
ρ Lρ (13.6)
vm2 vm2
2 2 2h .
The characteristic length for the force balance is defined as the ratio of the cross-
sectional area A and the wetted perimeter U. Since this ratio equals D/4 for a pipe,
296 Fundamentals of wood drying

4A/U is called the hydraulic diameter Dh.


Below a critical duct Reynolds number
laminar turbulent
ReDh,c the flow is laminar (ReDh,c = vDhρ/µ
≈ 3000 for flow between parallel plates
(Schlichting and Gersten 1997). Above
this value, disturbances can lead to a
turbulent flow regime where the velocity
fluctuates stochastically around a mean
quantity. Hence, the instantaneous Figure 13.5. Typical velocity
velocity may be represented as the sum of profiles for laminar and turbulent
the time-averaged value and the duct flow
fluctuating component:

v = v + v′ . (13.7)

The turbulence intensity Tu is related to the average fluctuating velocity and mean
velocity, and can be expressed for isotropic turbulence as:

Tu = v ′ 2 v . (13.8)

Typical mean velocity distributions for both laminar and turbulent regimes in
internal duct flows are shown in Figure 13.5. The model of a duct flow is just an
idealisation of the real configuration in the stack, since the stack layers are made up
of single boards, which do not form smooth channels. As illustrated in Figure 13.6,
the flow is disturbed by the blunt edge at the inlet and subsequent gaps and steps
within the stack. As a result, flow separation occurs, accompanied by a local sharp
increase of the shear stress as well as an intensified turbulence.

flow separation

developing flow
discontinuities produce
increased shear stress

Figure 13.6. Flow phenomena in the fillet space between stacked timber boards
Airflow within kilns 297

13.3 Measurement of flow field

An experimental investigation is useful to analyse the flow field in a timber kiln and
to eliminate possible sources of flow maldistribution. The measurement can take
place directly within the industrial kiln and can sometimes be incorporated into the
control system. However, for a more detailed examination, it is often necessary to
measure the flow patterns in small-scale models or water channels. The results can
be transferred to large-scale kilns using the principles of flow analogy, as
characterised by the Reynolds number.

13.3.1 Typical Measuring Equipment

Experimental fluid mechanics has made great progress in recent years. New optical
measuring principles as well as automation of signal processing and growing
computing power are the main reasons. However, the most important task remains
the same: the experimental acquisition of all necessary parameters that characterise
the flow of a fluid. This can involve measuring the flow field itself (velocity field
with direction and magnitude, temperature field, turbulence parameters) and
additionally measuring the flow domain boundary conditions like pressure, wall
shear stress, and heat transfer, as described in Table 13.1. For a more thorough
treatment of experimental fluid mechanics, the reader is referred to the standard
literature on this subject, e.g. by Ower and Pankhurst (1977), Perry (1982), and
Durst et al. (1981). In the following sections, only some selected equipment will be
highlighted and examples for their application in kiln airflow are mentioned.

Table 13.1. Flow field parameters and measuring devices


Pressure Probes for
Measurement Total pressure Pitot tube
Dynamic pressure Prandtl tube
Static pressure Disk probe
Pressure meters
Fluid manometer Inclined manometer
Membrane manometer Piezoresistive, Capacitive, Inductive
Velocity Pressure probes Prandtl tube, Spherical Pitot tube
Measurement Thermoelectric devices Hot-Wire/Film Anemometry
Optical devices Laser-Doppler Anemometry
Mechanical devices Vane Anemometer
Flow Surface visualisation Film techniques
Visualisation Tracer methods Particle, Smoke techniques
Optical methods Interferometry
298 Fundamentals of wood drying

13.3.2 Selected measuring techniques for airflow in kilns

The choice of a suitable measuring device depends on the flow parameters and the
size of the region to be investigated. From the numerous techniques mentioned
above, pressure tubes, hot-wire anemometry, and Laser-Doppler anemometry have
thus far been applied to airflow measurement in various kiln geometries.

a) Pressure probes

The measurement of pressure plays a central role in experimental fluid dynamics.


According to equation (13.3), the total pressure at any location is composed of static,
dynamic, and hydrostatic pressure. Specialised pressure probes come in various
sizes and shapes allowing each of these components to be quantified (Ower and
Pankhurst 1977). A typical use for pressure measurement is the characterisation of
friction forces. By obtaining the total pressure upstream and downstream of
obstacles, the representative pressure drop and friction factor (13.6) can be
calculated (see also “Stack pressure drop”). The total pressure can be measured with
the help of the well-known Pitot tube (Figure 13.7), which consists of a simple thin-
walled tube. An additional tube, flush with the wall surface, provides the static
pressure. The pressure difference ∆p between both probes yields the dynamic or
velocity pressure at the location of the Pitot tube’s nose so that the velocity can be
calculated as follows:

∆p
v= (13.9)
ρ air , where ρair is the fluid density.

An application example would be the velocity measurement in a channel separating


timber boards Ledig et al. (2001). A typical feature of pressure tubes is their
directional sensitivity. For a simple straight tube, the accuracy depends strongly on
the angle of attack of the probe, which must be within ±8° of the flow direction to
provide an accurate velocity measurement. However, a tapering of the wall
thickness of the probe at the tip can extend this range to ±20° (Figure 13.8). If the
flow direction is unclear and must be
0
determined, probes with multiple
openings can be utilised. One example is 0.1

0.2
(pα =0°− pα )/pα =0°

D 0.3

v∞ 0.4

0.5
L/D>20
0.6
Angle of attack α
0.7
pstat ptot -60° -50° -40° -30° -20° -10° 0° 10° 20° 30° 40° 50° 60°

Figure 13.7. Pitot tube and wall static Figure 13.8. Angle of attack influence
tube in a horizontal flow on measured pressure
Airflow within kilns 299

the five-hole spherical Pitot tube,


which permits a three-dimensional
measurement (Figure 13.9). The semi-
spherical head carries 5 evenly spaced
holes. During calibration, the relative
pressure differences across these open-
ings are recorded depending on the
flow angles α and β. During measure- Figure 13.9. Five-hole spherical Pitot
ment, the magnitude and direction of tube with coordinate system
the velocity vector can be interpolated
from these characteristic values. The
previously described probes produce a pressure signal. However, to record and
visualise this information, a conversion of the pressure into a suitable signal for
further processing is required. This task is performed by manometric devices.
Simple fluid manometers allow a direct visual reading, whereas membrane
manometers transform the pressure into an electric signal, which can be processed
by a computer.

b) Hot-wire anemometry

This measurement technique belongs to the group of thermoelectrical methods. It is


one of the classic systems for velocity measurement providing high accuracy and
flexibility. Its particular importance lies in the determination of rapid changes in
velocity, which allows the study of turbulence, characterised by parameters such as
the turbulence intensity Tu (13.8). Figure 13.10 presents the main components of
this device. Hot-wire anemometry is an indirect method which uses the influence of
an external flow on the heat transfer from a heated wire to measure the flow
velocity. The heated sensor which can be either wire, film or fibre-film, is kept at a
constant temperature (constant temperature anemometer - CTA). The magnitude of
the current required to achieve this varies with the external heat transfer and hence
fluid velocity v (Figure 13.11). Another less common operating mode is to maintain
a constant current I and monitor the changing wire temperature and thus resistance
RW (constant current anemometer - CCA). Single-sensor probes provide just one
velocity component directed normal to the wire. X-probes and Tri-axial probes
employ up to 3 wires and allow a measurement of two- or three-dimensional flows.

Figure 13.10. Typical CTA measuring chain Jorgensen (2002)


300 Fundamentals of wood drying

A number of specialised probes are Sensor Supports Sensor, Resistance RW


(Stainless Steel) (Thin Wire: Diameter ~5 µm
available for particular applications, Length ~1 mm)
such as flow in a boundary layer or
at high temperatures. Hot-wire ane-
mometry has been used by some
workers to measure velocity profiles Current I Velocity v
in the channels between the timber
boards Wu et al. (1995), Wiedemann Generated Heat QΩ Transferred Heat Qconv

(1989), Gillwald and Tschirnich


(1967). More in-depth information Figure 13.11. Principles of operation
on this subject can be found in text- for a hot-wire anemometer
books, e.g. by A.E. Perry (Perry
1982).

c) Laser-Doppler anemometry

In contrast to the above mentioned methods, which demand the insertion of a probe
into the flow path, Laser-Doppler anemometry (LDA) allows the non-evasive
examination of the flow field. It is a well-proven optical technique with a key
feature that it is capable of high spatial resolution. The main components of a Laser-
Doppler anemometer are drawn up in Figure 13.12. Most LDA units today are
equipped with a combined transmitter and detector unit, thus having the advance of
requiring access to the measurement zone from one side only. In the transmitting
optics, the generated monochrome laser is split into two separate beams, which are
focused at the measuring-volume. The velocity measurement is based on the
Doppler effect and therefore requires no calibration. This physical effect can be
illustrated with the help of Figure 13.13. A particle travelling through the region of
two intersecting laser beams scatters the light with a frequency proportional to its

Figure 13.12. Main components of a Laser-Doppler anemometer (Nitsche 1994)

Figure 13.13. Interference pattern of intersecting laser beams (Nitsche 1994)


Airflow within kilns 301

velocity. Because the angle Θ between the beams as well as their wave length λ are
known, the velocity measurement is reduced to the determination of the Doppler
frequency f (equation (13.10)), which is derived from the light signals arriving at the
photodetector unit.


v= (13.10)
2 sin (Θ / 2)

Since the LDA measurement relies on particles moving through the interference
pattern, special generators are used to produce smoke or liquid dispersions upstream
of the measurement points. More velocity components can be determined
simultaneously when additional laser beam pairs with different wavelengths are
focused at the same measuring volume. In most LDA systems, a Bragg cell produces
a frequency shift in one of the beams of each pair, which allows measurement of the
flow direction. Laser-Doppler anemometry has been applied to the study of flow
between rough-sawn timber boards (Leiker and Ledig 2001) (see also “Flow
between the board rows”). A comprehensive description of the different variations
of Laser-Doppler anemometry and their applications can be found in the literature
(Durst et al. 1981).

d) Flow visualisation

The previously discussed experimental methods are essentially point flow measure-
ment techniques. In contrast, visualisation methods provide the user with a global
flow picture. Flow visualisation has been an important tool for experimental fluid
dynamics because of its simplicity and straightforward application of flow manifes-
tation in familiar environmental patterns. Before proceeding with detailed measure-
ment and mathematical modelling, visualisation techniques help to obtain an under-
standing of the flow phenomena. Usually, the results are of a qualitative nature, but
the new possibilities of digital image processing allow quantitative results too. Visu-
alisation methods can be generally categorised into those techniques requiring the

Figure 13.14. Oil-Soot Film Flow Visualisation (negative image) for flow through
the stack of an end-mounted fan kiln: top view of board surfaces for middle stack
layer (Ledig and Militzer 1999a), dashed kiln outline only for clarity – not to scale
302 Fundamentals of wood drying

introduction of particles into the flow, and optical methods (Table 13.1). In the spe-
cial case of surface visualisation, the particles are attached to a wall of the flow
domain. Figure 13.14 shows an example of this technique in which a mixture of oil,
soot, and petroleum was used to examine the effect of end-mounted fans in a kiln.
The picture (negative image) shows the two-dimensional flow between two layers of
the timber stack. Large bright-coloured regions in the right half indicate low veloci-
ties in a recirculation zone. Dragged soot particles show the direction of the flow,
which makes 90°-turns at the stack entry and exit.

Several workers have used water channel models to simulate the two-dimensional
flow in a kiln section along the stack height and width (Nijdam and Keey 1999,
Krischer and Kroell 1978, Gillwald and Tschirnich 1967, Kroell 1954, Sturany
1952). The photos in Figure 13.15 were taken by Nijdam and Keey (2002), who
used cavitation-induced and hydrogen bubbles to visualise and measure the fillet-
space velocities for various designs of the right-angled bend after the fan. Other
researchers employed visualisation methods in investigating the flow between
boards in a wind tunnel. To study the time-dependent flow in an array of timber
boards, Lee (1990) has conducted visualisation experiments. He used smoke to
observe the formation of eddies in the horizontal gaps between adjacent boards in a
single row. An example of optical visualisation techniques can be found in a
publication by Longauer (1994) who applied interferometry to obtain temperature
profiles in a channel between two board layers. Further general material about
visualisation techniques can be found in the literature, e.g. by Merzkirch (1987).

(a) (b)

Figure 13.15. Flow patterns in a hydraulic kiln model with (a) no semicircular
section, (b) a semicircular section of c = 0.3 (c-ratio of section radius to maximum
height of ceiling space), ratio of plenum space width to the sum of the fillet-space
widths equals 1.385, equivalent average between-board air speed is 8 m/s at 400 K
(Nijdam and Keey 2002)
Airflow within kilns 303

13.4 Airflow simulation

The experimental methods outlined in 13.3.2 are important to obtain information


about the flow field in complex domains. However, because of the large dimensions
of typical industrial applications, detailed measurements have to be restricted to
selected locations, and results are only valid for the given configuration of kiln and
stack. An optimisation of the airflow in a kiln and the analysis of stacking strategies
will therefore require a simulation tool to perform case studies in a more cost-
effective way. Airflow simulation uses computer-based numerical methods to solve
the governing equations of fluid flow for more general cases, and thus complements
the experimental methods. This field is commonly called computational fluid
dynamics (CFD). Airflow simulations have been carried out by a number of re-
searchers to investigate both global flow patterns in kilns as well as certain localised
flow details. Kiln-wide simulations have been carried out to test the influence of kiln
and stack geometry parameters, or positions of fans and baffles (Hua et al. 2001,
Ledig and Militzer 1999a, Ledig and Militzer 1999b, Riepen and Gard (1996),
Arnaud et al. (1991). Details of kiln airflow studies using numerical methods include
flow in the plenum chamber Langrish (2002), bypass flows due to ragged stack ends
Langrish and Keey (1996), and local flow inside the stack in the region of gaps and
steps between horizontal boards (Sun 2001, Esfahanian 2000, Esfahanian et al.
1999, Salin et al. 1998, Langrish 1994, Langrish et al. 1992, Langrish et al. 1993).

13.4.1 Conservation equations

The basic equations for the movement of a fluid are derived from conservation laws
for mass, momentum, and energy of a fluid element. The derivation of these
equations is omitted here for reasons of brevity and is covered exhaustively in the
standard fluid mechanics literature. The equations will be given in index notation
and Cartesian coordinates for the case of an incompressible flow of a pure fluid,
which is a valid approximation for the conditions in timber kilns. A mass balance for
a fluid element yields the continuity equation:

∂vi ∂v1 ∂v2 ∂v3


= + + =0 (13.11)
∂xi ∂x1 ∂x2 ∂x3 .

Newton's 2nd Axiom states that the change of momentum of a body equals the sum
of the forces acting on it. Applying this law to a fluid element leads to Cauchy's
equation of motion. For a Newtonian fluid (4.1), where the stress components
depend linearly on the rates of deformation, the Navier-Stokes equations are6:

 ∂vi ∂vi  ∂p ∂   ∂v j ∂v 
ρ  + vj

= ρg i − + µ  + i 
 . (13.12)
 ∂t ∂x j  ∂x i ∂x j   ∂xi ∂x j 

6
For the special case of an inviscid flow (µ = 0, 1/Re→0) the Euler equation can be derived, which yields
for a steady flow and integrated along a streamline the previously stated Bernoulli equation (13.3).
304 Fundamentals of wood drying

The forces in the momentum balance can be categorised into body (e.g. here
gravitational) and surface forces due to pressure and stresses. If heat transfer is
included in the analysis, the energy equation must also be solved. This equation is
derived from the first law of thermodynamics, which asserts that the internal energy
of a closed system changes by transferred heat and work done on it. Using the
enthalpy formulation leads to the following expression:

 ∂h ∂h  ∂  ∂T  dp ∂v j  ∂v j ∂vi 
ρ  + vj = λ + +µ  +  (13.13)
 ∂t ∂x j  ∂x j  ∂x j  dt
 ∂xi  ∂x ∂x
 i j
 .

13.4.2 Turbulent flows

The airflow in kilns is largely of a turbulent nature. Turbulent flows are


characterised by fluctuating motions overlaying the mean flow. Although the equa-
tions in 13.4.1 describe the instantaneous fluid flow exactly, the resolution of the
high frequency oscillations by direct numerical simulation (DNS) is computationally
too expensive for most practical applications. The common approach is the so-called
“Reynolds averaging” of the conservation equations, in which the instantaneous
quantities are replaced by the sum of a mean and a fluctuating component (13.7) and
then time-averaged. This leads to the following forms of equations (13.11) and
(13.12), which are called the Reynolds-averaged Navier-Stokes (RANS) equations:

∂vi ∂v1 ∂v2 ∂v3


= + + =0 (13.14)
∂xi ∂x1 ∂x2 ∂x3 ,
 ∂v ∂v  ∂p ∂   ∂v j ∂vi  
ρ  i + v j i  = ρg i − + µ +  − ρ vi′v′j  (13.15)
 ∂t ∂x j  ∂x i ∂x j   ∂xi ∂x j 
  .

Different turbulence models are used to describe the additional unknowns in the
momentum equations − ρ vi′v′j , which are called the “Reynolds stresses”. A simple
model is the Boussinesq hypothesis, which relates the Reynolds stresses to the
velocity gradients:

 ∂v j ∂vi  1
− ρ vi′v′j = µ t  +  − ρ v′v′δ (13.16)
 ∂x ∂x  3 i i ij .
 i j 

Usually, the turbulent viscosity µt is modelled as a function of the turbulent kinetic


energy k = 1 / 2vi′v′j and the turbulent dissipation rate ε leading to two additional
transport equations, which are, in the case of the standard k-ε model (Laudner and
Spalding 1974) excluding buoyancy and compressibility effects, written as follows7:

7
Except for the Reynolds stresses, the overbar indicating mean values is dropped in common notation
hereafter.
Airflow within kilns 305

 ∂k ∂k  ∂  µt  ∂k  ∂vi  ∂vi ∂v j 
ρ  + vj

=  µ +   + µt + − ρε (13.17)
 ∂t ∂x j  ∂x 
j  σ k  ∂x j  ∂x j  ∂x j ∂xi  ,
 ∂ε ∂ε   ∂   µ  ∂ε  ε ∂vi  ∂vi ∂v j  ε2
ρ  + v j  ∂x 
= µ + t   + C1 µt + − C2 ρ (13.18)
 ∂t ∂ x j  j  σ ε  ∂x j  k ∂x j  ∂x j ∂xi  k ,
where µt = ρCµ k2/ε . (13.19)

In the above equations, σk and σε are the respective turbulent Prandtl numbers for k
and ε. The terms Cµ , C1 and C2 are constants in the high Reynolds number
formulation, whereas in a modified formulation they may vary with the turbulence
Reynolds number ( Ret = ρ k2/(εµ) ) to account for low-Reynolds-number effects.

Compared to the above-mentioned two-equation turbulence models, a more realistic


but computationally more expensive calculation is to solve transport equations for
each component of the Reynolds stress tensor (Reynolds stress models - RSM).
Finally, a solution method called large eddy simulation (LES) employs a
compromise between DNS and RANS methods by averaging only small-scale
motions while simulating “large” eddies directly.

13.4.3 Notes on use of CFD

Given that the partial differential equations describing fluid flow cannot be solved
analytically except for special cases, they are normally transformed into a system of
algebraic equations by numerical discretisation. Several methods can be applied for
this purpose, the most important of which are finite difference (FD), finite volume
(FV), and finite element (FE) methods. They are covered in general fluid mechanics
literature as well as in numerous CFD textbooks, e.g. Patankar (1980), Ferziger and
Peric (1997).

A numerical solution method contains the following main components:

Mathematical model: Defined by the set of equations and boundary conditions.


Discretisation method: Transforms differential equations to algebraic equations.
Numerical grid: Discrete representation of the problem domain.
Solution method: Schemes (mostly iterative) to solve algebraic equations
until convergence criteria are met.

In order to make CFD calculations feasible, it is often necessary to make


simplifications to the boundary conditions and mathematical modelling approaches.
These simplifications result in differences between the real experimental airflow
distribution in the kiln and the predicted airflow distribution. In addition to these
modelling errors, there are ‘numerical errors’ introduced by the solution procedure.
In Table 13.2 an overview is given of possible errors when CFD is applied to model
airflow in kilns.
306 Fundamentals of wood drying

Table 13.2. Errors introduced by CFD models for airflow in kilns


Solution of
Reality Mathematical Model
Mathematical Model
Kiln with: Model assumptions: Not exact due to:
- Airflow - Navier-Stokes with con- - Discretisation on
- Wet air stant properties computational grid
- Temperature - Pure fluid - Break off error of it-
variations - Air properties not known erative solver
- Non flat surfaces exactly
- Rough surfaces - Simplified geometry
- Turbulent flow - Turbulence model
- Stack with boards - Surface roughness not
known
- Stack model

Due to the approximate nature of the predicted airflow distribution (numerical


solution), the application of CFD for engineering purposes usually demands the
experimental verification of the modelling approach. However, the general character
of the numerical simulation model helps to plan and limit the extent of these
experiments. The main steps of a CFD supported flow analysis could then be as
follows:

1. Measurement of flow parameters for a base case.


2. Simulation of the given configuration and comparison with measured results.
3. Software parameter studies using CFD to find best possible configuration.
4. Experimental verification of numerically selected modification.

