89
i Wettability of modified wood MAZIAR SEDIGHI MOGHADDAM Doctoral Thesis KTH Royal Institute of Technology Stockholm 2015

MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

i

Wettability of modified wood

MAZIAR SEDIGHI MOGHADDAM

Doctoral Thesis

KTH Royal Institute of Technology

Stockholm 2015

Page 2: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

ii

KTH Royal Institute of Technology

School of Chemical Science and Engineering

Department of Chemistry

Division of Surface and Corrosion Science

and

School of Architecture and the Built Environment

Department of Civil and Architectural Engineering

Division of Building Materials

SE-100 44, Stockholm, Sweden

TRITA-CHE Report 2015:56

ISSN 1654-1081

ISBN 978-91-7595-707-4

SP Chemistry, Materials and Surfaces Publication nr: A-3582

© Maziar Sedighi Moghaddam, 2015. All rights reserved. No part of this thesis may be

reproduces without permission from the author.

The following papers are printed with permission:

Paper I: Copyright © 2014 Springer

Paper II: Copyright © 2013 American Chemical Society

Paper III and V: Copyright © 2015 De Gruyter

Printer at US-AB, Stockholm, Sweden 2015

Akademisk avhandling som med tillstånd av KTH i Stockholm framlägges till offentlig

granskning för avläggande av teknologie doktorsexamen fredagen den 20 november kl.

10:00 i sal B3, KTH, Brinellvägen 23, Stockholm. Avhandlingen försvaras på engelska.

Page 3: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

iii

To my love, Golrokh

Page 4: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

iv

Page 5: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

v

Abstract

Despite many excellent properties of wood which make it suitable for

many applications, it suffers from a number of disadvantages limiting its

use. For instance, modification is needed to reduce water sorption and to

improve decay resistance, dimensional stability and weathering

performance. In addition, wood/liquid interaction such as water

wettability on wood plays an important role in design and characteristics

of many processes and phenomena such as adhesion, coating,

waterproofing, wood chemical modification, and weathering. This thesis

focuses on enhancing the understanding of wetting of wood, with

emphasis on modified wood. The influence of surface chemical

composition of wood and its microstructural characteristics on wetting

and swelling properties has also been studied.

A multicycle Wilhelmy plate technique has been developed to evaluate

wetting properties of porous materials, such as wood, in which the

samples were subjected to repeated immersions and withdrawals in a

swelling liquid (water) and in a non-swelling liquid (octane). This method

was utilized to dynamically investigate contact angle, sorption and

swelling properties, as well as dimensional stability of unmodified,

chemically and surface modified wood samples. Scots pine sapwood and

heartwood samples were utilized to establish the principles of the

technique. Acetylated and furfurylated wood samples with different level

of modification were thereafter examined utilizing the developed

technique for wetting measurements. A perimeter model based on a

linear combination of the measured force and final change in sample

perimeter was suggested to evaluate the dynamic dimensional stability of

wood veneers. The feasibility of this method for studying dynamic

wettability was investigated by measuring the changes of advancing and

receding contact angles over repeated cycles on surface modified wood

samples, created by combining liquid flame spray and plasma

polymerisation methods. X-ray photoelectron spectroscopy (XPS) and X-

ray computed tomography (XCT) were employed to study the surface

chemical composition and microstructural properties of the samples,

respectively.

Three different kinetic regimes were observed in the wetting

measurements: i) fast wetting and spreading of the liquid on the wood

surface, ii) void filling and wicking and iii) swelling, which was the

Page 6: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

vi

slowest of the three. The multicycle Wilhelmy plate method was found to

be suitable for studying liquid penetration, sorption, and dimensional

stability of swelling materials. The results demonstrate that the wetting

properties of wood are highly affected by surface chemistry and

microstructure. It was shown that using both swelling and non-swelling

liquids in wetting measurements allow to distinguish between capillary

liquid uptake and swelling. Based on this, for chemically modified

samples, it was demonstrated that acetylation mostly reduces swelling,

while furfurylation reduces both swelling and capillary uptake. This is in

line with the microstructural study with X-ray computed tomography

where a significant change in the porosity was found as a result of

furfurylation, conversely acetylation left the total porosity values

unchanged. Wetting results for hydrophobised wood samples

demonstrate that the multi-scale roughness obtained by combination of

nanoparticle coating and plasma polymerization increased both the

hydrophobicity and the forced wetting durability compared to the micro-

scale roughness found on wood modified with plasma polymerisation

alone.