13.4.4 Examples of airflow simulation

In the following sections, two examples will illustrate the application of the
previously discussed CFD tools. Firstly, a 2D model is used to investigate the flow
in an overhead fan kiln. In the second section, results for the 3D flow in a kiln with
an end-mounted fan are presented.

a) Flow in an overhead fan kiln

The phenomenon investigated here is the global flow pattern in an industrial-scale


convective drying kiln. A 2D model has been set-up in the commercially available
CFD code TNO-WISH3D. The conservation equations for mass, momentum and
turbulence, in the form of the RANS equations (respectively (13.14), (13.15),
(13.17) and (13.18)) are solved for an incompressible and steady-state case with the
internal circulation fan as the only flow-generating source. The standard k-ε model
Airflow within kilns 307

Chamber
Plenum

Figure 13.16. Geometry of a high temperature kiln by Gooskens B.V., Netherlands

(13.4.2) is applied to model the turbulence. Near the walls, the turbulent boundary
layer is approximated by standard wall functions. The stack is modelled using a
porous medium model. This model simulates the average turbulent flow through a
porous stack rather than simulating the flow around all boards individually. The
porous medium model calculates the correct pressure drop and flow velocities
within the stack, based on the porosity of the stack and the hydraulic diameter of the
open space between the boards. This approach reduces the modelling effort while
preserving the required accuracy of the numerical solution.

To illustrate the application of airflow simulation, the airflow in the high


temperature kiln of Gooskens B.V. in Moerdijk is analysed for two cases:

- reference case: airflow in the kiln with a complete stack


- staggered stack case: airflow in the kiln partially loaded with
packages in a staggered orientation.

The geometry of the kiln, as modelled in TNO-WISH3D, is shown in Figure 13.16.


Note the relatively narrow plenum width in comparison to the space between the
inner ceiling and the top of the kiln. The airflow capacity of the fan is 5.0 m3/s per
meter kiln length. The board thickness is 50 mm and the sticker dimension 22 mm.
308 Fundamentals of wood drying

Figure 13.17. Air velocity distribution of the reference case: boards 50 mm, stickers
22 mm, grey contours show magnitude of air velocity (see legend)

Results of the reference case

The calculated airflow distribution is shown in Figure 13.17. The flow through the
stack is rather uniform with slightly higher velocities in the top part of the stack. It
deviates only 8% from its mean value (mean velocity is 4.2 m/s, standard deviation
is 0.34 m/s). The total stack width in this kiln is relatively narrow. The good flow
uniformity over the height of the kiln stack is a result of a uniform downward flow
at the entrance of the plenum chamber near the inner kiln ceiling. A vertical airflow
at the inner ceiling level avoids a swirling motion in front of the stack, which would
result in substantial flow non-uniformities especially in the upper portions of the
stack.

Variation on the reference case: “Staggered Stack”

If there is not enough timber available to fill up the drying kiln completely, kiln
operators often run a partially loaded kiln. The following situation has been
investigated to illustrate the effect of additional open space on the flow distribution
in a kiln. The same kiln geometry is used as in the reference case, but the length of
the timber packages is not sufficient to fill up the total length of the kiln. To avoid
bypassing effects, the kiln operator decides to place the packages in a staggered
position to fill the whole length of the kiln. The distance between the edge of the
Airflow within kilns 309

stack and the end wall is enormously


enlarged in this situation and, as a
1200 mm consequence, a large open space below
the kiln ceiling is created, which can
badly influence the uniformity of the air
velocity distribution (Figure 13.18).
1000 mm
In Figure 13.19, the calculated air
velocity distribution at cross section A-A
300 mm is shown. In the open area in front of the
stack, a recirculating flow will occur,
1200 mm which causes poor flow uniformity. Since
the total stack width is only half the total
width of the stack in this cross section,
the pressure drop over the width of the
stack is much lower than in the reference
A A case, thus further enhancing flow non-
uniformities (see also 13.5.2). Comparing
Figure 13.18. Top view of a partially Figure 13.17 to Figure 13.19 shows
filled kiln with staggered stacks clearly that the homogeneity in flow is
decreased dramatically by placing the
packages in a staggered position.

Figure 13.19. Air velocity distribution in a partially filled kiln with staggered stack
packages, grey contours show magnitude of air velocity (see legend): cross section
A-A in Figure 13.18
310 Fundamentals of wood drying

b) Flow in an end-mounted fan kiln

The phenomenon investigated here is the global flow pattern in a laboratory-scale


superheated steam vacuum dryer (Ledig and Militzer 1999b). A complete 3D model
was developed using the commercial pre-processor Gambit by Fluent Inc. Applying
the finite-volume solver Fluent 5 (Fluent… 1998) of the same manufacturer, the
RANS equations for mass and momentum (13.4.2) were solved for an
incompressible and steady-state case with the fan as the only flow generating source.
For turbulence modelling, a variant of the k-ε model was applied (Shih 1995). Near
the walls, the turbulent boundary layer was approximated by standard wall functions
(see – “stack entrance effects”).

Figure 13.20 depicts the absolute velocity distribution in a plane between the middle
board layers for the whole kiln. The main flow feature is the separation zone on the
lee side of the second sticker, which is also present in a sharp-angled bend (Figure
13.2). Consequently, gas velocities are lowest in the part of the stack nearest to the
fan and increase along the stack length. These results confirm the flow pattern
obtained from measured gas velocities (Ledig and Militzer 2000) and visualisation
studies in the same drying kiln (Figure 13.14). Applying a screen in front of the
stack entry face, which straightens and homogenises the flow along the stack length,
can prevent the formation of that separation zone (Ledig and Militzer 1999a).

Axial Fan

Stack Outline
Sticker Positions

Figure 13.20. Gas flow in a laboratory vacuum kiln with end-mounted fan: contour
plot of simulated absolute velocity field for a surface between board layers at
medium stack height (in m/s); main flow is horizontal, clock-wise
Airflow within kilns 311

13.5 Airflow analysis

While the previous sections introduced experimental and theoretical methods for
studying flow fields in a drying kiln, the following sections will present results of
their application in analysing the main flow features both inside and outside the
stack. Whereas the first section will focus on the details of the flow inside the stack,
the second section will deal with the kiln-wide flow situation providing some
engineering guidelines to help with an appropriate design of drying kilns.

13.5.1 Airflow in the stack

Fluid dynamics in the stack is characterised by several phenomena. At the stack


entrance, the behaviour is similar to the flow around blunt plates. Further
downstream, the upper and lower boundary layers are thickening and finally grow
together establishing a channel flow profile. To describe the situation correctly,
discontinuities along each airflow channel, which result from variations in board
thickness and gaps between boards, and board surface roughness effects should also
be addressed. Due to the analogy of momentum and heat/mass transfer the
subsequent descriptions have also consequences for the external heat and mass
transfer characteristics during drying, which are described in the chapter “External
heat and mass transfer”.

a) Flow between the board rows

The time-averaged flow in the channel between timber boards has been investigated
by Wiedemann et al. (1989), who applied hot-wire anemometry to measure mean
velocity profiles across the duct height. A similar experiment was conducted by Wu
et al. (1995) who recorded velocity profiles at various stations along the channel
length. The results, which are presented in this section, were obtained with a Laser-
Doppler anemometer (LDA), which allows the examination of a flow pattern
undisturbed by a measurement probe (Leiker and Ledig 2001). The main focus in
the next sections will be on the time-averaged flow pattern, which is of foremost
interest for the kiln-drying process. Some phenomena of the instantaneous flow will
be explained in “Discontinuities and instantaneous flow” together with the effects of
surface discontinuities in the channel.
35
y [mm]

30
4m/s
25
vX
20
15
10
5
0

0 x [mm] 20 40 60 80 100 120 140 160 180 200 220

Figure 13.21. Vector plot of measured mean x-velocities in the 220 mm long
entrance section of a channel (h=35 mm, vm=4.12 m/s, Re2h=19290) between 38 mm
thick boards (roughness K/(h/2)=0.022) in a model timber stack (535 mm wide,
460 mm long, 610 mm high), dashed lines (---) indicate separation zones
312 Fundamentals of wood drying

Stack entrance effects

Figure 13.21 shows the typical flow pattern in the entrance region of a timber stack.
At the inlet of the duct, the board rows represent blunt plates. From the stagnation
point in the front face of the plate, symmetrical boundary layers develop, which
detach at the edge if a critical, plate thickness based, Reynolds number
(Ret,c=vtρ/µ=100 Lane and Loehrke 1980) is exceeded. This condition is always met
for the typical board dimensions, air velocities, and turbulence levels in wood kilns
and hence a separation bubble is formed. The velocity profile in the middle of the
duct is constant with slightly higher velocities in the vicinity of the bubble. The flow
reattaches to the walls after the separation zone and new boundary layers grow.
After a sufficient length, both boundary layers join and form a developed duct
velocity profile. The reattachment length depends on the turbulence of the
approaching flow and the Reynolds number Ret and is in the range of 10 to 50mm
(Leiker and Ledig 2001, Kho et al. 1989, Ota et al. 1981, Kottke et al. 1977a) for the
configuration found in timber stacks where higher blockage ratios (t/h) lead to
shorter bubbles than for single plates. The characteristics of separated flow around
blunt plates are comprehensively reviewed by Ota (2000).

The development of the boundary layer can be characterised by the shape factor H
of the velocity profile, which is defined as the ratio of displacement thickness δ* to
momentum thickness δ**:

h2 h2
δ*  v  vx  vx 
δ ** ∫0  v∞  ∫v
H= = 1 − x dy 1 − dy (13.20)
0 ∞  v∞  .

Schlichting (1958) gives shape factors that range from 2.3 to 3.5 for laminar
boundary layers and from 1.3 to 2.2 for turbulent layers. Thus, the decrease of the
shape factor in Figure 13.22 indicates a transition in the boundary layer regime from
laminar to turbulent with growing distance. A turbulent boundary layer can be safely
assumed for Rex>20,000. The following charts will illustrate the different flow
regimes. For an average velocity of 0.25 m/s, no separation was detected and a
3
0.3 m/s 15mm
1.4 m/s
2.7 m/s
2.5 5 m/s
7 m/s
0.3 m/s 25mm
1.4 m/s
2 2.7 m/s
H

5 m/s
7 m/s
Figure 13.22. Shape factor
0.3 m/s 35mm H as a function of Rex =
1.5 1.4 m/s
2.7 m/s vmxρ/µ in the entrance
5 m/s section of a channel for
7 m/s
1 various heights h and
0 20000 40000 60000 80000 100000 120000
free-stream velocities v∞
ReX
Airflow within kilns 313

1 1
v /v ∞ v /v ∞
0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6
Blasius
0.5 10 mm 0.5
x =
20 mm
0.4 30 mm 0.4 (1/4)
/δ )
(y/delta)^(1/n)
(y
50 mm
x = 50 mm
0.3 70 mm 0.3
70 mm
90 mm
90 mm
0.2 110 mm 0.2
125 mm 110 mm
170 mm 125 mm
0.1 0.1 170 mm
220 mm
220 mm
0 0
0 1 2 3 4 5 6 0 0.2 0.4 0.6 0.8 1
y* y /δ

Figure 13.23. Measured boundary layer Figure 13.24. Measured boundary


mean velocities for various distances x layer mean velocities for a duct
from the entrance of a duct (h=35mm, (h=25mm, vm=1.2m/s,
vm=0.25m/s, Re2h=1160, K/(h/2)=0.022) Re2h=3880, K/(h/2)=0.031) and
and laminar theory power law correlation
1
laminar boundary layer developed shortly v /v ∞
0.9
after the leading edge as shown in Figure
13.23. The solution of Blasius’ equation 0.8
(Blasius 1908) is included for comparison,
0.7
and correlates well with the data. The
similarity parameter y* is defined as 0.6

0.5
y
y* = Re x (13.21)
x . 0.4 (1/5)
(y /δ )
(y/delta)^(1/n)
x = 50 mm
0.3
70 mm
For a velocity of vm=1.2m/s, the profile after 90 mm
reattachment in Figure 13.24 shows a 0.2 110 mm
transitional shape with a more laminar 0.1
125 mm
170 mm
gradient near the wall and a rather turbulent 220 mm
appearance in the core. A further increase of 0
0 0.2 0.4 0.6 0.8 1
the velocity results clearly in a turbulent y /δ

boundary layer, as shown in Figure 13.25, Figure 13.25. Measured boundary


which can be approximated by a simple layer mean velocities for a duct
power law: (h=35mm, vm=4.1m/s,
Re2h=19290, K/(h/2)=0.022) and
1n
vx  y  power law correlation
=  (13.22)
v∞  δ  .
314 Fundamentals of wood drying

In comparison with smooth walls, for which n=6…7, the velocity profile for a rough
wall follows a curve, for which n is between 4 and 5, which corresponds well with
the experimental data presented by Schlichting (1958). Although the power law
describes the shape of the turbulent core rather well, it fails to quantify the wall
shear, since the velocity gradient at y=0 is not defined. A more elaborate model of
the turbulent velocity profile, which takes into account the characteristic 3-layer
structure of a turbulent boundary layer, is able to address this deficiency:

Viscous sublayer, v+ = y+ , 0<y+<5 : (13.23)


Logarithmic and Wake layer, y+>30: v x+ = 1 ln y + + B − ∆B + Π w y  (13.24)
κ κ δ  ,
+ + 0.5
where v = vx/vτ , y = yvτρ/µ , vτ = (τW/ρ) .

The terms ∆B and w(y/δ) are called the roughness and wake functions (Π-wake
strength), respectively. Correlations for these terms can be found in fluid dynamics
handbooks. Typical values for the constants κ and B are 0.4 and 5.5. The surface
roughness has no influence on the boundary layer flow as long as the roughness
elements are submerged in the viscous sublayer, i.e. in the case where K+=Kvτρ/µ
≤5, the flow is dynamically smooth (∆B=0). The region 5<K+<70 is often referred to
as “transitionally rough”, and for K+≥70 the surface is called “fully rough”.

Channel flow characteristics

Following the entrance region, both boundary layers have grown together and a
developed duct velocity profile is formed. In the laminar case (Figure 13.26), a
1.0 1.0

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6
v /v ∞

v /v ∞

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2
laminare
Poiseuille equation U/Umax-theory
turbulent theory (13.26)
0.1 Kanalströmung 0.1
Messung:
experiment220 experiment
220 mm
mm
0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
y /h y /h

Figure 13.26. Fully developed laminar Figure 13.27. Fully developed turbulent
velocity profile in a duct (L= 220mm, mean velocity profile in a duct (L=220mm,
h=15mm, vm=0.23m/s, Re2h=460) h=25mm, vm=2.3m/s, Re2h=7800)
Airflow within kilns 315

parabolic distribution is established which is the so-called two-dimensional


Poiseuille flow:

vx 4y  y
= 1 −  (13.25)
vmax h  h .

In turbulent flow, the velocity profile displays a “fuller” shape than in the laminar
case, with higher velocity gradients near the wall, as shown in Figure 13.27. The
profile of the turbulent core for both smooth and rough walls can be correlated by
equation (13.26), which was proposed by Churchill and Chan (1995):

 9.025 y +  15 2 10 3
v + = 2.5 ln  + Y − Y , where Y = 2y/h . (13.26)
 1 + 0.301K  4 3
+

The transition from laminar to turbulent flow is influenced by the ratio of board
roughness to channel height. In the transitional region (Re2h=2500-7000), a higher
relative roughness leads to a turbulent regime at a lower Reynolds number Re2h, i.e.
for lower velocities at smaller channel heights (Leiker and Ledig 2001).

Discontinuities and instantaneous flow

As has been stated in “Stack entrance effects”, the flow detaches at the leading edge
where the air enters the stack. The separation bubble, however, is only stable at low
Reynolds numbers, and quickly evolves to a turbulent state at higher velocities, in
which large-scale vortex structures are shed from the reattachment region similar to
the case of flow around a thick plate (Ota 2000). These are convected downstream
and still influence the flow far away from the leading edge.

In the previous sections, the stack has been viewed as a pack of plates assuming no
horizontal gaps between the boards. This is an unrealistic idealisation because, in
practice, gaps will certainly form due to timber shrinking, even if the boards have
been stacked tightly together at the beginning of the drying process. These gaps and
any steps formed by recessed boards interrupt the duct walls, therefore disturbing
the boundary layers and increasing the turbulence and hence pressure drop of the
flow through the fillet space. As with the leading edge, the increased shear may lead
to the formation of eddies, which are continuously released from the gaps, thus
enhancing the transfer of momentum, heat, and mass in the boundary layer
downstream. The effect of gaps in wood stacks has been experimentally and
numerically investigated by Kho et al. (1989), Langrish (1994, 1998), Langrish et al.
(1992, 1993), Esfahanian (2000), Esfahanian et al. (1999), Sun (2001), and Hua et
al. (2001) (see also introduction to section 13.4).

Another feature characterising a real stack is the presence of non-aligned board


ends, which are caused by uneven board lengths. The resulting flow configuration is
a succession of expansions and contractions, while the overall larger channel sizes at
316 Fundamentals of wood drying

the ends of the stack produce a much lower resistance to flow than in the central
portion of the stack. This leads to bypass of airflow around the ends of the stack (see
also 13.5.3). The typical value of 10...30% bypass flow for a stack with well-aligned
ends can increase to as much as 100% for a stack with ragged edges, as reported by
Langrish and Keey (1996).

b) Stack pressure drop

The timber stack (Figure 13.28)


represents a flow resistance within the
drying kiln, which is balanced by h
continuous energy supply by the fans. v0
The average velocity in the channel vm is
related to the stack pressure drop ∆pStack t
via the pressure-loss coefficient ζ (13.2): L

ρ Figure 13.28. Geometry of a timber


∆pstack = ζ vm2 (13.27) stack section
2 .

The overall pressure loss coefficient ζ (13.28) contains the loss terms for the stack
inlet ζi and outlet ζo, which include the effects of contraction, velocity profile
development, and expansion. Within the stack, the coefficient ζf (13.31) accounts for
wall friction (see (13.6)) and ζg describes the loss due to gaps between boards:

ζ = ζi +ζ f +ζ g +ζo . (13.28)

The stack outlet loss coefficient is easily obtained from the Borda-Carnot equation:

(
ζ o = (1 − A1 A0 )2 = 1 − (1 + t h )−1 )
2
, (13.29)

where A0 and A1 are the cross sections of the larger and smaller duct, respectively.
The stack inlet loss coefficient ζi accounts for the loss due to contraction, possible
separation, and the loss due to the developing flow. Gerbery (1966) suggested an
empirical formula summarising the loss coefficients for the inlet and outlet of a
timber stack. A modified version of that equation, which agrees well with experi-
mental results by Gerbery (1966), Darakov (1977), and Leiker and Ledig (2001) for
the range of applicable velocities and sticker sizes between 10 and 50mm, is:

(
ζ io = 2 1 − (1 + t h )−1 ) 2
. (13.30)

For the flow resistance due to wall friction in the channel, excluding inlet and outlet
losses, the following expression is applicable:

ζ f = λ ⋅ L 2h = f ⋅ 2 L h , (13.31)
Airflow within kilns 317

where the friction factor (13.6) has to be determined. For laminar flow (Re2h < 2500)
(13.25), the average velocity is vm = ∆ph²/(12µL), and the friction factor is thus:

λlam = 96 Re 2 h . (13.32)

For the transition region, a good estimate of the friction factor can be obtained by
taking the average of the laminar friction factor at Re2h = 2500 and the turbulent
friction factor at Re2h=7000 for the given roughness. The friction factor for turbulent
flow (Re2h>7000) can be derived by integrating the velocity distribution over the
cross section since vm+ = (8/λ)0.5. Analogous to Colebrook’s correlation for flow in
smooth and naturally rough pipes (Colebrook 1938/39), a relationship can be de-
rived for parallel plates (Churchill and Chan 1994), which is, using the logarithmic
part of (13.26), as follows:

1  3.03 K h
= −2 lg  +  (13.33)
λ turb  Re 2 h λ turb 6.12  .

For practical purposes, an explicit correlation for the friction in smooth pipes, as
suggested by Colebrook, was used by Churchill to derive an equation for the
developed flow in smooth and rough pipes (Churchill 1973). This can also be done
for the flow in a channel; an equation explicit in λ for smooth walls is:

1  Re 
= −1.8 lg 2h  (13.34)
λturb  8.4  .

Therefore, the friction factor correlation for both smooth and rough walls is given
by:

1  8.4 
0.9
K h
= −2 lg   +  (13.35)
λ turb  Re 2 h  6.12  .

Friction factor plots according to equations (13.32) and (13.35) are presented in
Figure 13.29. Measured values of friction factors for timber stacks can be found in a
few papers (Longauer and wigo 1997, Wu et al. 1995, Darakov 1977, Viktorin
1968, Gerbery 1966). They are typically in the range from 0.04 to 0.08.

A problem that has not been addressed yet is the influence of measured timber
surface texture on the flow. It is unclear what equivalent roughness parameter corre-
sponds to the roughness height used in the previous equations (13.33), (13.35).
However, since the flow is only influenced by surface roughness if the elements pro-
trude outside the viscous sublayer, a parameter characterising the peak to valley
distance should be preferred to area-averaged values. Roughness heights reported in
the literature vary between 0.2 and 0.5 mm. A further complicating factor is that,
318 Fundamentals of wood drying

0.10
laminar turbulent
0.09

0.08

0.07
K /h =0.060
0.06
λ

K /h =0.030
0.05

K /h =0.015
0.04
K /h =0.005
0.03
K /h =0

0.02

0.01
1000 4000 7000 10000 13000 16000 19000 22000 25000 28000
Re2h
Figure 13.29. Friction factors λ for laminar (13.32) and turbulent (13.35) flow
between parallel plates as a function of Re2h=vm2hρ/µ and relative roughness K/h
even with careful stacking, the channel walls are not continuous but have disconti-
nuities where the boards are butted up to each other, which form small steps or gaps
that increase the measured pressure drop and therefore increase the calculated pres-
sure loss coefficient (see also “Discontinuities and instantaneous flow”). Most ex-
perimental values for the inner stack friction thus include the additional resistance
by gaps ζg. The few available friction data are slightly above the values obtained by
equation (13.35), which can be fitted by using an equivalent roughness of Kequ=1.5K.

13.5.2 Airflow in the kiln

After discussing the flow situation between the board rows in the previous section
13.5.1, the focus will be now on the kiln-wide flow distribution. As pointed out at
the start of the chapter, the flow features in a kiln are mainly influenced by fans,
bends, baffles, and the stack, which may consist of several packs. This analysis will
be based on a standard overhead configuration, as depicted in Figure 13.30. The
drawing shows a two-dimensional cross section of a kiln on the plane where the
main flow occurs. This model is sufficient to explain most of the flow features, since
the length perpendicular to the view extends normally to several metres and the
geometry changes across it are only minor. However, some remarks on three-
dimensional phenomena will be included at the end of this section. The following
study will be based on existing rules for dividing and combining manifolds.

a) Branch loss ratio of manifolds

Consider the dividing and combining manifolds given in Figure 13.31a. According
to Miller (1990), the uniformity of the flow distribution through the branches is
related to the Branch Loss Ratio ηbranch, defined by:
η branch = p v ∆p branches . (13.36)
Airflow within kilns 319

ceiling part

b Vɺ Fan
a ceiling
level

stack
height
wood stack part Hstack
plenum
chamber

stack width L stack


Figure 13.30. Geometry of an overhead-fan wood drying kiln

Here, the velocity pressure pv is calculated from the inlet manifold velocity:

p v = 1 2 ρ (v manifold )
2
. (13.37)

An alternative definition for the loss ratio relates the total cross-sectional areas of
branches and manifold:

Abranches
ηbranch , A = (13.38)
Amanifold .

In order to explain the flow distribution due to inertial effects, we assume the same
total pressure at each branch end and small frictional losses in the manifold
compared to those in the branches. As the flow in a short dividing manifold of
constant area feeds the branches, the velocity in the manifold decreases and the
static pressure rises (see equation (13.3)). This results in increasing branch velocities
towards the end of the manifold. In a short combining manifold, the velocity
increases as more fluid is added from the branches and therefore the static pressure
decreases. Hence, the velocity is highest in the branch closest to the manifold exit.
These characteristics are especially relevant when the branch loss ratio exceeds
unity, as shown in Figure 13.31a. For ηbranch<0.5 the distribution is uniform. If the
frictional losses in the manifold increase, e.g. for smaller cross-sections or
increasing length, they will dominate over the inertial effects and the velocity profile
will have its highest values close to manifold inlet and exit for dividing and
combining manifold respectively.
320 Fundamentals of wood drying

dividing manifolds combining manifolds

vmanifold

dividing combining
vbranch manifold
manifold

branches
short short medium short short medium
a) η < 0.5 η > 1.0 η < 0.5 η > 1.0 b)

Figure 13.31. Flow distribution in single manifolds of short and medium length
according to Miller (1990) (a) and configuration of linked manifolds in a kiln (b)

b) Airflow uniformity along the wood stack height

The theory describing flow uniformity in the branches of single manifolds can be
applied to wood drying kilns. Two characteristic parts will be distinguished, which
are labelled in Figure 13.30:
- Wood stack: The plenum chamber can be considered as the manifold, while the
stack's fillet spaces represent the branches of the manifold.
- Ceiling: The space above the inner ceiling can be considered to be a manifold,
while the entrance to the plenum chamber represents a branch.

Wood stack

The sketch in Figure 13.31b illustrates the flow configuration for a timber stack,
where a dividing manifold at the stack inlet is linked by sticker spaces (branches) to
a combining manifold at the stack outlet. The inertial and frictional effects in both
plenums now interact to influence the flow distribution across the branches, i.e. the
velocity profile over the height of the stack. For symmetric plenums, the inertial
effects on each side of the stack compensate each other, and hence the frictional
effects in the plenum chambers determine the flow distribution within the stack.8
The resulting velocity profile has its highest value in the top part of the stack as
displayed in Figure 13.31b.9 If the plenums have different sizes, the smaller plenum
decides whether the distribution will be that of a dividing or a combining manifold,
e.g. a smaller inlet plenum leads to the profile of a dividing manifold with higher

8
This effect of counterbalanced inertial forces also explains why the application of progressively
narrowing plenum chambers proves to be counterproductive for improving the uniformity of flow
across the stack height (see experiments in Nijdam and Keey 2002).
9
This distribution is also present in the plots of Figure 13.17 and Figure 13.33. It was experimentally and
theoretically obtained in Nijdam and Keey (1999) and Ledig et al. (2001).
Airflow within kilns 321

velocities at the bottom of the stack. This effect was clearly visible in the
experiments carried out by Arnaud et al. (1991) in a kiln section of an overhead fan
kiln.

As in the case of single manifolds, the branch loss ratio given by (13.36) and (13.38)
can also be used to assess the uniformity of flow in the present case of linked
manifolds. In fact, equation (13.38) together with the rule η < 1 is widely used in
practice by kiln designers. The dynamic pressure pv results from equation (13.37).
The pressure drop ∆pbranches is determined by the pressure drop over the stack width.
This value can be calculated using the formulas and graphs given in section 13.5.1b)
or other empirical equations from handbooks, e.g. Idelchik (1996) or VDI… (1994).
The total pressure drop over the stack is given by (13.27) and (13.28). The values for
inlet, outlet and friction loss coefficients can be calculated from equations (13.30)
and (13.31) respectively. Friction factors are obtained from (13.32)-(13.35). After
substituting these definitions into equation (13.36), the branch loss ratio for both
inlet and outlet manifolds becomes:

2
 hH stack   L 
ηstack =    ζ io + λ stack  (13.39)
 (h + t )a   2h  ,

in which the following mass balance is used:

va a = vh hNboards = vh h H stack (h + t ) . (13.40)

The formulation of the loss ratio using cross-sectional areas (13.38) leads to:

hH stack
η stack , A = (13.41)
(h + t )a .

According to Figure 13.31a, the flow through the stack should be uniform along the
stack height if ηstack<0.5. This agrees with results by Nijdam, whose experimental
and numerical results suggest maximum values of ηstack,A = 1 to avoid severe non-
uniformities in the velocity distribution (Nijdam and Keey 1999). The same had
been found in experiments by Arnaud and his co-workers (Arnaud et al. 1991).
Figure 13.32 shows these workers results in a plot of a velocity distribution
coefficient as a function of ηstack,A. Clearly, the velocity distribution becomes more
uniform as the branch loss ratio decreases. In comparison to the cases with blockage
ratios of t/h = 1, the velocity distribution for the same value of ηstack,A improves
further if the blockage ratio increases. This is explained with the help of equation
(13.39), since a higher blockage increases the value of the inlet/outlet pressure loss
coefficient ζio in the denominator and thus lowers ηstack.

Though the flow analysis has focused on a single stack configuration (Figure
13.31b), the results can be also applied to multi-track kilns. The loss ratio according
322 Fundamentals of wood drying

to (13.39) then implies that, for 0,35


t/h=1
kilns with a larger stack width L, a/b doubled
0,30
the plenum size can be narrower t/h=1.4
because of the higher pressure 0,25 a/b doubled

distribution coefficient d
drop through the sticker spaces.
0,20
Ceiling space
0,15

As shown in the flow model of


0,10
Figure 13.31b, it has so far been
assumed that the flow at the 0,05
ceiling level is uniform and
directed downwards. However, if 0,00
no flow guides are installed, the 0,0 1,0 2,0 3,0 η 4,0
stack ,A
flow through the right-angled
bend into the plenum will form a Figure 13.32. Flow distribution coefficient of
separation zone on the inside wall the velocity profile across the height of a
of the bend exit, as shown in timber stack as a function of ηstack,A for
Figure 13.2. The larger the blockage ratios t/h = 1, 1.4 (data from
horizontal velocity component at Arnaud et al. 1991)
the ceiling level, the more
extensive the swirl in front of the top part of the wood stack, which will block the
flow into the uppermost sticker spaces and reduce the uniformity of flow
considerably. To avoid this source of flow maldistribution, the ratio between the
plenum width and the distance between the inner ceiling and the top of the kiln
should obey a special design criterion. Considering the ceiling space to be a
manifold with a single branch, the branch loss ratio according to (13.38) is:

ηceiling , A = a b . (13.42)

If ηceiling,A < 0.5, the flow at the ceiling level should be uniform along the width of
the plenum chamber. The implication of this rule is that the horizontal flow near the
plenum entrance is sufficiently low in comparison to the vertical downwards flow
component. The results in Figure 13.32 support this conclusion, where for same
values of ηstack , the distribution worsens if the value of ηceiling,A is doubled. The flow
through a bend can be guided by contoured walls, which is commonly applied in
some kiln types. Nijdam and Keey (2002) measured the influence of semicircular
sections at the inside wall of the bend on the formation of a vortex in front of the top
fillet spaces (see also Figure 13.15). He found that ηceiling,A < 1 improves the flow
uniformity into the plenum entrance and reduces the size of the separation vortex,
when the right-angled bend is streamlined.
Airflow within kilns 323

Summary and additional remarks

This study on kiln-wide flow features has demonstrated that the uniformity of air
flow over the height of the wood stack is affected by the following branch loss
ratios:
- ηstack or ηstack,A , i.e. either a ratio comparing velocity pressure at plenum cham-
ber inlet to stack flow resistance properties (width, board and sticker dimen-
sions, surface roughness of the boards), or a ratio relating the total cross-
sectional areas of plenum and fillet spaces,
- ηceiling,A , i.e. a ratio relating the ceiling space, where the fans are situated, to the
width of the plenum chamber.

If ηstack> 1, the airflow distribution within the stack will be non-uniform. In this
case, the velocity profile across the stack height will have its maximum in the upper
part of the stack. If ηstack< 0.5, the pressure losses in the plenum are small compared
with the pressure drop across the wood stack ensuring a good flow uniformity across
the height of the stack. However, this requires that the vertical flow distribution over
the plenum chamber width at the ceiling level does not contain a large horizontal
velocity component. Otherwise, a swirl in front of the top of the stack can cause low
air velocities through the upper sticker channels. This vertical flow distribution at
the ceiling level can be ensured by either choosing the kiln dimensions such that
ηceiling,A < 0.5, or by using flow guides and baffles to reduce the horizontal
component.

Other sources of flow maldistribution that have not been addressed yet are three-
dimensional effects in the kiln. They are responsible for flow non-uniformities along
the length perpendicular to the main flow plane, i.e. along the board length. As
shown in the experimental results for the two-dimensional flow fields at the air-inlet
and outlet faces of a board-timber stack in a semi-scale kiln (Ledig 2001), the
vertical velocity distributions at the ends of the lengthwise stack are influenced by
end-wall and airflow bypass effects. Even flow straightening tube banks could not
eliminate separation zones appearing close to the top-ends of the lengthwise stack.
In these regions, the flow is considerably reduced by end-wall effects leading also to
possible secondary flow, as depicted in Figure 13.3. Furthermore, the velocity
profile at the inlet face contained peaks corresponding to the location of the fans.
Finally, a comparison of the inlet and outlet flow fields indicated redistribution of
the airflow within each fillet space on its way through the stack. Hence, conducting
airflow measurements at the outlet face of the stack can give a misleading picture of
flow uniformity.

13.5.3 Airflow design using engineering rules

The airflow through a wood stack should be as uniform as possible (along the stack
height, length and width, see Figure 13.30) in order to obtain the best drying
performance for the kiln. Non-uniform air velocity will not only extend drying times
324 Fundamentals of wood drying

Figure 13.33. CFD-


calculation of the air
velocity distribution in
a kiln designed with
rules for loss ratios

and costs, but can also introduce static instabilities of the stack and large variations
in the final moisture content throughout the stack, which may result in intolerable
timber degrade and low product quality. In the next subsections, the application of
the derived engineering rules is demonstrated, firstly, for the support of the kiln
design process and, secondly, for the improvement of airflow uniformity in existing
kilns.

a) Airflow and kiln design

For demonstration purposes, we consider an example case where the design will be
based on the following requirement specifications:
- Board thickness t: 50 mm.
- Sticker height h: 22 mm.
- Air velocity between the boards vh: 7.0 m/s.10
- Stack height Hstack: 5 m.
- Stack width L: 4.5 m.

To avoid a non-uniform airflow through the stack, the branch loss ratios for the
ceiling space and the wood stack should not exceed 1.0 and be preferably lower than
0.5. The ratios for this design have been chosen slightly above the lower threshold as
ηstack,A = 0.6 and ηceiling,A = 0.6. Applying the engineering rules according to
equations (13.39)-(13.42) results in a plenum chamber width a of 0.95 m and a kiln
ceiling space b of 1.5 m. Instead of building the kiln and verifying the air velocity

10
The investigated type of dryer is a high-temperature kiln, for which such high air velocities in the
sticker space are required.
Airflow within kilns 325

uniformity with measurements, a computer simulation using a CFD model was


performed for this particular kiln design. The calculated absolute velocity
distribution is shown in Figure 13.33. The CFD model predicts a high uniformity of
88% for the air velocity along the height of the stack (mean velocity between the
boards 7.0 m/s, standard deviation 0.8 m/s). This result illustrates that the
application of the design rules can be very useful to achieve a good airflow
distribution within a timber kiln.

b) Improving airflow in existing kilns

Important criteria for a uniform air velocity in the timber stack of an overhead-fan
kiln are:

1. The width of the plenum chamber on both sides of the stack. If the plenum
chamber is too narrow, the frictional effects will generate a relatively high pres-
sure drop down the stack height. As a result, the air velocity in the lower part of
the stack will be significantly lower than the mean velocity.

2. The velocity distribution at inner ceiling level. If the air velocity at the ceiling
level has a significant horizontal component, it is likely that swirl will occur in
front of the upper part of the stack. This separation zone causes a significantly
lower velocity in the uppermost part of the stack. The swirl can be avoided by
reducing the horizontal component in the velocity at ceiling level. This can be
achieved in two ways: (a) by enlarging the distance between the ceiling and the
kiln top wall and (b) by guiding the airflow through the bend with the help of
contoured flow guides or baffles.

In addition to the design of the kiln geometry, the stacking strategy is of high
importance. The following guidelines should be taken into consideration:

1. Bypassing air. All gaps between the stack and the kiln walls (and ceiling) should
be blocked to achieve a maximal flow through the stack. Bypassing through the
open spaces between individual packages should not be neglected.

2. Placement of the packages with respect to the kiln ceiling. Placement of the
packages should be such that the outer stack edge is aligned with the edge of the
inner ceiling wall. If the stack protrudes from the stacking space, the plenum is
narrowed leading to the discussed results. If the load is placed in a way that
leaves a wider plenum, the formation of a separation zone in front of the stack
below the inner ceiling will retard the flow in the upper stack's fillet spaces.

3. Placement of the packages with respect to each other. The positioning of packs
in a staggered manner will create locally large open spaces below the kiln’s inte-
rior ceiling, and therefore large areas of re-circulating flow, which result in poor
airflow uniformity.
326 Fundamentals of wood drying

13.6 Influence of airflow on wood moisture distribution

Thus far, the airflow analysis has concentrated on various features within a kiln that
affect the uniformity of airflow across the stack, which in turn affects the even
drying of timber. However, in the drying process, the air is cooled down as it carries
away the evaporated moisture from the wood surface, an effect which also
influences the uniform drying of the timber boards. On its way through the stack, the
drying potential of the air decreases. Even in the case of a perfect air flow
distribution across the stack, this decreasing drying potential leads to uneven drying
across the stack width. Airflow reversals can reduce but not eliminate this effect.
The next sections will address these phenomena and consider the impact of non-
uniform external drying conditions on local drying rates in the stack.

13.6.1 Airflow induced drying non-uniformities

There is a direct correlation between the magnitude of the air velocity in the stack
and the magnitude of the heat- and mass transfer coefficients. This correlation is
often described by so-called Nusselt relations. Examples of these Nusselt relations
are covered in the chapter “External heat and mass transfer”. As a consequence,
local variations in the airflow will have an influence on the drying rate and on the
spread in final moisture content of the boards. In general, the correlation between
the heat transfer coefficient and the air velocity is of the form (see also (13.45)):

hh = C v a , where a = 0.6 ... 0.8 . (13.43)

This relation shows that a variation in air velocity between 2 and 4 m/s with an
average of 3m/s will cause a variation in the heat transfer coefficient of
approximately 25% with respect to the average value in the stack. This will have a
significant impact on the drying rate, especially in the first drying period, where the
moisture content is above fibre saturation point.

To illustrate the effect of this variation in the heat transfer coefficient on the drying
rate, results of a simulation for a drying process of Sitka Spruce are presented in
Figure 13.34. The simulation was performed with TNO’s timber drying model
“Woody”. The moisture content evolution is shown for three different air velocities:
2, 3 and 4 m/s. The target moisture content was 20%. Starting from the same initial
moisture content, the boards would achieve this target value after 100 hours, if the
air velocity was 4 m/s. With an air velocity of 3 m/s, the drying process would take
12 hours longer, while the boards with an air velocity of 2 m/s would dry the slowest
and only achieve a moisture content value of 25%. For these boards the drying time
should be extended by at least 24 hours. If the target moisture content of the boards
is lower, e.g. 12%, the influence of the air velocity on the total drying time will be
less significant. Nonetheless, a variation in air velocity will create substantial
differences in the moisture content at specific locations in the timber stack.
Airflow within kilns 327

100

90 Tdry [C]
Twet [C]
80 MC (Vair = 2 m/s)
MC (Vair = 3 m/s)
[°C]

70
Temperature [°

MC (Vair = 4 m/s)
[%],Temperature

60

50

40
MC[%],
MC

30 Vair = 2 m/s

20 Vair = 4 m/s Vair = 3 m/s

10

0
0 10 20 30 40 50 60 70 80 90 100 110 120 130
Time [hours]
Figure 13.34. Variation in moisture content as a function of local air velocity in the
stack, drying simulation of Sitka Spruce from green (MC=75%) to target MC= 20%
with the timber drying model “Woody”

The drying conditions (temperature, relative humidity) are normally based on the
average moisture content of a kiln load. In those parts of the stack where the air
velocity is high, the drying conditions based on the drying schedule will be often too
harsh, which may result in timber quality degrade. As rule of thumb, the maximum
variation in air velocity should be within 15% of the average air velocity value.

13.6.2 Temperature drop across the stack during the first drying period

The following paragraphs give an overview of some theoretical aspects related to


drying rate uniformity along the width of the stack (see Figure 13.30). The drying
rate along the width of the stack decreases in the flow direction. During the drying
process, the energy transfer from the hot air to the wet wood surface will result in
moisture evaporation, and a subsequent cooling of the airflow. As a consequence,
the drying potential of the airflow decreases towards the exit of the stack. In the
worst case, the air leaves the stack with a temperature equal to the wood
temperature, reducing the drying rate at the exit of the stack to zero. The drop in air
temperature is significant especially during the first drying period (moisture content
above fibre saturation point), when the drying rate is high, and will reduce the
effective drying rate considerably along the stack width.

a) One-dimensional heat transfer

The temperature drop over the stack width can be calculated using the theory for
heat transfer by forced convection in a channel, assuming steady flow and a one-
328 Fundamentals of wood drying

TW
Ta,i
Ta
v dx
TW
x

Figure 13.35. One-dimensional heat transfer by forced convection in a channel

dimensional approach. This theory can be found in numerous textbooks, e.g.


Incropera and DeWitt (1996) and will therefore be presented here only as a short
summary. The energy balance for a differential section of a channel of length dx is
given by (see Figure 13.35):

dTa
Aρ a v c p , a = hhU (TW − Ta ) , Ta x =0
= Ta ,i . (13.44)
dx

TW and ρ are assumed to be constant over the entire length of the channel. This is a
good assumption during the first stage of drying, once the wood has been warmed
up to its initial steady value. The wall heat transfer coefficient hh in a fully
developed duct flow is calculated using (Incropera and DeWitt 1996):

λa
h h = 0.027 Re 0.80 Pr 0.33 . (13.45)
Dh

A dimensionless air temperature can be derived from equation (13.44) and (13.45),
as follows:

~ T − TW
Ta ≡ a = e −Cx (0< T~a <1, TW < Ta <Ta,i) , (13.46)
Ta ,i − TW

where

hhU 2 0.027
C= = ⋅ 0.2 0.67 . (13.47)
mɺ a c p , a h Re Pr

This dimensionless temperature difference represents the ratio of the local heat
transfer potential, or temperature driving force, to its maximum value at the inlet.
The product C·L stands for the number of heat transfer units with 1/C as the unit’s
length. Figure 13.36 shows the air temperature potential along the stack width for
different values of C·L, where L represents the width of the stack. Hence, the air
Airflow within kilns 329

~ 1
Ta
0.9

0.8

0.7

0.6

0.5

0.4

0.3

C×L=0.50
0.2
C×L=1.00
0.1 C×L=2.50
C×L=5.00
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L

Figure 13.36. Dimensionless mean temperature difference drop along the stack
width without flow reversal

entering the stack (at x=0) has a temperature of Ta,i and the air leaving the stack (at
x=L) has:
- almost reached the wet bulb temperature TW , if C·L>5.0 (no drying capacity
remains at the stack exit),
- only slightly deviated from its initial dry bulb temperature Ta,i , if C·L<0.5 (sub-
stantial drying capacity remains at the stack exit).