Keywords Wood, dimensional stability, dynamic wettability, surface

chemistry, microstructure, swelling, multicycle Wilhelmy plate method,

contact angle, sorption, acetylation, furfurylation, surface modification

Page 7: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

vii

Sammanfattning

Trots att trä har många utmärkta egenskaper, vilket gör det lämpligt för

många tillämpningar, lider trä av ett antal nackdelar som begränsar dess

användningsområden. Trä måste därför modifieras för att förbättra dess

motstånd mot nedbrytning, dess dimensionsstabilitet och dess

väderbeständighet, samt för att minska vattenabsorption. Dessutom

spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig

roll när det gäller att utforma och utvärdera olika processer och fenomen,

såsom vidhäftning, ytbehandling, tätskikt, trämodifiering och

nedbrytning vid utomhusexponering. Denna avhandling fokuserar på att

få en bättre förståelse av vätningsegenskaper hos trä, med betoning på

modifierat trä, med hjälp av nyutvecklade tekniker. Inverkan av den

ytkemiska sammansättningen av trä och dess mikrostrukturella

egenskaper på vätning och svällningsegenskaper har också studerats.

En multicykel Wilhelmy-platta-metod har utvecklats för att undersöka

vätningsegenskaperna hos porösa material, såsom trä, där proverna

utsattes för upprepad nedsänkning i och upplyftning ur en svällande

vätska (vatten), samt i en icke-svällande vätska (oktan). Användbarheten

av denna metod undersöktes dynamiskt med avseende på kontaktvinkel,

absorption och svällningsegenskaper, men även dimensionsstabilitet för

omodifierade, samt kemiskt och ytmodifierade träprover. Splintved och

kärnvedsprover från tall användes för att fastställa principerna för

tekniken. Acetylerade och furfurylerade träprover med olika grad av

modifiering undersöktes därefter med den utvecklade tekniken för

vätningsmätningar. En perimetermodell baserad på linjär kombination av

den uppmätta kraften och den slutliga perimetern föreslås för utvärdering

av den dynamiska dimensionsstabiliteten hos trä. Den utvecklade

metoden för dynamisk vätbarhet undersöktes genom att mäta

förändringar i avancerande och återgående kontaktvinklar under

upprepade vätningscykler på små träfanér ytmodifierade genom

kombination av aerosolbaserad beläggning och plasmapolymerisation.

Röntgenfotoelektronspektroskopi (XPS) och röntgendatortomografi (CT)

användes för att studera den ytkemiska sammansättningen, samt

mikrostrukturen hos proverna.

Tre olika kinetiska regimer observerades i vätningsmätningarna: i)

snabb vätning och spridning av vätskan på träytan, ii) kapillärt

Page 8: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

viii

vätskeupptag och fuktspridning och iii) svällning, vilket också var det

långsammaste förloppet. Den utvecklade tekniken, mångcyklisk

Wilhelmy-platta, föreslås som en lämplig metod för att studera

vätskepenetration, absorption och dimensionsstabilitet hos porösa,

svällande material. Resultaten visar att de vätande egenskaperna hos trä

påverkas kraftigt av ytkemi och mikrostruktur. Vid användning av en

svällande och en icke-svällande vätska i mätningarna är det möjligt att

skilja mellan kapillärt upptag och svällning. Baserat på detta påvisades

det att acetyleringen reducerade svällning, medan furfuryleringen

minskade både svällning och kapillärt upptag. Detta är i linje med

mikrostrukturstudien med röntgendatortomografi som påvisade att

furfuryleringen resulterade i en betydande förändring i porositet medan

acetyleringen innebar en oförändrad porositet. Vätningsförsöken med de

hydrofobiserade träproverna visade att den flerskaliga ytråheten ökade

hydrofobiciteten, liksom motståndet mot forcerad vätning jämfört med

enbart ytråhet på mikroskala.

Nyckelord Trä, dimensionsstabilitet, dynamisk vätbarhet, ytkemi,

mikrostruktur, svällning, multicykel Wilhelmy-platta, kontaktvinkel,

absorption, acetylering, furfurylering, ytmodifiering

Page 9: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

ix

List of appended papers

This thesis is based on the following papers referred to in the text by their

Roman numbers:

I. Wettability and liquid sorption of wood investigated

by Wilhelmy plate method

Sedighi Moghaddam, M., Claesson, P.M., Wålinder, M.E.P.,

Swerin, A., Wood Science and Technology 2014. 48, 161-176.