With a sticker dimension of h = 22 mm, the value of the constant C is approximately


0.5 m-1 in most kilns. In those cases, a value C·L≈0.5 represents a stack width of
L≈1.0 m, where C·L≈5.0 represents a stack width of L≈10 m. When flow reversals
are applied, the dimensionless mean air temperature difference is determined by:
~
Ta ,reversal = 0.5(e − Cx + e − C ( L − x ) ) , (13.48)

which is plotted in Figure 13.37.

b) Drying rate uniformity across the width of the stack

The drying rate is proportional to the heat flux to the wood assuming that the wood
temperature is constant or changing very slowly during the main part of a drying
schedule phase. This (mainly convective) heat flux to the wood surface can be
expressed as:
~
q = hh (Ta − TW ) = hh (Ta ,i − TW )T (C , x ) , (13.49)
330 Fundamentals of wood drying

1
~
Ta
0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2 C×L=0.50
C×L=1.00
0.1 C×L=2.50
C×L=5.00
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x/L

Figure 13.37. Dimensionless mean temperature difference drop along the stack
width using flow reversal

~
where T can be calculated from equation (13.46) or (13.48). The following
relationship is introduced to quantify the uniformity in drying rate:

max drying − rate − min drying − rate


uniformity = 1 − (13.50)
max drying − rate .

It follows that the drying rate uniformity is 1 if there is no difference in drying rate
within the stack along the airflow direction, and that it approaches zero if the
minimum drying rate is very small compared with the maximum value. Combining
(13.46) and (13.48), the uniformity can then be expressed as:

uniformityno reversal = e − CL , (13.51)


1
− CL
2e 2
(13.52)
uniformity reversal =
1 + e −CL .

Graphs showing the uniformity in drying rate along the width of the stack calculated
using (13.52) for the case with and (13.51) for the case without flow reversal are
given in Figure 13.38. The curves are drawn for cases with C=0.5 m-1. The figure
clearly illustrates that flow reversals improve the uniformity of the drying rate across
the width of the stack. For example, a drying rate uniformity of 35% is achieved in a
kiln without flow reversal for a stack width of 2.0 m. With flow reversals, the same
drying rate uniformity can be achieved for a larger, 6.5 m wide stack.
Airflow within kilns 331

100%

90% no flow reversal

80% with flow reversal

70%

60%
uniformity

50%

40%

30%

20%

10%

0%
0 1 2 3 4 5 6 7 8 9 10
L stack [m]

Figure 13.38. Comparison between the relative drying rate uniformity along the
stack width without and with flow reversal (C=0.5 m-1)

c) Concluding remarks

The drying rate uniformity along the width of the stack (in the airflow direction), as
defined by (13.50), is determined by the product C·L, which implies that the
uniformity is proportional to the sticker height h and inversely proportional to the
stack width L. There is only a weak dependence of the product C·L on the Reynolds
number, and hence the air velocity is of minor importance for the drying rate
uniformity along a sticker channel. Once the product C·L is known, the uniformity
can be calculated using equation (13.51) and (13.52).

During the first drying period, the wood temperature in the stack is rather uniform
and close to the wet bulb temperature. As explained in the previous sections, the
local drying rate will vary considerably in the stack during this drying phase. If no
flow reversal is applied, boards at the entrance side will dry with the maximal rate,
while the boards located at the exit of the stack will experience the minimum
temperature driving force and accordingly dry more slowly. If flow reversals are
applied with a frequency high in comparison to the first phase drying time, then the
minimum drying rate occurs in the middle of the stack. As shown in studies by
Nijdam and Keey (1996), in practice only a few early flow reversals proved to be
sufficient to improve uniformity in drying, and a further increase in flow reversal
number did not produce a better drying uniformity.
332 Fundamentals of wood drying

13.7 Acknowledgements

The authors wish to acknowledge the contribution of Dr. Justin Nijdam through his
review of the chapter. We would also like to thank Prof. K.-E. Militzer and several
colleagues at Dresden University for proofreading the manuscript.

The following material was reprinted by permission of the publishers:

1. From: Wood Science and Technology


(published by Springer-Verlag Heidelberg)

Figure 4 of NIJDAM, J.J. and KEEY, R.B., 2002, vol 36 (1), pp. 19-26

2. From: NITSCHE, W., 1994 - Stroemungsmesstechnik. Berlin, Heidelberg:


Springer-Verlag

Figure 3.23 p. 61
Figure 3.24 p. 62

3. From: JOERGENSEN, F.E. , 2002 - How to measure turbulence with hot-wire


anemometers - a practical guide. Publication no.: 9040U6151. Skovlunde,
Denmark: Dantec Dynamics A/S

Figure 1 p. 6
14 References
Adler P.M. 1992 – Porous media: geometry and transports. Butterworth-Heinemann Series in
Chemical Engineering.
Agoua E. 2001 – Diffusivité et perméabilité du bois: validation de méthodologies
expérimentales et prise en compte de paramètres morphologiques simples pour la
modélisation physique. PhD Thesis, ENGREF, Nancy.
Aguiar O., Perré P. 2000a – The “flying wood” test used to study the variability of drying
behaviour of oak. Proceedings of 2nd Workshop of COST Action E15 “Quality Drying of
Hardwood”, 11-13 September 2000, Sopron, Hungary, 10 pages.
Aguiar O., Perré P. 2000b – Pocesso de secagem acelerada de madeira baseado nas suas
propriedades reológicas, Patent n° 1265, Instituto nacional de proteçao industrial.
Amburgey T., Kitchens S., Baileys J. 2000 – Guidelines for the use of anti-sapstain
formulations. American Wood Preservers’ Association, N4 Task Force, p. 5.
Anderson N.T., McCarthy J.L. 1963 – Two – parameter isotherm equation for fiber – water
systems. Eng.Chem.Process Design Dev. 2, 103–105.
Anon. 1987 – Wood handbook. US Government Printing Office 0-495-044.
Anon. 2002 – A Dictionary of Biology, Oxford University Press, Market House Books Ltd
2000 from: http://www.xrefer.com/ (19.08.2002).
Anon. 2002 – University of Wisconsin, Forestry 415: http://treebiol.forest.wisc.edu/
forestry415/Tree Structure/Stems/ (19.08.02).
Anon. 2002 – Woodzone.com: http://www.woodzone.com/articles/reaction_wood.htm
(20.08.2002).
Anon. 2002 – http://www.kwic.com/~pagodavista/schoolhouse/land/pics/trunkcrs.jpg
(22.08.2002).
Anon. 2002 – NZFRI Forest Research, NZ - www.forestresearch.co.nz/images/photos/
af_sem_compwood.gif (19.08.2002).
Antti L., Perré P. 1999 – A microwave applicator for on line wood drying: temperature and
moisture distribution in wood. Wood Science and Technology 33: 123-138.
Antti L., Zhao H., Turner I. 1999 – An investigation of the heating of wood in an industrial
microwave applicator: theory and practice. Proceedings of 6th International IUFRO Wood
Drying Conference, 25-28 January 1999, Stellenbosch, South Africa, 11-17.
Arnaud G., Fohr J.-P., Garnier J.-P., Ricolleau C. 1991 – Study of the air flow in a wood
drier. Drying Technology 9(1), 183-200.
Arrighetti G. 2002 – http://www.itl.tn.cnr.it/Web/labimil/anatomcar.html (12.08.2002).
Ashworth J.C. 1977 – The mathematical simulation of batch-drying of softwood timber. PhD
Thesis, University of Canterbury, New Zealand.
Asikainen A. 2002 – Chemistry and discoloration of birch wood. In: Paavilainen, L. (ed.).
Finnish Forest Cluster Research Programme WOOD WISDOM (1998-2001). Final report,
196-206.
Atkins P.W. 1994 – Physical chemistry. 5th Edition, Oxford University Press.
334 Fundamentals of wood drying

Aufseß H. v. 1986 – Lagerverhalten von Stammholz aus gesunden und erkrankten Kiefern,
Fichten und Buchen. Holz als Roh- und Werkstoff 44: 325.
Avramidis S. 1989 – Evaluation of “three – variable” models for the prediction of equilibrium
moisture content in wood. Wood Science and Technology 23: 329–333.
Avramidis S. 1999 – Radio-frequency vacuum drying of wood. Proceedings of the First
COST Action E15 Wood Drying Workshop, 14 pages, Edinburgh, Scotland, 14 pages.
Avramidis S., Ellis S., Liu J. 1993 – The alleviation of brown stain in hem-fir through
manipulation of kiln-drying schedules. Forest Products Journal 43, 65-69.
Avramidis S., Englezos P., Papathanasiou T. 1992 – Dynamic nonisothermal transport in
hygroscopic porous media: moisture diffusion in wood. American Institute of Chemical
Engineers Journal 38, 1279-1287.
Avramidis S., Hatzikiriakos S.G., Siau J.F. 1994 – An irreversible thermodynamics model for
unsteady-state nonisothermal moisture diffusion in wood. Wood Science and Technology
28, 349-358.
Avramidis S., Hatzikiriakos S.G. 1995 – Convective heat and mass transfer in nonisothermal
moisture desorption. Holzforschung 47, 163-167.
Avramidis S., Kuroda N., Siau J.F. 1987 – Experiments in nonisothermal diffusion of
moisture in wood. Part II. Wood Science and Technology 21, 249-256.
Avramidis S., Siau J.F. 1987a – Experiments in nonisothermal diffusion of moisture in wood.
Part 3. Wood Science and Technology 21, 329-334.
Avramidis S., Siau J.F. 1987b – An investigation of the external and internal resistance to
moisture diffusion in wood. Wood Science and Technology 21, 249-256.
Avramidis S., Zwick R. 1996 – Commercial-scale RF/V drying of softwood lumber. Part 2.
Drying characteristics and lumber quality. Forest Products Journal 46, 27-36.
Babbitt J.D. 1940 – Observations of the permeability of hygroscopic materials to water
vapour. I. Observations at relative humidities less than 75%. Canadian Journal of
Research 18A, 113-120.
Banks W.B. 1968 – A technique for measuring the lateral permeability of wood, J. Inst. Wood
Science 4, 35-41.
Banks W.B. 1980 – The effect of cell wall swelling on the permeability of grand fir wood.
Wood Science and Technology 14, 49-62.
Banks W.B. 1981 – Addressing the problem of non-steady state liquid flow in wood. Wood
Science and Technology 15, 171-177.
Barkas W.W. 1942 – Wood water relationships – VII. Swelling pressure and sorption
hysteresis in gels. Transactions of the Faraday Society 38, 194-209.
Barrer R.M., Barrie J.A. 1958 – Sorption and diffusion in ethyl cellulose. Part IV. Water in
ethyl cellulose. Journal of Polymer Science 28, 377-386.
Barton G.M., Gardner J.A.F. 1966 – Brown stain formation and the phenolic extractives of
Western hemlock (Tsuga heterophylla (Raf.) Sarg.). Department of Forestry Publication
No. 1147. Roger Duhamel, Ottawa, 7-20.
Basilico C., Martin M. 1984 – Approche expérimentale des mécanismes de transfert au cours
du séchage convectif à haute température d'un bois résineux, Int. J. Heat Mass Transfer
27, 657-668.
Bauch J. 1984 – Discoloration in the wood of living and cut trees. IAWA Bulletin 5, 92-98.
References 335

Bauch J. 1986 – Verfärbungen von Rund- und Schnittholz und Möglichkeiten für
vorbeugende Schutzmaßnahmen. Holz-Zentralblatt 112, 2217-2218.
Bauch J., Schmidt O., Yazaki Y., Starck M. 1985 – Significance of bacteria for the
discoloration of Ilomba wood (Pycnanthus angolensis Exell.). Holzforschung 39,
249-252.
Bauch J., Hundt H.V., Weißmann G., Lange W., Kubel H. 1991 – On the causes of yellow
discolorations of oak heartwood (Quercus Sect. Robur) during drying. Holzforschung 45,
79-85.
Blasius H. 1908 – Grenzschichten in Flüssigkeiten mit kleiner Reibung. Z. Math. Physik 56
(1), pp. 1-37. English translation in NACA Techn. Mem. Nr. 1256.
Bois P. 1970 – Grey-brown chemical stain in southern hardwoods. USDA FS, FPU Technical
Report No 1, p. 4.
Bazant Z.P. 1985 – Constitutive equation of wood at variable humidity and temperature,
Wood Science and Technology 19, 159–177.
Bergfeld A., Bergmann R., von Sengbusch P. 2002 – Botany online - The Internet
Hypertextbook: http://www.biologie.uni-hamburg.de/b-online/ (12.08.02).
Bird R.B., Stewart W.E., Lightfoot E.N. 1960 – Transport phenomena. John Wiley & Sons
Inc. New York.
Bolton A.J., Petty J.A. 1977 – Influence of critical point and solvent exchange drying on the
gas permeability of conifer sapwood. Wood Science 9, 187-193.
Bolton A.J., Petty J.A. 1978 – The relationship between the axial permeability of wood to dry
air and to a non polar solvent. Wood Science and Technology 12, 111-126.
Bosshard H.H. 1956 – Über die Anisotropie der Holzschwindung. Holz als Roh- und
Werkstoff 14, 285-295.
Bramhall G. 1971 – The validity of Darcy’s law in the axial penetration of wood. Wood
Science and Technology 5, 121-134.
Bramhall G. 1976 – Fick’s law and bound water diffusion. Wood Science 8, 153-161.
Bramhall G. 1979 – Sorption diffusion in wood. Wood Science 12, 3-13.
Brandão A., Jankowsky I. 1990 – A screening to select kiln schedules, Wood drying working
party, IUFRO World Congress, August, Quebec, Canada.
Brandao A., Perré P. 1996 – The “Flying Wood” - A quick test to characterise the drying
behaviour of tropical woods. Proceedings of 5th International IUFRO Wood Drying
Conference, August 13-17, 1996, Québec, Canada, 315-324.
Brauner A.B., Loos W.E. 1968 – Colour changes in black walnut as a function of
temperature, time and two moisture conditions. Forest Products Journal 18: 29-34.
Brauner A., Conway E.M. 1964 – Steaming walnut for colour. Forest Products Journal 14:
525-527.
Brunauer et al. 1967 – The Langmuir and BET theories. The solid - gas interface. Marcel
Dekker Inc., New York, 77–103.
Brunner R. 1999 – Vakuumtrocknung im Wirtschaftlichkeitsvergleich, Teil 3, Holz-
Zentralblatt 125, 874-875.
Bucki M., Perré P. 2003 – Physical formulation and numerical modeling of high frequency
heating of wood. Drying Technology 21, 1151-1172.
336 Fundamentals of wood drying

Burmester A., Wille W.E. 1975 – Untersuchungen zur Formbeständigkeit von Teakholz. Holz
als Roh- und Werkstoff 33, 147-150.
Burtin P., Jay-Allemand C., Charpentier J.-P., Janin G. 2000 – Modifications of hybrid walnut
(Juglans nigra x Juglans regia) wood colour and phenolic composition under various
steaming conditions. Holzforschung 54, 33-38.
Canteri L., Martin M., Perré P. 1997 – A non-destructive method for quantifying the wood
drying quality used to determine the intra and inter species variability. Drying Technology
15, 1293-1325.
Cai L., Avramidis S. 1993 – A study on the separation of diffusion and surface emission
coefficients in wood. Drying Technology 15, 1457-1473.
Carlsson P., Esping B., Dahlblom O. 1996 – Optimization of the wood drying process.
Proceedings of 5th International IUFRO Wood Drying Conference, August 13-17, 1996,
Québec, Canada, 169-176.
Carlsson P., Arfvidsson J. 1999 – Optimized wood drying. Proceedings of 6th International
IUFRO Wood Drying Conference, 25-28 January 1999, Stellenbosch, South Africa,
139-147.
Cassens D.L. 1991 – Sap stain in hardwood logs and lumber. FNR-90 (Rev. 11/91),
Cooperative Extension Service Purdue University, West Lafayette, IN 47907, pp. 10.
Chang S.T., Wang S.Y., Su Y.C., Huang S.H., Kuo Y.H. 1999 – Chemical constituents and
mechanisms of discoloration of Taiwania (Taiwania cryptomerioides Hayata) heartwood.
Holzforschung 53, 142-146.
Charrier B., Haluk J.P., Janin G. 1992 – Prevention of brown discoloration in European
oakwood occurring during kiln drying by a vacuum process: Colorimetric comparative
study with a traditional process. Holz als Roh- und Werkstoff 50: 433-437.
Charrier B., Haluk J.P., Metche M. 1995 – Characterisation of European oakwood constitutes
acting in the brown discoloration during kiln drying. Holzforschung 49, 168-172.
Chen Y., Choong E.T., Wetzel D.M. 1995 – Evaluation of diffusion coefficient and surface
emission coefficient by an optimization technique. Wood and Fiber Science 27, 178-182.
Chen Y., Choong E.T., Barnes H.M. 1995 – Effect of selected water-soluble bulcking
chemicals on moisture diffusion and dimensional stability of wood. Forest Products
Journal 45, 84-90.
Chen Y., Choong E.T., Wetzel D.M. 1996 – A numerical analysis technique to evaluate the
moisture-dependent diffusion coefficient on moisture movement during drying. Wood and
Fiber Science 28, 338-345.
Chen G., Keey R.B., Walker J.C.F. 1997 – The drying stress and check development on high-
temperature kiln seasoning of sapwood Pinus radiata boards. Holz als Roh- und Werkstoff
55, 59-64.
Chen P., Schmidt P.S. 1997 – Mathematical modeling of dielectrically-enhanced drying. In:
Mathematical modeling and numerical techniques in drying technology. (Ed. Turner, I.,
Mujumdar A.S.) Marcel Dekker, New York.
Chen P.Y.S., Workman E. Jr. 1980 – Effect of steaming on some physical and chemical
properties of black walnut heartwood. Wood and Fiber 11, 218-227.
Cho H.H., Jabbari M.Y., Goldstein R.J. 1994 – Mass transfer with flow through an array of
rectangular cylinders. J. Heat Transfer 116, 904-911.
References 337

Choong E.T. 1962 – Movement of moisture through a softwood (Abies sp.) in the hygroscopic
range. Ph.D. Thesis, State University of New York, Syracuse, NY.
Choong E.T. 1963 – Movement of moisture through a softwood in the hygroscopic range.
Forest Products Journal 13, 489-498.
Choong E.T. 1965 – Diffusion coefficient of softwoods by steady-state and theoretical
methods. Forest Products Journal 15, 21-27.
Choong E.T., Kimbler O.K. 1971 – A technique of measuring water flow in woods of low
permeability. Wood Science 4, 32-36.
Choong E.T., Shupe F.T., Chen Y. 1999 – Effect of steaming and hot-water soaking on
extractive distribution and moisture diffusivity in southern pine during drying. Wood and
Fiber Science 31, 143-150.
Choong E.T., Skaar C. 1969 – Separating internal and external resistance to moisture removal
in wood drying. Wood Science 1, 200-202.
Choong E.T., Skaar C. 1972 – Diffusivity and surface emissivity in wood drying. Wood and
Fiber 4, 80-86.
Choong E.T., Tesoro F.O., Manwiller F.G. 1974 – Permeability of twenty-two small diameter
hardwoods growing on southern pine sites, Wood and Fiber 6, 91-101.
Choong, E.T., Tesoro, F.O. 1989 – Relationship of capillary pressure and water saturation in
wood. Wood Science and Technology 23, 139–150.
Christensen G.N. 1960 – Kinetics of sorption of water vapour by wood. I The effect of sample
thickness. Australian Journal of Applied Science 11, 295-304.
Christensen G.N., Kelsey K.E. 1959a – The rate of sorption of water vapour in wood. Holz als
Roh- und Werkstoff 17, 178-188.
Christensen G.N., Kelsey K.E. 1959b – Die Sorption von Wasserdampf durch die chemischen
Bestandteile des Holzes. Holz als Roh- und Werkstoff 17, 189–203.
Churchill S.W., Chan C. 1995 – Theoretically based correlating equations for the local
characteristics of fully turbulent flow in round tubes and between parallel plates. Ind. Eng.
Chem. Res. 34 (4), 1332-1341.
Churchill S.W., Chan C. 1994 – Improved correlating equations for the friction factor for
fully turbulent flow in round tubes and between identical parallel plates, both smooth and
naturally rough. Ind. Eng. Chem. Res. 33(8), 2016-2019.
Churchill S.W. 1973 – Empirical expressions for the shear stress in turbulent flow in
commercial pipe. AIChE Journal 19(2), 375-376.
Claesson J., Arfvidsson J. 1996 – A new method to evaluate flow coefficients from moisture
distributions. Proceedings of 5th International IUFRO Wood Drying Conference, August
13-17, 1996, Québec, Canada, 515-521.
Cloutier A., Fortin Y., 1991 – Moisture content-water potential relationship of wood from
saturated to dry conditions. Wood Science and Technology 25, 263-280.
Cloutier A., Fortin Y. 1993 – A model of moisture movement in wood based on water
potential and the determination of the effective water conductivity. Wood Science and
Technology 27, 95-114.
Colebrook C.F. 1938/39 – Turbulent flow in pipes with particular reference to the transition
between the smooth and rough pipe laws. J. Inst. Civil Eng. 11, 133-156.
338 Fundamentals of wood drying

Comstock G.L. 1963 – Moisture diffusion coefficients in wood as calculated from adsorption,
desorption and steady-state data. Forest Products Journal 13, 97-103.
Comstock G.L. 1967 – Longitudinal permeability of wood to gases and nonswelling liquids,
Forest Products Journal 17, 41-46.
Comstock G.L. 1970 – Directional permeability of softwoods. Wood and Fiber 1, 283-289.
Comstock G.L., Côté W.E. 1968 – Factors affecting permeability and pit aspiration in
coniferous sapwood. Wood Science and Technology 2, 279-291.
Constant T., Perré P., Moyne C. 1996 – Drying with internal heat generation: Theoretical
apects and application to microwave heating. AIChE Journal 42, 359-368.
Coper D.N.E., Ashpole D.K. 1959 – The heats of sorption of cellulosic fibres from
calorimetric measurements. J. Text. Inst. 50, 223–232.
CTBA 1995 – Dictionnaire trilingue des bois rond et des bois scié. CTBA, Paris.
Crank J. 1951 – A theoretical investigation of the influence of molecular relaxation and
internal stress on diffusion in polymers. Journal of Polymer Science 11, 151-168.
Crank J. 1975 – The mathematics of diffusion. Oxford University Press, Oxford.
Crank J., Park G.S. 1968 – Diffusion in polymers. Academic Press, London, New York.
Crews D.L. 1965 – Structural variables and the differential transverse shrinkage of wood.
PhD Thesis, SUNY Coll. For., Syracuse NY.
Cudinov B.S. 1984 – Voda v drevesine. Nauka, Novosibirsk.
Cur N., Sparrow E.M. 1978 – Experiments on heat transfer and pressure drop for a pair of
colinear, interrupted plates aligned with the flow. Int. J. Heat Mass Transfer 21,
1069-1080.
Cur N., Sparrow E.M. 1979 – Measurements of developing and fully developed heat transfer
coefficients along a periodically interrupted surface. J. Heat Transfer 101, 211-216.
Dahlblom O. 1987 – Constitutive modelling and finite element analyse of concrete structures
with regard to environmental influence. Lund Institute of Technology. Division of
Structural Mechanics, Report TVSM.1004. Lund Sweden.
Dahlblom O., Ormarsson S., Petersson H. 1996 – Simulation of wood deformation processes
in drying and other environmental loading. Annals Forest Sci. 53, 857-866.
Dahlblom O., Petersson H., Ormarsson S. 1994 – Numerical simulation of the development of
deformation and stress in wood during drying. Proceedings of 4th IUFRO International
Wood Drying Conference, August 9-13, 1994, Rotorua, New Zealand, 165-180.
Dahlblom O., Peterssson H., Ormarsson S. 2001 – Full 3-D FEM-simulations of drying
distortions in spruce boards based on experimental studies. Proceedings of 7th IUFRO
International Wood Drying Conference, July 9-13, 2001, Tsukuba, Japan, 246-251.
Danckwerts P.V., Anolik C. 1962 – Mass transfer from a grid packing to an air stream. Trans.
Inst. Chem. Engrs. 40, 203-213.
Darakov M. 1977 – Investigation of the hydraulic resistance of stacks of surfaced timber with
horizontal spaces between the boards. Naucni trudove / Serija Mechanicna technologija na
durvesinata 22, 87-97.
Darcy H. 1856 – Les Fontaines Publiques de la Ville de Dijon. Dalmont, Paris. 647 p. & atlas.
Davis J.R., Ilic J., Wells P. 1993 – Moisture content in drying wood using direct scanning
Gamma-ray densitometry. Wood and Fiber Science 25, 153-162.
References 339

Denig J., Wengert E., Simpson W. 2000 – Drying hardwood lumber. USDA F.S., FPL.
General Technical Report. Sept. 2000, FPL –GTR-118: 138.
Dent R.W. 1987 – A multilayer theory for gas sorption. I. Sorption of a single gas. Text. Res.
J. 47, 145–152.
Dietrichs H.H. 1964 – Chemisch-physiologische Untersuchungen über die Splint-Kern-
Umwandlung der Rotbuche (Fagus sylvatica [L.]) - ein Beitrag zur Frage der
Holzverkernung. Mitteilungen der BFH 58, 141.
Dinwoodie J.M. 1981 – Timber. Its nature and behaviour. Van Nostrand Reinhold Company,
New York.
Dullien F.A.L. 1992 – Porous media - Fluid transport and pore structure. Second Edition,
Academic Press, London.
Durst F., Melling A., Whitelaw J.H. 1981 – Principles and practice of Laser-Doppler
anemometry. London: Academic Press.
Dushman S., Lafferty J.M. 1962 – Scientific foundations of vacuum technique. John Wiley &
Sons Inc. New York.
Eckert E.R.G., Drake R.M. 1981 – Heat and mass transfer. Robert E. Krieger Publishing
Company. Malabar, FL.
Esfahanian A., Dost S., Tabarrok B. 1999 – A numerical study of fluid flow and heat/mass
transfer over a stack of planks. Drying Technology 17, 1495-1510.
Esfahanian A. 2000 – A numerical modelling study of transport phenomena in wood drying.
PhD Thesis. Victoria (Canada): University of Victoria.
Fengel D., Wegener G. 1989 – Wood. Chemistry, ultrastructure, reactions. Berlin, New York:
Walter de Gruyter.
Feist W., Hon D. 1984 – Chemistry of weathering and protection. In: The chemistry of solid
wood. American Chemical Society, 401-451.
Ferziger J.H., Peric M. 1997 – Computational methods for fluid dynamics. 2. Ed. Springer
Berlin, Heidelberg, New York.
Fluent Inc. 1998 – FLUENT 5 User’s Guide. Lebanon, NH.
Fortin Y., Tremblay C., Fafard M., Duplain G. 2001 – Wood drying modelling based on the
water potential concept: simulation results for conventional kiln drying of black spruce
lumber. Proceedings of 7th IUFRO International Wood Drying Conference, July 9-13,
2001, Tsukuba, Japan, 28-33.
Fortuin G., Welling J., Hesse Ch., Brückner G. 1988a – Verfärbung von Eichenschnittholz bei
der Trocknung (1). Holz-Zentralblatt 114: 1606-1608.
Fortuin G., Welling J., Hesse Ch., Brückner G. 1988b – Verfärbung von Eichenschnittholz bei
der Trocknung (2). Holz-Zentralblatt 114: 1621-1622.
Fox D., Mortimer L.M., Weirsberger A. (Eds.) 1965 – Chemistry of the organic solid state.
Interscience Publications, Chapter 6, 509-627.
Frey-Wyssling A. 1940a – Die Anisotropie des Schwindmasses auf dem Holzquerschnitt.
Holz als Roh- und Werkstoff 3, 43-45.
Frey-Wyssling A. 1940b – Die Ursachen der anisotropen Schwindung des Holzes. Holz als
Roh- und Werkstoff 3, 349-353.
Fujita H., Kishimoto A. 1958 – Diffusion-controlled stress relaxation in polymers. II. Stress
relaxation in swelling polymers. Journal of Polymer Science 28, 547-567.
340 Fundamentals of wood drying

Gerbery J. 1966 – Aerodynamic resistances in lumber stacks. Drevo 21(10), 351-353, 356.
Gillwald W., Tschirnich J. 1967 – The effect on kiln drying of the air-layer close to the
boards. Holztechnologie 81, 29-35.
Gjerdrum P. 1997 – Distortions in industrially kiln dried timber. Norske Skog Teknikk,
Lysaker, Report 31 (in Norwegian, English summary).
Gjerdrum P. 2000a – Shrinkage in industrially kiln dried timber. Treteknisk Informasjon,
Oslo 4/2000, 27-28 (in Norwegian).
Gjerdrum P. 2000b. – Cost efficient timber drying. Proceedings of 2nd Workshop of COST
Action E15 “Quality Drying of Hardwood”, 11-13 September 2000, Sopron, 5 pages.
Goldstein R.J., Yoo S.Y., Chung M.K. 1990 – Convective mass transfer from a square
cylinder and its base plate. Int. J. Heat Mass Transfer 33, 9-18.
Grace C. 1996 – Drying characteristics of heartwood. ME (Chen) Thesis, University of
Canterbury, NZ.
Greenhill W.L. 1937 – The design of the circulating system of commercial timber seasoning
kilns. J. Counc. Sc. Ind. Res. (Aust.) 10, 131-135.
Grosser D. 1977 – Die Hölzer Mitteleuropas. Springer Berlin, Heidelberg, New York.
Hanhijärvi A. 1997 – Perpendicular-to-grain creep of Finnish softwoods in high temperature
drying conditions. Experiments and modelling in temperature range 95–125°C. VTT,
Technical Research Centre of Finland, VTT Publications 301.
Hanhijärvi A. 1998 – Deformation properties of Finnish spruce and pine wood in tangential
and radial directions in association to high temperature drying. Part I. Experimental
techniques for conditions simulating the drying process and results on shrinkage,
hygrothermal deformation, modulus of elasticity and strength. Holz als Roh- und
Werkstoff 56, 373-380.
Hanhijärvi A. 1999 – Deformation properties of Finnish spruce and pine wood in tangential
and radial directions in association to high temperature drying. Part II: Experimental
results under constant conditions. Holz als Roh- and Werkstoff 57, 365-372.
Hanhijärvi A. 2000a – Deformation properties of Finnish spruce and pine wood in tangential
and radial directions in association to high temperature drying. Part III. Experimental
results under drying conditions (mechano-sorptive creep). Holz als Roh- und Werkstoff
58, 63–71.
Hanhijärvi A. 2000b – Deformation properties of Finnish spruce and pine wood in tangential
and radial directions in association to high temperature drying. Part IV. Modelling. Holz
als Roh- und Werkstoff 58, 211–216.
Hanhijärvi A., Hunt D. 1998 – Experimental indication of interaction between viscoelastic
and mechano-sorptive creep. Wood Science and Technology 32, 57–70.
Hanhijärvi A., Mackenzie-Helnwein P. 2003 – Computational analysis of quality reduction
during drying of lumber due to irrecoverable deformation. I: Orthotropic viscoelastic-
mechanosorptive-plastic material model for the transverse plane of wood. Journal of
Engineering Mechanics 129, 996–1005.
Hatton A.P., Quarmby A. 1963 – The effect of axially varying and unsymmetrical boundary
conditions on heat transfer with turbulent flow between parallel plates. Int. J. Heat Mass
Transfer 6, 903-914.
Hatton A.P., Quarmby A., Grundy I. 1964 – Further calculations on the heat transfer with
turbulent flow between parallel plates. Int. J. Heat Mass Transfer 7, 817-823.
References 341

Hart C.A. 1964 – Principles of moisture movement in wood. Forest Products Journal 14,
207-214.
Hartley I.D. 2000 – Application of the Guggenheim-Anderson-deBoer sorption isotherm
model to Klinki pine (Araucaria klinki Lauterb). Holzforschung 54, 661–663.
Hartley D.I., Avramidis S. 1993 – Analysis of wood sorption isotherm using clustering
theory. Holzforschung 47, 163-167.
Hartley D.I., Avramidis S. 1994 – Water clustering phenomenon in two softwoods during
absorption and desorption processes. Journal of the Institute of Wood Science 76,
467-474.
Hartley I.D., Avramidis S. 2002 – Sorption hysteresis of Western Canadian softwood species.
J. Inst. Wood Science 16, 63-64.
Hartley I.D., Kamke F.A., Peemoeller H. 1992 – Cluster theory for water sorption in wood.
Wood Science and Technology 26, 83-99.
Häussler W. 1973 – Lufttechnische Berechnungen im Mollier-i,x-Diagramm. Theodor
Steinkopff, Dresden.
Hayashi K., Nakamura K., Kanagawa Y., Yasujima M. 1994 – Improvement of dryability and
its distribution in squared timber by local steam explosion. Proceedings of 4th IUFRO
International Wood Drying Conference, August 9-13, 1994, Rotorua, New Zealand,
359-365.
Haygreen J.G., Bowyer J.L. 1982 – Forest products and wood science. Iowa State University
Press.
Haygreen J.G., Bowyer J.L. 1996 – Forest products and wood science. An introduction. Third
Edition. Iowa State University Press/Ames.
Haygreen J.G., Bowyer J.L. 2002 – www.forestry.auburn.edu/elder/fp339/ch3/ch3.html
(06.06.02).
Haynes F.D., Ashton G.D. 1980 – Turbulent heat transfer in large aspect channels. J. Heat
Transfer 102, 384-386.
Helm R.F. 2002 – (Virgina Tech, Blacksburg); Wood 3434 2002 -
http://www.chemistry.vt.edu/chem-dept/helm/home.html (06.06.2002).
Heady R.D. 1997 – The wood anatomy of Callitris Vent. (Cupressaceae) an SEM Study.
Hillis W.E. 1985 – Biosynthesis of tannins. In: Biosynthesis and biodegradation of wood
components. Ed. T. Higuchi. Academic Press, USA, 325-348.
Hillis W.E. 1984 – High temperature and chemical effects on wood stability. Wood Science
and Technology 18, 281-293.
Hilpert R. 1933 – Wärmeangabe von geheizten Drähten und Rohren im Luftstrom. Forsch.
Gebiete Ingenieurw. 4, 215-224.
Hisada T. 1986 – Creep and set behaviour of wood related to kiln drying. (In Japanese). Bull.
For. & For. Prod. Res. Inst. 7, 31–130.
Horton R.K., Resch H. 1976 – Kiln-stick thickness effect on drying 2-inch Western hemlock
lumber. Forest Products Journal 26(3), 35-40.
Hua L., Bibeau E., He P. Gartshore I., Salcudean M., Bian Z., Chow S. 2001 – Modelling of
airflow in wood kilns. Forest Products Journal 51(6), 74-81.
342 Fundamentals of wood drying

Hukka A. 1996 – A mathematical model for simulation of softwood drying in temperatures


above boiling point of water with special attention to the boundary conditions. Drying
Technology 14, 1719-1732.
Hukka A. 1999 – The effective diffusion coefficient and mass transfer coefficient of Nordic
softwoods as calculated from direct drying experiments. Holzforschung 53, 534-540.
Hukka A., Oksanen O. 1999 – Convective mass transfer coefficient at wooden surface in jet
drying of veneer. Holzforschung 53, 204-208.
Hundt H.V. 1990 – Verfärbungen von Eichenkernholz und von Ilomba während der
Trocknung und Entwicklung vorbeugender Schutzmaßnahmen. Dissertation im
Fachbereich Biologie der Universität Hamburg, 118 S.
Hunt D.G. 1989 – Linearity and non-linearity in mechano-sorptive creep in softwood in
compression and bending. Wood Science and Technology 23, 323–333.
Hunt D., Gril J. 1996 – Evidence of physical ageing phenomenon in wood. Journal of
Materials Science Letters 15, 80–82.
Hunt D.G., Shelton C.F. 1988 – Longitudinal moisture-shrinkage coefficients of softwood at
the mechano-sorptive creep limit. Wood Science and Technology 22, 199–210.
Hunter A.J. 1996 – A complete theoretical isotherm for wood based on capillary
condensation. Wood Science and Technology 30, 179–192.
Husain S.S., Zamir K. 1968 – Studies on the effect of temperature and pH on the growth of 6
imperfect fungi. Sci. Ind. (Karachi) 6, 436-442.
Hrutfiord B.F., Luthi R. 1981 – Chemistry and of oregonin. In: Chemistry and morphology of
wood and wood components. Proceedings of International Symposium on Wood and
Pulping Chemistry, Stockholm, Vol. 1, 95-98.
Hwang K.S., Sung H.J., Hyun J.M. 1996 – Mass transfer measurements from a blunt-faced
flat plate in a uniform flow. Int. J. Heat Fluid Flow 17(2), 179-182.
Idelchik I.E. 1996 – Handbook of hydraulic resistance. Third Edition. Begell House, New
York.
Ifju G. 1973 – Influence of steaming on the properties of red oak. Part I. Structural and
chemical changes. Wood Science 6, 87-94.
Igarashi T. 1985 – Heat transfer from a square prism to an air stream. Int. J. Heat Mass
Transfer 28, 171-181.
Igarashi T. 1986 – Local heat transfer from a square prism to an air stream. Int. J. Heat Mass
Transfer 29, 777-784.
Igarashi T. 1987 – Fluid flow and heat transfer around rectangular cylinders (the case of
width/height ratio of a section of 0,33 ~ 1,5). Int. J. Heat Mass Transfer 30, 893-901.
Incropera F.P., DeWitt D.P. 1985 – Fundamentals of heat and mass transfer. John Wiley &
Sons.
Incropera F.P., DeWitt D.P. 1996 – Fundamentals of heat and mass transfer. Fourth Edition.
John Wiley & Sons, New York.
Inoue, M., Tanahashi, M., Rowell, R. 1993 – Steam or heat fixation of compressed wood.
Wood and Fiber Science 25, 224-235.
ISO 4469. 1981 – Wood – determination of radial and tangential shrinkage. First edition –
1981-11-01. Ref. No. ISO 4469-1981 (E). International Organization for Standardization.
References 343

ISO 4858. 1982a – Wood – determination of volumetric shrinkage. First edition – 1982-12-
01. Ref. No. ISO 4858-1982 (E). International Organization for Standardization.
ISO 4860. 1982b – Wood – determination of volumetric swelling. First edition – 1982-12-01.
Ref. No. ISO 4860-1982 (E). International Organization for Standardization.
ISO 4859. 1982c – Wood – determination of radial and tangential swelling. First edition –
1982-12-01. Ref. No. ISO 4859-1982 (E). International Organization for Standardization.
Irvine G.M. 1984 – The glass transition of lignin and hemicellulose and their measurement by
differential thermal analysis. Tappi Journal; Wood Chemistry 67, 118-121.
Jacobs P.M., Jones F.R. 1990 – Diffusion of moisture into two-phase polymer. Part 3.
Clustering of water in polyester resins. Journal of Material Science 25, 2471-2475.
Joergensen F.E. 2002 – How to measure turbulence with hot-wire anemometers - a practical
guide. Publication no.: 9040U6151. Skovlunde, Denmark: Dantec Dynamics A/S.
Johansson M., Perstorper M., Kliger R., Johansson G. 2001 – Distortion of Norway spruce
timber. Part 2. Modelling twist. Holz als Roh- und Werkstoff 59, 155-162.
Johansson A., Fyhr C., Rasmuson A. 1997 – High temperature convective drying of wood
chips with air and superheated steam, International Journal of Heat and Mass Transfer 40,
2843-2858.
Jost W. 1960 – Diffusion in solids, liquids, gases. Academic Press Inc., New York.
Joyet P. 1992 – Comportement differe du materiau bois dans le plan transverse sous des
conditions hydriques evolutives (Doctoral dissertation n° 812) L’Université Bordeaux I
(In French).
Joyet P., Meunier T. 1996 – Drying green oak under vacuum with superheated steam without
discoloration and drawback: industrial results. Proceedings of 5th International IUFRO
Wood Drying Conference, August 13-17, 1996, Québec, Canada, 169-176.
Kamke F.A., Casey L.J. 1988 – Gas pressure and temperature in the mat during flakeboard
manufacture, Forest Products Journal, 38, 41-43.
Kangas J. 1990 – Sahatavaran kuivauksen laadun parantaminen. Virumiskokeiden raportti
1990. (Improvement of the quality of timber drying. Report of creep experiments 1990).
Unpublished report. Technical Research Centre of Finland (VTT), Laboratory of
Structural Engineering, Espoo (In Finnish).
Karimi A., Perré P. 1997 – Un dispositif de mesure du débit et de la progression du liquide
dans la direction longitudinale du bois, Les Cahiers Scientifiques du Bois, Volume 1:
149-158, numéro spécial Instrumentation, ARBOLOR, Nancy.
Kauman W.G., Ananias, R.A., Gutiérrez M., Valenzuela H. 1994 – Non-Darcian permeability
in Chileau Tepa (Laurelia philippiana), Holzforschung 48, 77-81.
Kays W.M., Leung E.Y. 1963 – Heat transfer in annular passages. Hydrodynamically
developed turbulent flow with arbitrarily prescribed heat flux. Int. J. Heat Mass Transfer
6: 537-557.
Kawai S. 1980 – Moisture movement and drying stresses in wood, Doctor Dissertation,
Faculty of Agriculture, Kyoto.
Keey R.B. 1994 – Heat and mass transfer in kiln drying: A review. Proceedings of 4th IUFRO
International Wood Drying Conference, August 9-13, 1994, Rotorua, New Zealand,
22-44.
Keey R.B. Langrish T.A.G., Walker J.C.F. 2000 – Kiln-drying of lumber. Springer Berlin,
Heidelberg, New York.
344 Fundamentals of wood drying

Keey R.B., Pang S. 1994 – The high-temperature seasoning of softwood boards. Drying
Technology 12, 1297-1322.
Kelsey K.E. 1956 – The sorption of water vapour by wood. Australian Journal of Applied
Science 8, 42-54.
Keylwerth R. 1964 – Untersuchungen über freie und behinderte Quellung 4. Untersuchungen
über den Quellungsverlauf und die Feuchtigkeitsabhängigkeit der Rohdichte von Hölzern.
Holz als Roh- und Werkstoff 22, 255-258.
Kho P.C.S. 1993 – Mass transfer from in-line slabs. Ph.D. Thesis, University of Canterbury,
New Zealand.
Kho P.C.S., Keey R.B., Walker J.C.F. 1989 – Effects of minor board irregularities and air
flows on the drying rate of softwood timber boards in kilns. Proceedings of 2nd IUFRO
International Wood Drying Symposium, July 23-28, 1989, Seattle, USA, 150-157.
Kishimoto A., Fujita H. 1958 – Diffusion-controlled stress relaxation in polymers. III. Stress
relaxation in swollen polymers. Journal of Polymer Science 28, 569-585.
Kishimoto A., Fujita H., Odani H., Kurata M., Tamura M. 1960 – Succesive differential
absorption of vapors by glassy polymers. Journal of Physical Chemistry 65, 234-242.
Kifetew G. 1997 – Application of the earlywood-latewood interaction theory to the shrinkage
anisotropy of Scots pine. Proceedings of the International Conference on Wood-Water
Relations, 16-17 June 1997, Copenhagen, Denmark. COST Action E8, 165-171.
Kline E., Conners R., Aramon P. 1999 – Technology to Sort Lumber by Colour and Grain for
Furniture Parts. Proceedings of “Quality Lumber Drying in the Pacific Northwest:
Vertical Integration = Improved Profit” Conference, 30/9 - 2/10/1999, Seattle,
Washington. Forest Products Society, 67-73.
Kobayashi Y., Miura I., Kawai Y. 1999 – High performance drying using combination of HF
and hot air under atmospheric pressure. Proceedings of 6th International IUFRO Wood
Drying Conference, 25-28 January 1999, Stellenbosch, South Africa, 18-21.
Koch, G., Bauch J. 2000 – Discoloration in European beechwood (Fagus sylvatica L.) during
storage and drying. Proceedings of the 2nd COST E-15 Workshop on “Quality Drying of
Hardwood” 11-13, September 2000, Sopron, Hungary, pp. 5.
Koch G., Bauch J., Puls J., Schwab E., Welling J. 2000 – Vorbeugung gegen Verfärbungen
von Rotbuche. Holz-Zentralblatt 126, 74-75.
Koch G., Bauch J., Puls J., Welling J. 2002 – Ursachen und wirtschaftliche Bedeutung von
Holzverfärbungen. AFZ-DerWald 57, 315-318.
Koch G., Puls J., Bauch J. 2003 – Topochemical characterisation of phenolic extractives in
discolored beechwood (Fagus sylvatica L.). Holzforschung 57, 339-345.
Kollmann F.F.P. 1963 – Zur Theorie der Sorption. Forsch. Gebiete Ingenieurw. 29, 33-41.
Kollmann F.F.P., Côté W.A. 1968 – Principles of Wood Science and Technology. I Solid
Wood. Berlin, New York: Springer.
Kollmann F., Keylwerth R., Kübler H. 1951 – Verfärbung des Vollholzes und der Furniere
bei der künstlichen Holztrocknung. Holz als Roh- und Werkstoff 9, 382-391.
Korte H.E., Offermann W., Puls J. 1991 – Characterisation and preparation of substituted
xylo-oligosaccharides from steamed birchwood. Holzforschung 45, 419-424.
Košíková B., Hricovíni M., Cosentino C. 1999 – Interaction of lignin and polysaccharides in
beech wood (Fagus sylvatica) during drying processes. Wood Science and Technology
33, 373-380.
References 345