II. Multicycle Wilhelmy Plate Method for Wetting

Properties, Swelling and Liquid Sorption of Wood

Sedighi Moghaddam, M., Wålinder, M.E.P., Claesson, P.M.,

Swerin, A., Langmuir 2013. 29, 12145-12153.

III. Wettability and swelling of acetylated and

furfurylated wood analyzed by multicycle Wilhelmy

plate method

Sedighi Moghaddam, M., Wålinder, M.E.P., Claesson, P.M.,

Swerin, A., Holzforschung 2015. doi:10.1515/hf-2014-0196

IV. Wetting hysteresis induced by temperature changes:

supercooled water on hydrophobic surfaces

Heydari, G., Sedighi Moghaddam, M., Tuominen, M.,

Fielden, M., Haapanen, J., Mäkelä, J.M., Claesson, P.M.,

Submitted

V. Hydrophobisation of wood surfaces by combining

liquid flame spray (LFS) and plasma treatment:

dynamic wetting properties

Sedighi Moghaddam, M., Heydari, G., Tuominen, M.,

Fielden, M., Haapanen, J., Mäkelä, J.M., Wålinder, M.E.P.,

Claesson, P.M., Swerin, A., Accepted in Holzforschung 2015.

doi:10.1515/hf-2015-0148

VI. Microstructure of chemically modified wood using

X-ray computed tomography scanning in relation to

wetting properties

Sedighi Moghaddam, M., Van den Bulcke, J., Wålinder,

M.E.P., Claesson, P.M., Van Acker, J., Swerin, A., Submitted

Page 10: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

x

The author´s contribution to the papers:

I. All the planning, experimental work, evaluation and major

part of writing.

II. All the planning, experimental work (except XPS),

evaluation and major part of writing.

III. All the planning, experimental work (except chemical

modification of wood samples), evaluation and major part of

writing.

IV. Minor part of planning, experimental work, evaluation and

writing.

V. Major part of planning, experimental work (except plasma

modification and XPS), evaluation and major part of writing.

VI. All the planning, major part of experimental work,

evaluation and writing.

This thesis also contains unpublished results.

Page 11: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xi

Summary of papers

Paper I

A set of methods based on the Wilhelmy plate technique was

introduced for measuring dynamic wettability, void filling and liquid

sorption into porous and swellable solids, such as wood. Different kinetic

regimes, the fastest one associated with contact angle change and the

slowest regime associated with liquid sorption by capillary and diffusion,

were observed. Comparing different approaches to investigate sorption

properties suggested the “imbibition to constant depth” method as the

preferred one, since contact angle relaxation effects can be ignored. A

multi cycle Wilhelmy plate method was developed to investigate the

dynamic wettability properties of wood veneers.

Paper II

The use of the method developed and described in Paper I (multi-cycle

Wilhelmy plate) was exemplified by studies of wood veneers of Scots pine

sapwood and heartwood to investigate wetting properties, swelling and

dimensional stability. The samples were subjected to repeated immersion

and withdrawal in a swelling liquid (water) and in a non-swelling liquid

(octane). A model based on a linear combination of the measured force

and final change in sample perimeter was suggested, and validated to

elucidate the dynamic perimeter change of wood veneer samples. The

data revealed lower contact angle, higher liquid uptake, higher swelling,

less dimensional stability and higher water contamination during the

measurement for pine sapwood than pine heartwood due to different

structure and chemical composition (level of extractives and lignin).

Paper III

The multicycle Wilhelmy plate method was utilized to compare

wetting, dimensional stability and sorption properties of chemically

modified wood samples, obtained either by acetylation or furfurylation.

By combining the results of water and octane uptake from wetting

measurements would be possible to distinguish between capillary liquid

uptake and swelling. It was found that acetylation reduces the water

uptake mainly by reducing the swelling. In contrast, furfurylation reduces

both swelling and the void volume in the sample. The results revealed

that furfurylated Southern yellow pine with weight percentage gain (a

Page 12: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xii

measure of the level of modification) of 45% had lower water and octane

uptake, lower swelling, and higher dimensional stability than other

samples.