Kottke V., Blenke H., Schmidt K.G. 1977a – The influence of nose section and turbulence
intensity on the flow around thick plates in parallel flow. (In German). Wärme- und
Stoffübertragung 10, 159-174.
Kottke V., Blenke H., Schmidt K.G. 1977b – Determination of the local and average mass
transfer on thick plates in parallel flow with flow separation and reattachment. (In
German). Wärme- und Stoffübertragung 10, 217-232.
Koumoutsakos A., Avramidis S. 2002 – Mass transfer characteristics of western hemlock and
western red cedar. Holzforschung 56, 185-190.
Kreber B., Byrne A. 1994 – Discoloration of fem-fir wood: a review of the mechanisms.
Forest Products Journal 44, 35-42.
Kreber B., Fernandez M., Mc Donald A.G. 1998 – Migration of kiln brown stain precursors
during drying of radiata pine sapwood. Holzforschung 52, 441-446.
Kreber B., Haslett A.N. 1997 – A study of some factors promoting kiln drown stain formation
in radiata pine. Holz als Roh- und Werkstoff 55, 215-220.
Kreber B., Haslett A.N., Mc Donald A.G. 1999 – Kiln brown stain in radiata pine: A short
review on cause and methods for prevention. Forest Products Journal 49, 66-69.
Kreber B., Schmidt E., Byrne T. 1994 – Methyl bromide fumigation to control non-microbial
discolorations in western hemlock and red alder. Forest Products Journal 44, 63-67.
Krischer O., Kroell K. 1978 – Trocknungstechnik. 2.Bd. Trockner und Trocknungsverfahren.
2. Ed., Springer Berlin, Heidelberg, New York.
Kroell K. 1954 – Stroemungstechnische Massnahmen im Trocknerbau. Chem.-Ing.-Tech.
26(3), 132-140.
Kubinsky E., Ifju G. 1973 – Influence of steaming on the properties of red oak. Part I.
Structural and chemical changes. Wood Science 6, 87-94.
Lamvik M. 1982 – Drying of lumber: a method for evaluating a drying schedule. Drying’82,
A. S. Mujumdar Editor, Hemisphere Publ. Co., 227-232.
Lane J.C., Loehrke R.I. 1980 – Leading edge separation from a blunt plate at low Reynolds
number. J. Fluids Engng. 102, 494-496.
Langrish T.A.G. 1994 – Assessing the variability of mass-transfer coefficients in stacks of
timber with the aid of a numerical simulation. Proceedins of 4th IUFRO International
Wood Drying Conference, August 9-13, Rotorua, New Zealand, 122-129.
Langrish T.A.G. 1998 – The significance of the gaps between boards in determining the
moisture-content profiles in the drying of hardwood timber. Proceedings of 11th
International Drying Symposium, August 19-22, 1998, Halkidiki, Greece, Vol. B,
1546-1553.
Langrish T.A.G. 1999 – The significance of the gaps between boards in determining the
moisture content profiles in the drying of hardwood timber. Drying Technology 17(7/8),
1481-1494.
Langrish T.A.G. 2002 – Progress in the modelling of air flow patterns in timber kilns. Drying
Technology 20(9), 1789-1802.
Langrish T.A.G., Keey R.B., Kho P.C.S., Walker J.C.F. 1993 – Time-dependent flow in
arrays of timber boards: Flow visualization, mass-transfer measurements and numerical
simulation. Chem. Eng. Sci. 48(12), 2211-2223.
Langrish T.A.G, Keey R.B. 1996 – Improved timber kiln efficiency. CFX Update (12), 10.
346 Fundamentals of wood drying

Langrish T.A.G., Kho P.C.S., Keey R.B., Walker J.C.F. 1992 – Experimental measurement
and numerical simulation of local mass-transfer coefficients in timber kilns. Drying
Technology 10(3), 753-781.
Larson P.R., Kretschmann D.E., Clark A. III, Isebrands J.G. 2001 – Formation and properties
of juvenile wood in southern pines. A synopsis. USDA, FPL-GTR-129.
Launder B.E., Spalding D.B. 1974 – The numerical computation of turbulent flows. Computer
Methods in Applied Mechanics and Engineering 3, 269-289.
Le Bray Y., Prat M. 1999 – Three-dimensional pore network simulation of drying in capillary
porous media. Int. J. Heat Mass Transfer 42, 4207-4224.
Ledig S.F., Militzer K.-E. 1999a – Measurement and simulation of flow fields in convection
timber kilns. Proceedings of the 1st COST ACTION E15 Wood Drying Workshop. 13-14
October 1999, Edinburgh, UK, 9 pages.
Ledig S.F., Militzer K.-E. 1999b – Simulating the gas flow in a convective vacuum kiln.
Proceedings of 6th International IUFRO Wood Drying Conference, 25-28 January 1999,
Stellenbosch, South Africa, 232.
Ledig S.F., Militzer K.-E. 2000 – Measured gas velocity and moisture content distribution in
a convective vacuum kiln. Drying Technology 18(8), 1817-1832.
Ledig S.F., Nijdan J.J., Keey R.B. 2001 – Airflow distributions in the fillet spaces of a timber
stack. Drying Technology 19(8), 1697-1710.
Lee H.S. 1990 – Flow visualisation on high temperature wood drying. B.E. (Chem. And
Process) Report. New Zealand: University of Canterbury.
Leicester R.H. 1971 – A rheological model for mechano-sorptive deflections of beams. Wood
Science and Technology 5, 211–220.
Leiker M., Ledig S.F. 2001 – Air velocity profile in a duct between rough-sawn boards of
timber. Proceedings of 7th IUFRO International Wood Drying Conference, July 9-13,
2001, Tsukuba, Japan, 66-71.
Lide D.R. 1995 – Handbook of Chemistry and Physics. 76th Edition, The Chemical Rubber
Co.
Liese W. 1958 – Der Schutz des Buchenstammholzes gegen Risse, Einlauf und Verstocken.
Holz-Zentralblatt 84, 91-92.
Longauer J. 1994 – Temperature profiles of narrow interspace of the drying stack monitored
using holographic interferometry. Proceedings of VIII Drying Symposium, 20-22 June
1994, Warsaw, Poland, Vol. 2, 105-111.
Longauer J., wigo J. 1997 – Aerodynamika hydrotermických zariadení drevopriemyslu
(Aerodynamics of hydrothermal drying facilities in the wood-processing industry).
Zvolen (Slovakia): TU Zvolen.
Lowery D.P. 1979 – Vapor pressure generated in wood during drying, Wood Science 5,
73-80.
Long F.A., Thompson L.J. 1955 – Diffusion of water vapour in polymers. Journal of Polymer
Science 15, 413-426.
Luostarinen K., Luostarinen J. 2001 – Discoloration and deformations of birch parquet boards
during conventional drying. Wood Science and Technology 35, 517-528.
Luostarinen K., Möttönen V., Asikainen A., Luostarinen J. 2002 – Birch (Betula pendula)
wood discoloration during drying. Effect of environmental factors and wood location in
the trunk. Holzforschung 56, 348-354.
References 347

Magel E.A., Höll W. 1993 – Storage carbohydrates and adenine nucleotides in trunks of
Fagus sylvatica L. in relation to discolored wood. Holzforschung 47, 19-24.
Malmquist L. 1964 – On the heat transfer in convective drying of wood in superheated steam.
(In German). Holz als Roh- und Werkstoff 22, 95-106.
Marinescu I., Campean M., Marinescu N., Pescarus P. 2001 – Experimental results
concerning water movement in beech wood. Proceedings of 7th IUFRO International
Wood Drying Conference, July 9-13, 2001, Tsukuba, Japan, 210-215.
Mårtensson A. 1992 – Mechanical behaviour of wood exposed to humidity variations.
Doctoral dissertation, Lund Institute of Technology, Lund, Sweden.
Mårtensson A., Svensson S. 1997 – Stress–strain relationship of drying wood. Part 1.
Development of a constitutive model, Holzforschung 51, 472–478.
Mason E.A., Malinaukas A.P. 1983 – Gas transport in porous media: the dusty-gas model.
Elsevier, Amsterdam.
Mauget B., Perré P. 1999 – A large displacement formulation for anisotropic constitutive
laws, Eur. J. of Mech. A/Solids 18, 859-877.
McCormick D.C., Lessmann R.C., Test F.L. 1984 – Heat transfer to separated flow regions
from a rectangular prism in a cross stream. J. Heat Transfer 106, 276-283.
McCurdy M.C., Keey R.B. 2002 – The effect of growth-ring orientation on moisture
movement in the high-temperature drying of softwood boards. Holz als Roh- und
Werkstoff 60, 363-368.
McDonald A.G., Fernandez M., Kreber B. 1997 – Chemical and UV-VIS spectroscopic study
on kiln brown stain formation in radiata pine. Proceedings of the 9th International
Symposium of Wood and Pulping Chemistry. June 9-12, 1997, Montreal, Canada,
70.1-70.5.
McDonald A.G., Fernandez M., Kreber B., Laytner F. 2000 – The chemical nature of kiln
brown stain in radiata pine. Holzforschung 54, 12-22.
McIntosh D.C. 1955 – Shrinkage of red oak and beech. Forest Products Journal 5, 355-359.
McMillen J. 1975 – Physical characteristics of seasoning discolorations in sugar maple
sapwood. USDA FS, FPL Research Paper, FPL 248, p. 31.
McMillen J.M. 1976 – Control of reddish-brown coloration in drying maple sapwood. Res.
Note FPI-0231, USDA For. Serv., Forest Products Lab., Madison, WI, pp. 8.
Mcnamara S.W., Hart C.A. 1971 – An analysis of internal and average diffusion coefficients
for unsteady-state movement of moisture in wood. Wood Science 4, 37-45.
Merriam-Webster OnLine Unabridged Dictionary. 2003 – http://www.m-w.com.
Merzkirch W. 1987 – Flow visualization. Academic Press Inc.
Meyer R. W. 1971 – Influence of pit aspiration on earlywood permeability of Douglas-fir.
Wood and Fiber 2, 328-339.
Miller D.S. 1990 – Internal flow systems. Second Ed. Cranfield: BHRA (Information
Services).
Miller D.J., Knutson D.M., Tocher R.D. 1983 – Chemical brown staining of Douglas-fir
sapwood. Forest Products Journal 33, 44-48.
Millett M.A. 1952 – Chemical brown stain in sugar pine. Forest Products Journal 2, 232-236.
348 Fundamentals of wood drying

Milota M.R. 1994 – Mass transfer coefficients as a function of position in a lumber kiln.
Proceedings of 4th IUFRO International Wood Drying Conference, August 9-13, 1994,
Rotorua, NewZealand, 114-121.
Milota R.M. 1999 – Drying wood: the past, present and future. Proceedings of 6th
International IUFRO Wood Drying Conference, 25-28 January 1999, Stellenbosch, South
Africa, 1-10.
Mitchell R.L., Seborg R.M., Millet M.A. 1953 – Effect of heat on the properties and chemical
composition of Douglas-fir wood and its major components. Journal of FPRS, 38-42,
72-73.
Mohager S., Toratti T. 1993 – Long term bending creep of wood in cyclic relative humidity.
Wood Science and Technology 27, 49-59.
Morén T. 1993 – Creep, deformation and moisture redistribution during air convective wood
drying and conditioning. PhD Thesis, Luleå University of Technology, Sweden.
Morén T., Sehlstedt-Persson M. 1992 – Cupping of center boards during drying due to
anisotropic shrinkage. Proceedings of the 3rd IUFRO Conference on Wood Drying, 18-21
August 1992,Vienna, Austria, 160-164.
Morrow N.R. 1970 – Physics and thermodynamics of capillary, Industrial and Engineering
Chemistry 62, 32-56.
Mrowec S. 1980 – Defects and diffusion in solids. Elsevier Scientific Publishing Company.
Murmanis L. 1975 – Formation of tylosis in felled Quercus rubra L. Wood Science and
Technology 9, 3-14.
Muth G. 2002 – Pacific Union College http://www.puc.edu/Faculty/Gilbert_Muth/
(22.08.2002).
Navi P., Pittet V., Plummer C.J.G. 2002 – Transient moisture effects on wood creep. Wood
Science and Technology 36, 447-462.
Newman A.B. 1931a – The drying of porous solids. Diffusion and surface emission
equations. Transactions of the AIChE 27, 203-220.
Newman A.B. 1931b – The drying of porous solids: diffusion calculations. Transactions of
the AIChE 27, 310-333.
Newns A.C. 1956 – The sorption and desorption kinetics of water in a regenerated cellulose.
Transactions of the Faraday Society 52, 1533-1545.
Nijdam J.J., Keey R.B. 1996 – Influence of local variations of air velocity and flow direction
reversals on the drying of stacked timber boards in kilns. Trans IChemE 74 (Part A),
882-892.
Nijdam J.J., Keey R.B. 1999 – Airflow behaviour in timber (lumber) kilns. Drying
Technology 17(7/8), 1511-1522.
Nijdam J.J., Keey R.B. 2000 – The influence of kiln geometry on flow maldistribution across
timber stacks in kilns. Drying Technology 18(8), 1865-1877.
Nijdam J.J., Keey R.B. 2002 – An experimental study of airflow in lumber kilns. Wood Sci.
Technol. 36(1), 19-26.
Nisho S. 1976 – Estimation of drying stress in wood by cup method. II Comparison of cup
method with slice method. Mokuzai Gakkaishi 22, 180-186.
Nitsche W. 1994 – Stroemungsmesstechnik. Springer Berlin, Heidelberg, New York.
References 349

Oldham N.D., Wilcox W.W. 1981 – Control of brown stain in sugar pine with
environmentally acceptable chemicals. Wood and Fiber 13, 182-191.
Olek W., Weres J. 2001 – The inverse method for diffusion coefficient identification during
water sorption in wood. Proceedings of COST Action E15 3rd Workshop on “Softwood
Drying to Specific End-Uses”. 11-13 June 2001, Helsinki, Finland, pp 1-7.
Online Dictionary based on Webster's Revised Unabridged Dictionary. 2003 –
http://dict.die.net
Ormarsson S. 1995 – A finite element study of the shape stability of sawn timber subjected to
moisture variations. Division of Structural Mechanics, Lund University, Sweden.
Ormarsson S., Cown D., Dahlblom O. 2003 – Finite element simulations of moisture related
distortion in laminated products of Norway spruce and radiata pine. Proceedings of 8th
International IUFRO Wood Drying Conference, 24-29 August 2003, Brasov, Romania,
27-32.
Ormarsson S., Dahlblom O., Petersson H. 1996 – Influence of Annual Ring Orientation on
Shape Stability of Sawn Timber. Proceedings of 5th International IUFRO Wood Drying
Conference, August 13-17, 1996, Québec, Canada, 427-436.
Ota T. 2000 – A survey of heat transfer in separated and reattached flows. ASME Appl.
Mech. Rev. 53(8), 219-235.
Ota T., Asano Y., Okawa J. 1981 – Reattachment length and transition of the separated flow
over blunt plate. Bull. JSME 24, 941-947.
Ota T., Kon N. 1974 – Heat transfer in the separated and reattached flow on a blunt flat plate.
J. Heat Transfer 96, 459-462.
Ota T., Kon N. 1979 – Heat transfer in the separated and reattached flow over blunt flat
plates – effect of nose shape. Int. J. Heat Mass Transfer 22, 197-206.
Ower E., Pankhurst R.C. 1977 – The measurement of air flow. 5. Ed. Pergamon Press Oxford,
New York.
Özisik M.N., Cotta R.M., Kim W.S. 1989 – Heat transfer in turbulent forced convection
between parallel-plates. Can. J. Chem. Engng. 67, 771-776.
Panshin A. J., de Zeeuw C. 1980 – Textbook of wood technology. 4th ed. McGraw–Hill, New
York.
Pang S. 1996 – External heat and mass transfer coefficients for kiln drying of timber. Drying
Technology 14, 859-871.
Pang S., Haslett A.N. 1995 – The application of mathematical models to the commercial high-
temperature drying of softwood lumber. Drying Technology 13, 1635-1674.
Pang S., Keey R.B., Walker J.C.F. 1994 – Modelling of the high-temperature drying of mixed
sap and heartwood boards. Proceedings of 4th IUFRO International Wood Drying
Conference, August 9-13, 1994, Rotorua, New Zealand, 430-439.
Patankar S.V. 1980 – Numerical heat transfer and fluid flow. Hemisphere New York.
Park G.S. 1953 – An experimental study of the influence of various factors on the time
dependent nature of diffusion in polymers. Journal of Polymer Science 11, 97-113.
Park G.S. 1968 – The glassy state and slow process anomalies. Chapter 5 in “Diffusion in
Polymers”, Edited by Crank, J. and Park, G.S., Academic Press, New York.
350 Fundamentals of wood drying

Passard J., Perré P. 2001 – Creep tests under water-saturated conditions: do the anisotropy
ratios of wood change with the temperature and time dependency? Proceedings of 7th
International IUFRO Wood Drying Conference, July 9-13, 2001, Tokyo, Japan, 230-237.
Pentoney R.E. 1953 – Mechanisms affecting tangential vs. radial shrinkage. J. For. Prod. Res.
Soc. 3, 27-32.
Peralta P.N., Skaar C. 1993 – Experiments on steady-state nonisothermal moisture movement
in wood. Wood and Fiber Science 25, 124-135.
Perré P. 1987 – Measurements of softwoods permeability to air: importance upon the drying
model. Heat and Mass Transfer 14, 519-529.
Perré P. 1992 – Transferts couplés en milieux poreux non-saturés. Possibilités et limitations
de la formulation macroscopique. Habilitation à Diriger des Recherches, INPL, Nancy.
Perré P. 1995 – Drying with internal vaporization: introducing the concept of Identity Drying
Card. Drying Technology 13, 1077-1097.
Perré P. 1996 – The numerical modelling of physical and mechanical phenomena involved in
wood drying: an excellent tool for assisting with the study of new processes. Tutorial.
Proceedings of 5th International IUFRO Wood Drying Conference, August 13-17, 1996,
Québec, Canada, 9-38.
Perré P. 1997 – Image analysis, homogenization, numerical simulation and experiment as
complementary tools to enlighten the relationship between wood anatomy and drying
behavior. Drying Technology 15, 2211-2238.
Perré P. 1998 – The use of homogenisation to simulate heat and mass transfer in wood:
towards a double porosity approach. Keynote lecture. Proceedings of 11th International
Drying Symposium, August 19-22, 1998, Thessaloniki, Greece, 57-72.
Perré P. 1999 – How to get a relevant material model for wood drying simulation?
Proceedings of the First COST Action E15 Wood Drying Workshop, Edinburgh,
Scotland, 27 pages.
Perré P. 2001a – Wood as a multi-scale porous medium: observation, experiment, and
modelling. Keynote lecture. Proceedings of 1st International Conference of the European
Society for Wood Mechanics, April 19-21, 2001, Lausanne, Switzerland, 403-422.
Perré P. 2001b – The drying of wood: the benefit of fundamental research to shift from
improvement to innovation. Keynote lecture. Proceedings of 7th IUFRO International
Wood Drying Conference, July 9-13, 2001, Tsukuba, Japan, 2-13.
Perré P. 2003 – The role of wood anatomy in the drying of wood: “Great oaks from little
acorns grow”, Keynote lecture. Proceedings of 8th International IUFRO Wood Drying
Conference, 24-29 August 2003, Brasov, Romania, 11-24.
Perré P. 2004 – Evidence of dual scale porous mechanisms during fluid migration in
hardwood species. Part 2. A dual scale computational model able to describe the
experimental results. Chinese J. Chem. Eng. 12, 783-791.
Perré P. 2005 – MeshPore: a software able to apply image-based meshing techniques to
anisotropic and heterogeneous porous media. Drying Technology 23, 1993-2006.
Perré P. 2007 – Multiscale aspects of heat and mass transfer during drying. Transport in
Porous Media 66, 59-76.
Perré P., Agoua E. 2002 – Mass transfer in MDF (medium density fiberboard): identification
of structural parameters from permeability and diffusivity measurements. Proceedings of
References 351