Paper IV

The influence of temperature on the wetting states of hydrophobic

wood surfaces was studied utilizing contact angle measurements during a

freeze-thaw cycle. Introduction of multi-scale roughness on the wood

surfaces increased the initial contact angle and its stability during a

freeze-thaw cycle. The data elucidate that a multi-scale roughness reduces

the penetration of supercooled water into surface depressions, and also

enhances the freezing delay at low degrees of supercooling. A flat surface,

having similar surface chemistry as the modified wood surface, is

however more efficient in enhancing the freezing delay than the rougher

wood surface.

Paper V

A thin transparent superhydrophobic layer on Scots pine veneer

surfaces was achieved by combining plasma polymerisation and liquid

flame spray (LFS) for controlling surface wettability and water repellency.

The use of the multi-cycle Wilhelmy plate method for studying dynamic

wettability was exemplified by measuring the changes of advancing and

receding contact angles over repeated cycles on surface modified wood

samples. The effects of surface chemistry (using different plasma polymer

layers) and surface roughness (micro-scale and multi-scale rough

surfaces) on the wetting properties were investigated. Multi-scale rough

surfaces (micro-nano roughness) showed a contact angle greater than

150o with relatively low hysteresis, while the surfaces treated only with

plasma polymers revealed a contact angle of around 130o with high

hysteresis. Thus, the multi-scale roughness increased the hydrophobicity

as well as the forced wetting durability compared to micro-scale

roughness alone.

Paper VI

The wetting properties of chemically modified samples (Paper III)

were related to the microstructural properties as determined by using the

X-ray computed microtomography technique. The 2D and 3D

microstructure of acetylated and furfurylated softwood and hardwood

Page 13: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xiii

were visualized and some anatomical features were investigated such as

total porosity, cell wall thickness and maximum opening of tracheid

lumens. Significant changes in the wood structure were observed for

furfurylated sapwood samples mainly indicated by a change in tracheid

shape and filling of tracheids by furan polymer, whereas no

microstructural changes were noted for acetylated samples. Furfurylation

significantly decreased the porosity of the sample in both earlywood and

latewood regions; whereas for acetylated samples the total porosity of

modified and unmodified samples was rather similar. This is in line with

results of wetting showing that furfurylation reduced both swelling and

capillary uptake in contrast to acetylation which reduced mostly swelling.

Page 14: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xiv

Table 1: Summary of appended papers including materials, methods and evaluated properties

Paper I Paper II Paper III Paper IV Paper V Paper VI

Materials

Wetting liquids Water Octane

Water Octane

Water Octane

Water Water Water Octane

Substrate Unmodified wood, pine sapwood

Unmodified wood, Scots pine sapwood and heartwood

Acetylated and furfurylated wood, Unmodified wood, SYP, maple, beech

Surface modified wood and silica, nanoparticles and plasma polymer treated samples

Surface modified wood and silica, nanoparticles and plasma polymer treated samples

Acetylated and furfurylated wood, Unmodified wood, SYP, maple

Methods

Wetting methods Single cycle Wilhelmy plate Imbibition at constant depth Immersion to constant depth (without withdrawal) Full immersion Multicycle Wilhelmy plate