13th International Drying Symposium, 27-30 August 2002, Beijing, China, Volume A,
446-455.
Perré P., Aguiar O. 1999 – Fluage du bois "vert" à haute température (120°C):
expérimentation et modélisation à l'aide d'éléments de Kelvin thermo-activés. Annals of
Forest Science 56, 403-416.
Perré P., Degiovanni A. 1990 – Simulation par volumes finis des transferts couplés en milieux
poreux anisotropes: séchage du bois à basse et à haute température. Int. J. Heat and Mass
Transfer 33, 2463-2478.
Perré P., Karimi A. 2002 – Fluid migration in two species of beech (Fagus silvatica and
Fagus orientalis): a percolation model able to account for macroscopic measurements and
anatomical observations. MADERAS: Cience & Tecnologia 4(1), 50-68.
Perré P., Keey R. 2006 – Drying of wood: principles and practice. In: 3rd Handbook of
Industrial Drying, Dekker, New York, Chapter 36, 821-877.
Perré P., Martin M. 1994 – Drying at high temperature of heartwood and sapwood: theory,
experiment and practical consequence on kiln control. Drying Technology 12, 1915-1941.
Perré P., Moser M., Martin M. 1993 – Advances in transport phenomena during convective
drying with superheated steam or moist air. Int. J. Heat and Mass Transfer 36, 2725-2746.
Perré P., Passard J. 1995 – A control-volume procedure compared with the finite-element
method for calculating stress and strain during wood drying. Drying Technology, Special
Issue “Mathematical Modelling and Numerical Techniques For The Solution of Drying
Problems” 13, 635-660.
Perré P., Passard J. 2004 – A physical and mechanical model able to predict the stress field in
wood over a wide range of drying conditions. Drying Technology 22, 27-44.
Perré P., Rémond R. 2006 – A dual-scale computational model of kiln wood drying including
single board and stack level simulation. Drying Technology 24, 1069–1074.
Perré P., Thiercelin F. 2004 – Evidence of dual scale porous mechanisms during fluid
migration in hardwood species. Part I. Using the attenuation of a polychromatic x-ray
beam to determine the evolution of moisture content during imbibition of beech. Chinese
J. Chem. Eng. 12, 773-782.
Perré P., Rémond R., Aléon D. 2007 – Energy saving in industrial wood drying addressed by
a multi-scale computational model: board, stack and kiln. Drying Technology 25, 75–84.
Perré P., Thiercelin F., Aguiar O. 2000 – Prototype high temperature/high pressure kiln for
the evaluation of wood drying schedules. Drying Technology 18, 1849-1863.
Perré P., Turner I.W. 1996 – Using a set of macroscopic equations to simulate heat and mass
transfer in porous media: some possibilities illustrated by a wide range of configurations
that emphasize the role of internal pressure. In: Numerical methods and Mathematical
modelling of the Drying Process (Edited by I.W.Turner and A. Mujumdar), Marcel
Dekker, New York, 83-156.
Perré P., Turner I.W. 1997 – Microwave drying of softwood in an oversized waveguide.
AIChE Journal 43, 2579-2595.
Perré P., Turner I.W. 1999 – A 3D version of TransPore: A comprehensive heat and mass
transfer computational model for simulating the drying of porous media. Int. J. Heat Mass
Transfer 42, 4501-4521.
352 Fundamentals of wood drying

Perré P., Turner I.W. 2000 – The use of numerical simulation as a cognitive tool for studying
the microwave drying of softwood in an over-sized waveguide. Wood Science and
Technology 33, 445-464.
Perré P., Turner I.W. 2001a – Determination of the material property variations across the
growth ring of softwood for use in a heterogeneous drying model. Part I: Capillary
pressure, tracheid model and absolute permeability. Holzforschung 55, 318–323.
Perré P., Turner I. W. 2001b – Determination of the material property variations across the
growth ring of softwood for use in a heterogeneous drying model. Part II: Use of
homogenisation to predict bound water diffusivity and thermal conductivity.
Holzforschung 55, 417–425.
Perré P., Turner I. W. 2002 – A heterogeneous wood drying computational model that
accounts for material property variation across growth rings. Chemical Engineering
Journal 86, 117-131.
Perré P., Turner I. W., Passard J. 1999 – 2-D solution for drying with internal vaporization of
anisotropic media. AIChE Journal 45, 13-26.
Perry A.E. 1982 – Hot-wire anemometry. Oxford University Press Oxford, New York.
Persson K. 1997 – Modelling of wood properties by a Micromechanical approach. Division of
Structural Mechanics, Lund University, Sweden.
Plath E., Plath L. 1955 – Papierchromatographische Untersuchungen an Dämpfkondensaten
von Rotbuche. Holz als Roh- und Werkstoff 13, 226-237.
Plumb O.A., Brown C.A., Olmstead B.A. 1984 – Experimental measurement of heat and
mass transfer during convective drying of southern pine. Wood Science and Technology
18, 187-204.
Prat M. 1995 – Isothermal drying of non-hygroscopic capillary-porous materials as an
invasion percolation process. Int. J. Multiphase Flow 21, 875-892.
Prat M. 2000 – Recent advances in pore-scale models for drying of porous media.
Proceedings of 12th International Drying Symposium, August 28-31, 2000,
Noordwijkerhout, The Netherlands, Paper No. 366.
Princhananda C. 1966 – A study of some aspects of moisture sorption dynamics in wood.
Ph.D. Thesis, State University of New York, Syracuse, NY.
Požgaj A., Chovanec D., Kurjatko S., Babiak M. 1993 – Štruktúra a vlastnosti dreva. Príroda,
a.s. Bratislava.
Puls J. 1993 – Substrate analysis of forest and agricultural wastes. In: Bioconversion of Forest
and Agricultural Plant Residues. Eds. J.N. Saddler. CAP International, Wallingford, UK,
13-32.
Ranta-Maunus A. 1973 – A theory for the creep of wood with application to birch and spruce
plywood. Technical Research Centre of Finland, Building Technology and Community
Development, Helsinki, Publication 4.
Ranta-Maunus A. 1975 – The viscoelasticity of wood at varying moisture content. Wood
Science and Technology 9, 189–205.
Ranta-Maunus A. 1992 – Determination of drying stresses in wood when shrinkage is
prevented: Test method and modelling. Proceedings of the 3rd IUFRO Conference on
Wood Drying, 18–21 August 1992,Vienna, Austria, 139–144.
Ranta-Maunus A. 1993 – Rheological behaviour of wood in directions perpendicular to the
grain. Materials and Structures 26, 362–369.
References 353

Ranta-Maunus A., Kortesmaa M. 1988 – An analysis of the state of stress in timber caused by
moisture gradient. in Proceedings of IUFRO Timber Engineering Meeting, Turku,
Finland, 113-116.
Ranta-Maunus A., Forsén H., Hanhijärvi A., Hukka A., Partanen J. – 1995 Sahatavaran
kuivauksen simulointi. (Simulation of timber drying.) VTT Technical Research Centre of
Finland, VTT Julkaisuja 805.
Refai Ahmed G., Yovanovich M.M. 1997 – Experimental study of forced convection from
isothermal circular and square cylinders and toroids. J. Heat Transfer 119, 70-79.
Rémond R., Perré P., Mougel E. 2005 – Using the concept of thin dry layer to explain the
evolution of thickness, temperature and moisture content during convective drying of
Norway spruce boards. Drying Technology 23, 249-271.
Resch H., Gautsch E. 2000 – Drying of lumber in vacuum using dielectric heating.
Proceedings of 2nd Workshop of COST Action E15 “Quality Drying of Hardwood”,
11-13 September 2000, Sopron, Hungary, 10 pages.
Resch H., Hansmann C., Pokorny M. 2000 – The colour of wood from white oak.
Holzforschung 54, 13-15.
Ressel B.J. 1994 – State-of-the-art on vacuum drying of timber. Proceedings of 4th IUFRO
International Wood Drying Conference, August 9-13, 1994, Rotorua, New Zealand,
255-262.
Riepen M., Gard W. 1996 – Analysis and optimisation of fast timber drying processes
through the use of computer simulations. Proceedings of 10th International Drying
Symposium, July 30 – August 2, 1996, Kraków, Poland, Vol. A, 703-710.
Rogers C.E. 1965 – Physics and chemistry of organic solid state. Vol II, Chapter 6.
Interscience Publishers.
Rohsenow W.M., Hartnett J.P., Cho H.H. 1998 – Handbook of heat transfer. McGraw-Hill
Co.
Rosen H.N. 1974 – Penetration of water into hardwoods. Wood and Fiber Science 5, 275-287.
Rosen H.N. 1976 – Exponential dependency of the moisture diffusion coefficient on moisture
content. Wood Science 8, 174-179.
Rosen H.N. 1978 – The influence of external resistance on moisture adsorption rates in wood.
Wood and Fiber 10, 218-228.
Rosenkilde A., Glover P. 2002 – High resolution measurement of the surface layer moisture
content during drying of wood using a novel magnetic resonance imaging technique.
Holzforschung 56, 312-317.
Rosenkilde A., Soderstrom O. 1996 – Measurements of moisture transport coefficients in
wood during drying. Proceedings of 5th International IUFRO Wood Drying Conference,
August 13-17, 1996, Québec, Canada, 523-528.
Rowell R.M., Lichtenberg R.S., Larsson P. 1993 – Stability of acetylated wood to
environmental changes. Wood and Fibre Science 25, 359-364.
Sakakibara M., Endo K. 1976 – Analysis of heat transfer for turbulent flow between parallel
plates. Intern. Chem. Engin. 16(4), 728-733.
Salamon M. et al. 1975 – Sorption studies of softwoods. Dept. Envir. Can. For. Serv. Ottawa,
publ .1346.
354 Fundamentals of wood drying

Salin J.-G. 1989 – Remarks on the influence of heartwood content in pine boards on final
moisture content and degrade. Proceedings of 2nd IUFRO International Wood Drying
Symposium, July 23-28, 1989, Seattle, USA.
Salin J.-G. 1992 – Numerical prediction of checking during timber drying and a new
mechano-sorptive creep model. Holz als Roh- und Werkstoff 50, 195–200.
Salin J.-G. 1996a – Mass transfer from wooden surfaces. Proceedings of 10th International
Drying Symposium, July 30 – August 2, 1996, Kraków, Poland, Vol. A, 711-718.
Salin J.-G. 1996b – Prediction of heat and mass transfer coefficients for individual boards and
board surfaces. A review. Proceedigs of 5th IUFRO International Wood Drying
Conference, August 13-17, 1996, Quebec City, Canada, 49-58.
Salin J.-G. 1999 – Simulation models: from a scientific challenge to a kiln operator tool.
Proceedigs of 6th International IUFRO Wood Drying Conference, 25-28 January 1999,
Stellenbosch, South Africa, 177-185.
Salin J.-G. 2001 – Global modelling of kiln drying, taking local variations in the timber stack
into consideration. Proceedings of 7th IUFRO International Wood Drying Conference,
July 9-13, 2001, Tsukuba, Japan, 34-39.
Salin J.-G. 2002a – Theoretical analysis of mass transfer from wooden surfaces. Proceedings
of 13th International Drying Symposium, August 27-30, 2002, Beijing, China, Vol. C,
1826-1834.
Salin J.-G. 2002b – Unpublished calculations.
Salin J.G. 2005 – Drying of Sapwood Analysed as an Invasion Percolation Process, 3rd
Nordic Drying Conference, Karlstad, Sweden.
Salin J.-G.; Hinnela J., Ohman G. 1998 – Calculation of drying behaviour in different parts of
a timber stack. Proceedings of 11th International Drying Symposium, August 19-22,
1998, Halkidiki, Greece, 1603-1610.
Salin J.-G., Öhman G. 1998 – Calculation of drying behaviour in different parts of a timber
stack. Proceedings of 11th International Drying Symposium, August 19-22, 1998,
Halkidiki, Greece, Vol. B, 1603-1610.
Salmén L. 1984 – Viscoelastic properties of in situ lignin under water-saturated conditions.
Journal of Materials Science 19, 3090-3096.
Salmén N.L., Fellers C. 1982 – The fundamentals of energy consumption during viscoelastic
and plastic deformation of wood. Transactions of the Technical Section, A Journal of Pulp
and Paper Science 8: TR93–99.
Sandberg D. 1998 – Value activation with vertical annual rings – material, production,
products. PhD Thesis, Royal Institute of Technology, Stockholm, Sweden.
Sandland K.M., Tronstad S. 2001 – Possibilities to control deformations in wood during
drying to meet the requirements from timber end-users. Proceedings of COST Action E15
3rd Workshop on “Softwood Drying to Specific End-Uses”. 11-13 June 2001, Helsinki,
Finland, 10 pages.
Sandland K.M., Tronstad S. 2002 – Increased yield by reduced cupping – Reflections and
initial experiments of press drying. Proceedings of 4th European COST Action E15
Workshop, May 30-31, 2002, Santiago de Compostela, Spain, Paper No. 3.
Sarni F., Moutounet M., Puech J.-L., Rabier P. 1990 – Effect of heat treatment on oak wood
extractable compounds. Holzforschung 44, 461-466.
References 355

Sawabe O. 1974 – Studies on the thermal softening of wood. III. Effects of the temperature on
the bending creep of dry Hinoki wood. Mokuzai Gakkaishi 20, 517–522.
Schill V., Korger V., Stahl W. 1994 – First results concerning the application of the I/D
process in the final drying of hardwoods. Proceedings of 4th IUFRO International Wood
Drying Conference, August 9-13, 1994, Rotorua, New Zealand, 374-381.
Schlichting H., Gersten K. 1997 – Grenzschicht-Theorie. 9. Ed. Springer Berlin, Heidelberg,
New York.
Schlichting H. 1958 – Grenzschicht-Theorie. 3. Ed., G. Braun Karlsruhe.
Schlueter R., Fessel F. 1939 – Neue praktische Erfahrungen bei der kuenstlichen
Holztrocknung, Trockentechnik. Holz als Roh- und Werkstoff 2(5), 169-195.
Schmidt E. 1997 – Penetration of fumigants into logs for pest eradication and stain
prevention. Proceedings of 1997 Annual Conference on “Methyl Bromide Alternatives
and Emissions Reductions”. U.S. E.P.A. and U.S.D.A., 3-5/11/1997, San Diego,
California, 90, 1-2.
Schmidt E., Christopherson E., Highley T., Freeman M. 1995 – Trials of new treatments for
prevention of kiln brownstain of white pine (Pinus strobus). IRG on Wood Preservation,
Section 3, 26th Annual Meeting, Helsingor, Denmark, pp. 5.
Schmidt E.L., Ambugrey T.L., Kitchens S.C. 1998 – Mill trial confirms control of lumber
graystain and sticker shadow after fumigation of southern hardwood logs with methyl
bromide. Forest Products Journal 48, 50-52.
Schmidt E. 1969 – Properties of water and steam in SI – units. Springer Berlin, Heidelberg,
New York.
Seborg R.M., Tarkow H., Stamm J. 1953 – Effect of heat upon the dimensional stabilization
of wood. Journal of FPRS, 59-67.
Seeling U. 2002 – Universität Freiburg http://www.forst.uni-freiburg.de/fobawi/fob/
fob_lehre/211b/material/01_grundlagenholz.pdf (19.08.2002).
Sehlstedt-Persson S.M.B. 1995 – High-temperature drying of Scots pine. A comparison
between HT- and LT-drying. Holz als Roh- und Werkstoff 53, 95-99.
Sherwood T.K. 1929 – The drying of solids-II. Industrial and Engineering Chemistry 21,
976-980.
Shih T.-H., Liou W.W., Shabbir A., Zhu J. 1995 – A new k-ε eddy-viscosity model for high
Reynolds number turbulent flows - model development and validation. Computers Fluids
24(3), 227-238.
Siau J.F. 1984 – Transport processes in wood. Springer Berlin, Heidelberg, New York.
Siau J.F. 1995 – Wood: influence of moisture on physical properties. Department of Wood
Science and Forest Products. Virginia Polytechnic Institute and State University,
Blacksburg, VA.
Siau J.F, Avramidis S. 1993 – Application of a thermodynamic model to nonisothermal
diffusion experiment. Wood Science and Technology 27, 95-114.
Siau J.F, Avramidis S. 1996 – The surface emission coefficient for wood. Wood and Fiber
Science 28, 178-185.
Siau J.F., Babbiak M. 1983 – Experiments on nonisothermal moisture in wood. Wood and
Fiber Science 15, 40-46.
356 Fundamentals of wood drying

Siau J.F., Jin Z. 1985 – Nonisothermal moisture diffusion experiments analyzed by four
alternative equations. Wood Science and Technology 19, 151-157.
Siau J.F., Bao F., Avramidis S. 1986 – Experiments in nonisothermal diffusion of moisture in
wood. Wood and Fiber Science 18, 84-89.
Simpson W.T. 1973 – Predicting equilibrium moisture content of wood by mathematical
models. Wood and Fiber 5, 41–49.
Simpson W.T. 1979 – Sorption theories for wood. Symposium on wood moisture content –
temperature and humidity relationships. VPI&SU, Blacksburg, Virginia, 36–46.
Simpson W. 1991 – Dry kiln operator’s manual. Agric. Handbook AH-188, Madison, WI,
USDA, FS, FPL.
Skaar C. 1954 – Analysis of methods for determining the coefficient of moisture diffusion in
wood. Forest Products Journal 4(12), 403-410.
Skaar C. 1958 – Moisture movement in beech below the fiber saturation point. Forest
Products Journal 8(12), 352-357.
Skaar C. 1988 – Wood-water relations. Springer-Verlag Heidelberg, New York, Berlin.
Skaar C., Babiak M. 1982 – A model for bound water transport in wood. Wood Science and
Technology 16, 123-138.
Skarvelis M. 1996 – Utilization of solar energy for drying sawn timber. Ph.D. Thesis,
Aristotelian University of Thessaloniki, Department of Forestry and Natural Environment,
Wood Technology Sector.
Skarvelis M., Karaoulanis T. 2001 – Softwood lumber (domestic and imported) in Greece:
Problems related to wood drying conditions. Proceedings of COST Action E15 3rd
Workshop on “Softwood Drying to Specific End-Uses”. 11-13 June 2001, Helsinki,
Finland, pp. 7.
Smith W.B., Smith A., Neauhauser E.F. 1994 – Radio-frequency/vacuum drying of red oak:
energy, quality, value. Proceedings of 4th IUFRO International Wood Drying Conference,
August 9-13, 1994, Rotorua, New Zealand, 263-270.
Söderström O., Salin J.-G. 1993 – On determination of surface emission factors in wood
drying. Holzforschung 47, 391-397.
Sörensen A. 1969 – Mass transfer coefficients on truncated slabs. Chem. Eng. Sci. 24,
1445-1460.
Spalt H.A. 1997 – Water – vapor sorption by woods of high extractive content. Symposium
on wood moisture content – temperature and humidity relationships. VPI&SU,
Blacksburg, Virginia, 55–61.
Sparrow E.M., Hajiloo A. 1980 – Measurements of heat transfer and pressure drop for an
array of staggered plates aligned parallel to an air flow. J. Heat Transfer 102, 426-432.
Spolek G.A., Plumb O.A. 1980 – A numerical model of heat and mass transport in wood
during drying. Porceedings of the 2nd International Drying Symposium, Montreal,
Canada, Vol. 2, 84-92.
Spolek G.A., Plumb O.A. 1981 – Capillary pressure in softwoods, Wood Science and
Technology 15, 189– 99.
Stamm A.J. 1956a – Diffusion of water into uncoated cellophane I. From rates of water
vapour adsorption and liquid water absorption. Journal of Physical Chemistry 60, 76-82.
References 357

Stamm A.J. 1956b – Diffusion of water into uncoated cellophane II. From steady-state
diffusion measurements. Journal of Physical Chemistry 60, 83-86.
Stamm A.J. 1959 – Bound-water diffusion into wood in the fiber direction. Forest Products
Journal 9, 27-32.
Stamm A.J. 1960a – Bound-water diffusion into wood in across-the-fiber direction. Forest
Products Journal 10, 524-528.
Stamm A.J. 1960b – Combined bound-water and water-vapor diffusion into sitka spruce.
Forest Products Journal 10, 644-648.
Stamm A.J. 1961 – Wood and cellulose-liquid relationships. Proceedings of the 50th Annual
Pacific Section of TAPPI, Portland (USA), 18-19.
Stamm A.J. 1963 – Permeability of wood to fluids. Forest Products Journal. 13, 503-507.
Stamm A.J. 1964 – Wood and cellulose science. The Ronald Press Company. New York.
Stamm A.J. 1967 – Movement of fluids in wood – Part II. Diffusion. Wood Science and
Technology 1, 205-230.
Stamm A.J., Hansen L.A. 1937 – Ind. Eng. Chem. 29: 831.
Stamm A.J., Loughborough W.K. 1942 – Variation in shrinking and swelling of wood. Trans.
Am. Soc. Mech. Eng. 64, 379–386.
Stamm A.J., Nelson R.M. 1961 – Comparison between measured and theoretical drying
diffusion coefficients for southern pine. Forest Products Journal 11, 536-543.
Stanish M.A., Schajer G.S., Kayihan F. 1986 – A mathematical model of drying for
hygroscopic porous media. AICHE Journal 32, 1301-1311.
Starkweather H.W. 1963 – Clustering of water in polymers. Polymer Letters 1, 133-138.
Starkweather H.W. 1975 – Some aspects of water clusters in polymers. Macromolecules 8,
476-479.
Stich K., Ebermann R. 1984 – Peroxidase and Polyphenoloxidase Isoeenzyme im Splint und
Kernholz der Eiche. Holzforschung 38, 239-242.
Sturany H. 1952 – Strömungsfragen beim Trocknen von Holz. Holz als Roh- und Werkstoff
10(5), 201-207.
Sun Z.F. 2001 – Numerical simulation of flow in an array of in-line blunt boards: mass
transfer and flow patterns. Chem. Eng. Sci. 56(5), 1883-1896.
Svensson S. 1995 – Strain and shrinkage force in wood under kiln drying conditions I.
Measuring strain and shrinkage under controlled climate conditions. Equipment and
preliminary results. Holzforschung 49, 363–368.
Svensson S. 1996 – Strain and shrinkage force in wood under kiln drying conditions II. Strain,
shrinkage and stress measurements under controlled climate conditions. Holzforschung
50, 463–469.
Svensson S., Mårtensson A. 2002 – Simulation of drying stresses in wood. II: Convective air
drying of sawn timber. Holz als Roh- und Werkstoff 60, 72–80.
Svensson S., Toratti T. 2002 – Mechanical response of wood perpendicular to grain when
subjected to changes of humidity. Wood Science and Technology. 36, 145-156.
SWST 2002 – Society of Wood Science and Technology http://www.swst.org/teach/set2/
struct1.html (06.08.02).
358 Fundamentals of wood drying

Tatsutani K., Devarakonda R., Humphrey J.A.C. 1993 – Unsteady flow and heat transfer for
cylinder pairs in a channel. Int. J. Heat Transfer 36, 3311-3328.
Taylor F.W. 1997 – Prevention of stain associated with kiln drying. In: Prevention of
discolorations in hardwood and softwood logs and lumber. Forest Products Society,
Madison, WI, pp. 90.
Terazawa S. 1965 – Methods for easy determination of kiln drying schedules of wood. Wood
Industry. 20, 216-226.
Terziev N. 1996 – Low-molecular weight sugars and nitrogenous compounds in Scots pine.
PhD Thesis, Swedish University of Agricultural Sciences, Uppsala.
Terziev N., Boutelje J. 1998 – Effect of felling time and kiln-drying on colour and
susceptibility of wood to mould and fungal stain during an above-ground field test. Wood
and Fiber Science 30, 360-367.
Terziev N., Ekstedt J. 1997 – Discoloration of white coatings applied on kiln- and air-dried
Scots pine lumber. Holz als Roh- und Werkstoff 55, 77-81.
Tesoro F. O., Kimbler O.V., Choong E.T. 1972 – determination of the relative permeability of
wood to oil and water. Wood Science 5, 21-26.
Tesoro F. O., Choong E.T., Kimbler O.V. 1974 – Relative permeability and the gross pore
structure of wood. Wood and Fiber 6(3), 226-236.
Test F.L., Lessmann R.C. 1980 – An experimental study of heat transfer during forced
convection over a rectangular body. J. Heat Transfer 102, 146-151.
Timell T.E. 1986 – Compression wood in gymnosperms. Springer Berlin, Heidelberg, New
York.
Tjeerdsma B.F., Boonstra M., Militz H. 1998 – Thermal modification of non-durable wood
species. 2. Improved wood properties of thermally treated wood. The International
Research Group on Wood Preservation, Section 4. Proceedings of 29th Annual Meeting,
Maastricht, June 14-19, 1998.
Theander O., Bjurman J., Boutelje J.B. 1993 – Increase in the content of low-molecular
weight carbohydrates at lumber surfaces during drying and correlations with nitrogen
content yellowing and mould growth. Wood Science and Technology 27, 381-389.
Toratti T. 1992 – Creep of timber beams in a variable environment. Helsinki University of
Technology, Lab. of Structural Engineering and Building Physics, Espoo, Report 31.
Tremblay C., Cloutier A., Fortin Y. 2000a – Determination of the effective water conductivity
of red pine sapwood. Wood Science and Techology 34, 109-124.
Tremblay C., Cloutier A,. Fortin Y. 2000b – Experimental determination of the convective
heat and mass transfer coefficients for wood drying. Wood Science and Technology 34,
253-276.
Trénard Y. 1980 – Comparaison et interprétation de courbes obtenues par porosimétrie au
mercure sur diverses essences de bois. Holzforschung 34, 139-146.
Tronstad S. 1971 – The influence of sawing pattern on the economy of sawmilling.
Norwegian Institute of Wood Technology (in Norwegian, unpublished work).
Trnka M., Ladomerský J. 1985 – Sorp né vlastnosti smrekového dreva po prirodzenom a
umelom sušení. Drevársky Výskum 104, 35-44.
Trübswetter T. 1995 – Die Trocknung heller Laubhölzer und ihre Problematik. Holz-
Zentralblatt 121, 2194-2198.
References 359

Tsoumis G. 1983 – Domi, Idiotites kai Axiopoiisi tou Xylou [Structure, properties and
utilization of wood]. Aristotelian University of Thessaloniki, p. 655.
Tuttle F. 1925 – A mathematical theory of the drying of wood. Franklin Institute Journal 200,
609-614.
Turner I. W., Perré P. 1996 – A synopsis of the strategies and efficient resolution techniques
used for modelling and numerically simulating the drying process. In: Turner I.,
Mujumdar, A.S. (Eds.) Mathematical modeling and numerical techniques in drying
technology. Marcel Dekker, New York.
Turner I.W., Puiggali J.R., Jomaa W. 1998 – A numerical investigation of combined
microwave and convective drying of a hygroscopic porous material: A study based on
pine wood. Trans IChemE 76, Part A, 193-209.
Van Meel D.A. 1958 – Adiabatic convection batch drying with recirculation of air, Chemical
Engineering Science 9, 36-44.
VDI Wärmeatlas 1994 – 7. Ed. Düsseldorf: VDI-Verlag, Lb3.
Viktorin Z. 1968 – Widerstandscharakteristik von Holzstapeln. Luft- und Kaeltetechnik (6),
267-270.
Voight H., Krischer O., Schaus H. 1940 – Feuchtigkeitsbewegung bei der
verdunstungstrocknung von holz. Holz als- Roh und Werkstoff 3, 305-321.
Wadso L. 1993 – Studies of water vapor transport and sorption in wood. Doctoral Thesis,
Lund University, Lund, Sweden.
Wagenführ R. 1996 – Holzatlas. Fachbuchverlag Leipzig.
Wagenführ R. 1999 – Anatomie des Holzes. DRW-Verlag, Leinfelden-Echterdingen.
Wang S.Y., Cho C.L. 1993 – Equilibrium moisture content of six wood species and their
influences. Mokkuzai Gakkaishi 39, 126–137.
Wang Z., Choong E. 1994 – Effect of presteaming on drying stresses of red oak using a
coating and bending method. Wood and Fiber 26, 527-535.
Ward J.C., Pong W.Y. 1980 – Wetwood in trees: a timber resource problem. Gen. Techn.
Report PNW-112, U.S. For. Serv., Pacif. Northwest For. and Range Exp. Sta.
Wassipaul F., Fellner J. 1987 – Verfärbung von Eichenschnittholz bei der künstlichen
Trocknung. Holzforschung und Holzverwertung 40, 1-5.
Wassipaul F., Fellner J. 1992 – Eichenverfärbung bei der Trocknung mit niederen
Temperaturen. Holzforschung und Holzverwertung 44, 86-88.
Watt I.C. 1960 – Kinetic studies of the wool-water system. Part I. The influence of water
concentration. Textile Resources Journal 30, 443-450.
Watt I.C., Algie J.E. 1961 – Kinetic studies of the wool-water system. Part III. The
mechanism of sorption at high water concentrations. Textile Resources Journal 31,
793-799.
Wegener G., Fengel D. 1987 – Investigation on colour changes resulting from drying of
European oakwood. Proceedings of Fourth International Symposium on Wood and
Pulping Chemistry 1, 121-123.
Welling J. 2000 – Timber drying research – Industrial needs and scientific expectations.
Proceedings of 2nd Workshop of COST Action E15 “Quality Drying of Hardwood”,
11-13 September 2000, Sopron, Hungary, 7 pages.
360 Fundamentals of wood drying

Welling J., Wöstheinrich A. 1995 – Reduzierung von Verfärbungen durch Heißdampf-


Vakuumtrocknung. Holz-Zentralblatt 121, 145-146.
Welty R. J., Wicks E.C., Wilson R.E. 1984 – Fundamentals of momentum, heat, and mass
transfer. 3rd Edition. John Wiley & Sons. New York.
Wengert E.M. 1979 – Psychrometric relationships and equilibrium moisture content of wood
at temperatures below 212°F (100°C). Symposium on wood moisture content –
temperature and humidity relationships. VPI &SU, Blacksburg, Virginia, 4–11.
Wengert E.M. 1988 – The Wood Doctor’s RX. Virginia Polytechnic Institute and State
University, Department of Forest Products.
Wengert E.M. 1990 – Chemical stain and stain control in hardwood lumber drying. In:
Hamel, M.P. (Ed.): Drying softwood and hardwood lumber for quality and profit. For.
Prod. Soc., USA.
Wengert E.M. 1992 – Causes and cures for stains in dried lumber: sticker stain, chemical stain
and blue stain. Univ. of Virginia, School of Natural Resources, Department of Forestry.
Forestry Facts No. 64. p. 6.
Wengert E.M., Meyer D. 2002 – Woodweb: www.woodweb.com/knowledge_base_images/
zp/warp_68-1.gif (22.08.2002).
Wengert G., Denig J. 1995 – Lumber drying - today and tomorrow. Forest Products Journal
45, 22-30.
Werner H. 1972 – Aufgaben der Luft- und Waermeverteilung in Konvektionstrocknern.
Chem.-Ing.-Tech. 44(8), 570-576.
Wheeler E.A., Baas P., Gason P.E. 1989 – IAWA list of microscopic features for hardwood
identification. IAWA Bulletin 10, 219-332.
Wiberg P. 1996 – CT-scanning during drying. Moisture content distribution in Pinus
silvestris. Proceedings of 5th International IUFRO Wood Drying Conference, August
13-17, 1996, Québec, Canada, 231-235.
Wiberg P. 2001 – X-ray CT-scanning of wood during drying. Doctoral Thesis, Luleå
University of Technology, Sweden.
Wiberg, P., Sehlstedt, S.M.B., Morén.T.J. 2000 – Heat and mass transfer during sapwood
drying above the fibre saturation point. Drying Technology 18, 1647-1664.
Wiedemann H.G.R., Nassif N.M., Gostelow J.P. 1989 – Study of the air flow profile inside a
timber drying kiln. Proceedings of 4th Australasian Conference on Heat and Mass
Transfer. Christchurch, NZ, 693-701.
Wilhelm L.R. 1976 – Numerical calculation of psychrometric properties in SI units.
Transactions of the ASAE 19, 318-321, 325.
Wu Q., Milota M.R. 1995 – Rheological behavior of Douglas-fir perpendicular to the grain at
elevated temperatures. Wood and Fiber Science 27, 285–295.
Wu Q., Milota M.R. 1996 – Mechano-sorptive deformation of Douglas-fir specimens under
tangential tensile stress during moisture adsorption. Wood and Fiber Science 28, 128-132.
Wu Q., Oliver A.R., Doe P.E. 1995 – A study of the boundary layer flow through a lumber
drying stack. Drying Technology 13(8/9), 2011-2026.
Yahiaoui K. 1991 – A rheological model to account for mechano-sorptive behaviour. In:
Martensson A., Ranta-Maunus A. and Seoane I. (Eds.). Presentations at the COST 508
Wood mechanics Workshop on Fundamental aspects of creep in wood, Lund, Sweden,
20–21 March 1991 (Commission of the European Communities), 27–35.
References 361

Yazaki Y., Bauch J., Endeward R. 1985 – Extractive components responsible for the
discoloration of Ilomba wood (Pycnanthus angolensis Exell). Holz als Roh- und
Werkstoff 43, 359-363.
Yeo H. 2001 – Evaluation of mass transfer in wood utilizing a colorimetric technique and
numerical analysis. Ph.D. Thesis, State Univ. of New York, Syracuse, New York.
Yoo S.Y., Goldstein R.J., Chung M.K. 1993 – Effects of angle of attack on mass transfer from
a square cylinder and its base plate. Int. J. Heat Mass Transfer 36, 371-381.
Yeo H., Smith W.B., Hanna R.B. 2001 – Evaluation of mass transfer in wood utilizing a
colorimetric technique and numerical analysis. Proceedings of 7th IUFRO International
Wood Drying Conference, July 9-13, 2001, Tsukuba, Japan, 200-205.
Yeo H., Smith W.B., Hanna R.B. 2002a – Determination of surface moisture content of wood
utilizing a colorimetric technique. Wood and Fiber Science 34, 419-424.
Yeo H., Smith W.B., Hanna R.B. 2002b – Mass transfer in wood evaluated with a
colorimetric technique and numerical analysis. Wood and Fiber Science 34, 657-665.
Yilgor N., Unsal O., Kartal S.N. 2001 – Physical, mechanical and chemical properties of
steamed beech wood. Forest Products Journal 51, 89-93.
Zhao H., Turner I.W. 1996 – An analysis of the finite-difference time-domain method for
modeling the microwave heating of dielectric materials within a three dimensional cavity
system, Journal of Microwave Power and Electromagnetic Energy 31, 199-214.
Zhao H., Turner I.W., TorgovnikovG. 1998 – An experimental and numerical investigation of
the microwave heating of wood. J. Microwave Power and Electromagnetic Energy 33,
121-133.
Zhao H., Turner I.W. 2000 – The use of a coupled computational model for sStudying the
microwave heating of wWood. J. Appl. Math. Modelling 24, 183-197.
Ziegler H. 1968 – Biologische Aspekte der Kernholzbildung. Holz als Roh- und Werkstoff
26, 61-68.
Zigrikas A. 2000 – Sticker staining in iroko strips (a practical solution to the problem).
Proceedings of 2nd Workshop of COST Action E15 “Quality Drying of Hardwood”,
11-13 September 2000, Sopron, Hungary, 4 pages.
Zimmermann G. 1974 – Untersuchungen über Art und Ursachen von Verfärbungen an
Bergahorn-Stammholz (Acer pseudoplatanus L.) Forstw. Cbl. 93, 247-261.
Zimmerman M. H. 1983 – Xylem structure and the ascent of sap. Springer-Verlag, Berlin.
Zobel B.J., Sprague J.R. 1998 – Juvenile wood in forest trees. Springer Berlin, Heidelberg,
New York.
Zombori B.G., Kamke F.A., Watson L.T. 2003 – Simulation of the internal conditions during
the hot-pressing process. Wood and Fiber Science 35, 2-23.
Zycha H. 1952 – Der Schutzanstrich für lagerndes Buchenholz. Holz-Zentralblatt 78, 1865.
15 List of symbols
15.1 Latin letters

A Area m2
ai j k l Linear elastic constitutive relation MPa
a Kiln plenum width m
b Kiln ceiling space m
C Constant m-1
c Molar concentration mol m-3
cp Specific heat J kg-1 K-1
D Diffusivity tensor m2 s-1
D Diameter m
Dh Hydraulic Diameter, 4A/U m
d Velocity distribution coefficient
E Young modulus MPa
F(u) Complete set of non-linear functions
f Doppler frequency s-1
f Fanning friction factor, λ/4
g Gravitational acceleration m s-2
H Relative humidity of the air
H Shape factor, δ∗/δ∗∗
H Stack height m
h Specific enthalpy J kg-1
h Channel height m
∆ hw Differential heat of sorption J kg-1
hh Heat transfer coefficient W m-2 K-1
hm Mass transfer coefficient m s-1
J Flux expression kg m-2 s-1
J Jacobian Matrix
K Absolute permeability m2
k Relative permeability
K Surface roughness m
K+ Dimensionless surface roughness, Kvτρ/µ m
k Turbulent kinetic energy m2 s-2
L Length m
Lv Latent heat of evaporation J kg-1
M Molar mass kg mol-1
mɺ Mass flow rate kg s-1
n Unit normal
n Number of samples
p Pressure Pa
pv Velocity pressure Pa
364 Fundamentals of wood drying

∆p Pressure difference Pa
Pr Prandtl number, µcp/λ
Qɺ Heat flow rate W
q Heat flux W m-2
R Gas constant J mol-1 K-1
Re2h Reynolds number based on 2h, v2hρ/µ
Ret Reynolds number based on t, vtρ/µ
Ret Turbulence Reynolds number, ρ k2/(εµ)
Rex Reynolds number based on x, vxρ/µ
S Volume saturation
Sv Entropy J mol-1 K-1
T Temperature K
Tɶ Dimensionless temperature difference
Tu Turbulence intensity
t Time s
t Board/Plate thickness m
u Displacement field m
U Wetted duct perimeter m
Vɺ Volumetric flow rate m3 s-1
Vd Variation coefficient of velocity distribution
v Velocity vector m s-1
v Time-mean velocity m s-1
v´ Fluctuation velocity m s-1
v+ Dimensionless velocity, v/vτ m s-1
vm Mixed-mean velocity m s-1
vx Velocity component in x direction m s-1
vτ Friction velocity, τW ρ m s-1
X Moisture content of wood (dry basis) kg kg-1
x Molar fraction
x Rectangular coordinate; m
Distance in flow direction
Y Moisture content of moist air (dry basis) kg kg-1
Y Dimensionless distance from wall, y/h
y Rectangular coordinate; m
Distance normal to surface
y* Dimensionless distance from wall, y x Re x
+
y Dimensionless distance from wall, yvτρ/µ
z Rectangular coordinate; elevation m
List of symbols 365

15.2 Greek Symbols

α Flow angle °
β Flow angle °
δ Boundary layer thickness m
δij Kronecker delta, tensor
δ∗ Displacement thickness m
δ∗∗ Momentum thickness m
ε Volume fraction
εij Scalar component of the strain tensor
ε Turbulent dissipation rate m2 s-3
η Branch loss ratio
ζ Pressure drop coefficient
Θ Angle of intersection rad
λ Thermal conductivity W m-1 K-1
λ Darcy friction factor, 4f
λ Wave length m
µb Chemical potential of bound water J kg-1
µ Dynamic viscosity kg m-1 s-1
µt Turbulent viscosity kg m-1 s-1
ν Poisson ratio
ν Kinematic viscosity m2 s-1
ρ Density kg m-3
σij Scalar component of the stress tensor MPa
σ Surface tension N m-1
τW Wall shear stress N m-2
Φ Volumetric source term W m-3
φ Porosity m3 m-3
ϕ Phase potential
χ Depth scalar m
Ψ Conserved quantity
ω Mass fraction

15.3 Superscripts and Subscripts

A Area
a Air
am Arithmetic mean
b Bound water
c Capillary
c Critical value
eff Effective property
elas Elastic part of the strain tensor
eq Equilibrium
366 Fundamentals of wood drying

FL Flux limited
FSP Fibre saturation point
f Friction
g Gas phase
g Gap
i General index for tensors
i Inlet
j General index for tensors
L (Pressure) Loss
l Location
max Maximum
mec Strain tensor due to mechanical stress
mem Strain tensor due to the memory effect of wood material
Nb Neighbouring CV point
o Outlet
P CV centroid point
P (Pressure) Rise
p Pressure
s Solid phase
stat Static pressure
T Temperature
tot Total strain tensor or total pressure
v Vapour
W Wall
w Liquid phase
x Strain tensor due to hygro-expansion
∞ Value outside the boundary layer in the free stream
1 Initial value

15.4 Mathematical operations

v Vectors are denoted by bold letters or as


vi in index notation.
v Matrices (order 2 tensors) are denoted by bold letters with double bar
vij or in index notation*.
f mean value of f
fɶ dimensionless state of f
df
Derivative of f with respect to x
dx
∂f
Partial derivative of f with respect to x
∂x
Df
Substantial derivative of f
Dt
*In index notation Einstein’s summation convention is applied.
About the book
The present book originates from the work of WG 2 (Fundamentals aspects of wood
drying) of COST action E15 (Advances in the drying of wood). It should be considered
as a the state-of-the-art of fundamental knowledge concerning the phenomena involved
in wood drying. The book contains 13 chapters :
1 - From fundamental to practice: the interaction chain
2 - Wood anatomy, an introduction
3 - Psychometrics
4 - Sorption isotherms in wood
5 - Shrinkage, swelling and warp caused by moisture changes
6 - Bound water migration in wood
7 - Fluid migration in wood
8 - Creep deformation in drying wood
9 - External heat and mass transfer
10 - Coupled heat and mass transfer
11 - Stress development
12 - Discoloration of wood during drying
13 - Airflow within kilns

Written by 14 European, 1 Australian and 1 Canadian experts, Fundamentals of Wood


Drying is a reference book whose major aim is to provide sufficient knowledge to
scientists and engineers to analyse and improve existing situations. The book also may
be a first step in the development of innovative procedures that hopefully will enable
drying practitioners to imagine, simulate, improve, and implement any new idea...

About the editor


Patrick Perré is Professor at ENGREF (National Engineering School of Forestry), Nancy,
where he teaches the physics and mechanics of trees and of wood. Vice-president of
ARBOLOR the association for wood science in Lorraine, France, and fellow of IAWS
(International Academy of Wood Science), his scientific skills include physics, anatomy,
mechanics, chemical engineering and applied mathematics. His research expertise
covers wood anatomy/wood properties and relationships, transfer in wood, wood
rheology, wood drying and the thermal treatment of wood. On this complex object of
study, he definitely considers that theory and practice must be combined. He also tries to
promote multidisciplinary approaches in his research. Author or co-author of 81 papers
in referred journals, 15 invited lectures with full text and 78 communications with full text
at international conferences, 10 book chapters and 3 patents, Patrick Perré was
awarded for the Best Paper in Drying technology Journal in 1996 and for Innovation in
Drying Research in 2004. A graduate of the prestigious Ecole Polytechnique which
collects the top engineering students throughout France, he received his PhD degree in
Physics from the University of Paris 7 in 1987 and the Habilitation à Diriger des
Recherches in Mechanical Engineering from INPL (Institut National Polytechnique de
Lorraine), Nancy, France, in 1992.

Printed in France ISBN :


Price : 45 € inclusive of tax

EUROPEAN COST A.R.BO.LOR.

You might also like