Multicycle Wilhelmy plate

Multicycle Wilhelmy plate

Sessile drop method, freezing measurement

Multicycle Wilhelmy plate

Multicycle Wilhelmy plate

Characterisation methods

XPS XPS, AFM XPS, AFM, SEM XCT

Evaluated properties and proposed models and mechanisms

- CA - Dynamic sorption

and swelling - Different kinetic

regimes of wetting

- CA, - Dynamic

sorption and swelling

- Dimensional stability

- Wetting-Surface chemistry relationship

- CA - Dynamic

sorption - Dimensional

stability - Swelling and

capillary uptake

- CA as a function of temperature

- Freezing delay - Surface chemical

composition - Surface

topography

- Dynamic advancing and receding CA

- Dynamic sorption - Surface chemical

composition - Surface

topography

- Swelling and capillary uptake

- 2D and 3D visualisation

- Quantitative measurement of anatomical features

- Wetting-microstructure relationship

Page 15: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xv

Nomenclature

Abbreviations

2D Two dimensional

3D Three dimensional

Acet Acetylated sample

ADSA Axisymmetric drop shape analysis

AFM Atomic force microscopy

ATR-IR Attenuated total reflectance infrared

Avg Average

CA Contact angle

CAH Contact angle hysteresis

CCD Charge-coupled device

CF4 Tetrafluoromethane

CT Computed tomography

CTRL Control

CWT Cell wall thickness

EW Earlywood

F Fiber

FA Furfuryl alcohol

Furf Furfurylated sample

h Hour

HMDSO Hexamethyldisiloxane

L Longitudinal

LFS Liquid flame spray

LW Latewood

min Minute

mm Millimeter

nm Nanometer

NMR Nuclear magnetic resonance

PFH Perfluorohexane

POTS 1H, 1H, 2H, 2H- perfluoroalkyltriethoxysilanes

R Radial

Ra Ray cells

s Second

SD Standard deviation

SEM Scanning electron microscope

SYP Southern yellow pine

Page 16: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xvi

T Tangential

Tr Tracheid

UV Ultraviolet

V Vessel

WPG Weight percentage gain

XCT X-ray computed tomography

XPS X-ray photoelectron spectroscopy

Symbols

A Area [m2] F Force [N] FA Advancing force [N] FA,n Advancing force at cycle n [N] Ff Final force [N] Ff,n Final force at cycle n [N] ∆𝐹𝑓 Final force difference between last cycle and first cycle [N]

FR Receding force [N] FR,n Receding force at cycle n [N] Fw(t) Time-dependent force [N] g Gravitational acceleration [m/s2] h Immersion depth [m] ln Change in liquid mass uptake at cycle n [%] loct,10 Final octane uptake at cycle 1o [%] lw,20 Final water uptake at cycle 2o [%] ∆𝑙20 Wetting term for calculating the swelling [%] mi Initial mass [kg] P Perimeter [m] Pn Veneer perimeter after cycle n [m] ∆𝑃 Total perimeter change [m] Wa Work of adhesion [J/m2] θ Contact angle [o] θA Advancing contact angle [o] θA,n Advancing contact angle at cycle n [o] θR Receding contact angle [o] θR,n Receding contact angle at cycle n [o] ρ Density [kg/m3] ρoct Density of octane [kg/m3] ρw Density of water [kg/m3] 𝛾 Surface tension [N/m] 𝛾𝑆 Surface free energy of solid [J/m2]

Page 17: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xvii

𝛾𝐿 Surface free energy of liquid [J/m2] 𝛾𝐿𝐺 Surface tension of liquid-gas [N/m] 𝛾𝑆𝐺 Surface tension of solid-gas [N/m] 𝛾𝑆𝐿 Surface tension of solid-liquid [N/m]

Page 18: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xviii

Page 19: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xix

Table of Contents

Abstract........................................................................................ v

Sammanfattning ........................................................................ vii

List of appended papers ............................................................ ix

Summary of papers .................................................................... xi

Nomenclature ............................................................................ xv

Table of contents ...................................................................... xix

1. Introduction ............................................................................. 1 1.1. Aims and Objectives ............................................................................ 1 1.2. Wood structure and chemistry ......................................................... 2 1.3. Wetting ................................................................................................... 6 1.3.1. Basic concepts ......................................................................................... 6

1.3.2. Wood wettability measurements ............................................................... 8

The importance of wood wettability ................................................................. 8

Measurement techniques ................................................................................ 9

Challenges in wood wettability measurements .............................................. 10

Wettability studies on wood using Wilhelmy plate method ............................ 11

1.4. Wood modification ............................................................................. 12

1.4.1. Chemical modification ............................................................................ 13

Acetylation ..................................................................................................... 14

Furfurylation ................................................................................................... 14

1.4.2. Surface modification ............................................................................... 15

2. Experimental .......................................................................... 17

2.1. Wilhelmy plate method ....................................................................... 17

2.1.1. Contact angle detemination .................................................................... 18

2.1.2. Multicycle Wilhelmy plate method .......................................................... 18

2.1.3. Other wetting methods ........................................................................... 19

2.2. Sessile drop method ........................................................................... 20

2.3. X-ray photoelectron spectroscopy..................................................... 21

2.4. X-ray computed tomography .............................................................. 22

2.5. Atomic force microscopy .................................................................... 22

2.6. Materials ............................................................................................... 23

2.6.1. Wood samples ........................................................................................ 23

2.6.2. Wetting liquids ........................................................................................ 24

3. Results and discussion......................................................... 25

3.1. Comparison of Wilhelmy and sessile drop methods ...................... 25

3.2. Developed techniques for wettability study of wood ...................... 27

3.2.1. Immersion techniques ............................................................................ 27

Page 20: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

xx

3.2.2. Multicycle Wilhelmy plate method ........................................................... 29

3.3. Applications of multicycle Wilhelmy plate method ......................... 31

3.3.1. Dynamic contact angle ........................................................................... 31

Contact angle of chemically modified wood ................................................... 32

Contact angle of surface modified wood ........................................................ 33

3.3.2. Dynamic swelling and liquid sorption ...................................................... 34

Unmodified wood ........................................................................................... 35

Chemically and surface modified wood.......................................................... 36

3.3.3. Dimensional stability ............................................................................... 38

3.3.4. Dissolution of extractives ........................................................................ 41

3.4. Factors affecting wood wetting properties....................................... 42

3.4.1. Surface chemical composition ................................................................ 42

Scots pine sapwood and heartwood .............................................................. 43

Surface modified wood samples .................................................................... 46

3.4.2. Microstructure ......................................................................................... 47

Unmodified pine sapwood and heartwood ..................................................... 47

Chemically modified wood ............................................................................. 48

3.4.3. Temperature effects of wetting properties .............................................. 54

4. Conclusions .......................................................................... 57

5. Future work ........................................................................... 59

6. Acknowledgements .............................................................. 61

References ................................................................................ 63

Page 21: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 1

1. Introduction

1.1. Aims and Objectives

The wood/liquid interaction has a critical role in many wood applications

and it highly impacts the wood performance in various conditions. Better

understanding of wood wettability may help to improve the design of

different wood modifications, as well as to predict the wood

characteristics such as adhesion, coating properties, waterproofing, and

biodegradation properties.

The aim of this thesis is to gain more insight about wood wetting

properties with emphasis on modified wood. The specific objective is to

establish a link between wetting properties and surface chemical

composition and microstructural properties of wood. The study also

focuses on establishment of some novel techniques based on the

Wilhelmy plate method to study dynamic wetting properties of modified

wood samples. It will highlight phenomena such as dynamic contact

angle, dynamic sorption and swelling, dimensional stability and

dissolution of extractives on unmodified, chemically modified and surface

modified wood samples. The study relates the capillary/swelling

properties with microstructural characteristics with emphasis on

acetylated and furfurylated wood. It also investigates how to create a

multi-scale rough superhydrophobic wood surface by combining an

aerosol method (liquid flame spray) and a plasma polymerisation

technique, and the dynamic wetting properties of these surface modified

wood samples.

Page 22: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

2 | INTRODUCTION

1.2. Wood structure and chemistry

Wood, also known as secondary xylem, is a porous three dimensional,

hygroscopic, viscoelastic, anisotropic biopolymer composite material,

composed of polymers and chemicals formed in small microstructural

elements. There is a large variability in the properties of wood which can

be explained by its complex, heterogeneous and varying chemical

composition, micro- and macroscopic structure [1-3]. Three structurally

different directions can be defined for wood: (L) the longitudinal

direction, along the grain or the fiber direction, (R) the radial direction,

across the grain and perpendicular to the annual rings, and (T) the

tangential direction, across the grain and tangential to the annual rings.

These directions also define the cross, radial and tangential sections (see

Figure 1-1).

Figure 1-1: 3-D photomicrograph of the structure of a softwood/Southern yellow pine). Three principal directions and sections of wood are shown. The image is from own work.

Wood can be divided into the two main classes: 1) softwoods

(gymnosperms) such as pine and spruce; and 2) hardwoods

(angiosperms) such as maple and beech (see Figure 1-2). Softwood and

hardwood are different in terms of their xylem structure and cell types

[3]. Softwoods have a simpler anatomy than hardwoods and they are

typically composed of axially extended tracheids (90-95%) and radially

Page 23: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 3

extended ray cells (5-10%) [4]. The role of tracheids is mechanical

support and longitudinal water transport (conduction) to the top of the

tree for the photosynthesis, while ray cells enable radial liquid transport.

Conversely, hardwoods have more advanced and complex structure that

contain vessels, fibers (libriform fibers and fiber tracheids) and

parenchyma (longitudinal oriented and ray cells). The vessels (also

known as “pores”) are responsible for water conduction with the

mechanical support of the fibers, and parenchymas mainly store

carbohydrates. Furthermore, the water transport between the wood cells

is enabled by so-called pits in the cell wall, in softwood often in the form

of so-called boarded pits, functioning similar to a “valve”. Hardwoods

generally differ in terms of the proportion of different component cells.

For instance, maple (Acer platanoides) contains 7% vessels, 76% fibers

and 17% rays with very low share of longitudinal parenchyma, while

beech (Fagus sylvatica L.) has 39% vessels, 40% fibers, 16% rays and 5%

longitudinal parenchyma [5].

Figure 1-2: Microstructural images of different sections of a softwood sample (southern yellow pine) (a) transverse, (b) radial, (c) tangential sections; and of a hardwood sample (maple): (d) transverse, (e) radial, (f) tangential sections. For softwood structural components are tracheids (Tr) and ray cells (Ra), while a hardwood consist of vessels (V), fibers (F) and rays (Ra). The scale bars correspond to 100 µm. The images are from own work.

Page 24: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

4 | INTRODUCTION

From the aspects of the growing season, wood is macroscopically

composed of earlywood (EW) (lighter) and latewood (LW) (darker).

Typically, tracheids in EW have a considerably higher porosity, i.e. they

have thinner cell walls and wider lumen than tracheids in LW.

Furthermore, the main function of longitudinal tracheids in EW is the

conduction of water and nutrients, while tracheids in LW have

mechanical and supportive function [4].

In addition, the wood xylem can be divided into sapwood and

heartwood. The former, which is normally lighter, contains living and

dead cells. Heartwood, which is the inner layers of wood and has

generally darker colour, entirely contains dead cells. The main function of

sapwood is transport of liquid and nutrients and of the heartwood to

provide mechanical support. [5]. The heartwood is generally less

permeable and more durable, since extractives are deposited in the cells

during the wood formation [4]. Therefore, heartwood generally contains a

higher amount of extractives than sapwood. Additionally, the type of

extractives can differ in sapwood and heartwood. For instance in Scots

pine, heartwood contains more resin acids, fatty acids and pinosylvin,

while sapwood contains higher amounts of triglycerides [6-7]. It is well

understood that more bordered pits are aspirated (i.e. closed) in

heartwood in softwoods, and in some hardwoods the vessels are blocked

by tyloses [4]. These differences in extractives level and microstructure

for sap- and heartwood, may lead to a lower and slower liquid sorption in

heartwood compared to sapwood.

Wood is chemically composed of macromolecular compounds

(cellulose, hemicellulose and lignin) and components with low molecular

weight (extractives) [3]. Macromolecular compounds are located in the

cell walls and known as the structural components, and nonstructural

components have low molecular weight (extractives). The cell wall can be

considered as a reinforced composite in which the cellulose microfibrils

act as the reinforcements, embedded in a matrix of lignin and

hemicelluloses. In general wood species contain about 40-50% cellulose,

20-30% hemicellulose, 20-35% lignin and up to 10% extractives [8].

Softwoods normally contain more cellulose and lignin than hardwoods

[3].

Cellulose and hemicelluloses are polysaccharides. Cellulose as the

main molecular component of wood, is a linear polymer of D-

anhydroglucopyranosyl residues linked via β(1→4) glycosidic bonds [3].

Page 25: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 5

Several number of cellulose molecules can closely connect by hydrogen

bonds and form highly crystalline microfibrils as the reinforcing elements

in the cell wall, causing inertness and stability of the cellulose [9]. The

degree of polymerisation for cellulose varies from 2000-4000 in the

primary layer of the cell wall to 10000 in the secondary level [8].

Hemicelluloses, unlike cellulose, are heteropolysaccharides (i.e. they

contain several types of saccharides) with lower molecular weight (degree

of polymerisation of 150-200). They are normally branched amorphous

molecules. The hemicelluloses are located in the cell wall matrix between

the cellulose microfibrils and together with the cellulose, and they have

structural supportive function in the wood. They are also believed to

function as a link, or compatibilizer, between cellulose and lignin [3].

Cellulose and hemicelluloses in wood contain large amounts of hydroxyl

groups (OH). Polar liquids such as water can build hydrogen bonds with

the OH groups of the cell walls, causing swelling of the cell walls. This is

the main reason for the hygroscopic nature of wood.

Lignin is an amorphous aromatic polymer with a complex structure

consisting of various phenylpropane units connected by either ether (C-

O-C) or carbon-carbon (C-C) bonds in a 3-D network. Unlike cellulose,

lignin does not have any specific and single repeating unit (monomer)

and is built based on the polymerisation of three different monomers

called coumaryl, coniferyl and sinapyl alcohols. Softwoods mainly contain

coniferyl alcohol (G-lignin), while hardwood lignins usually are

copolymers of coniferyl and synapyl alcohols (SG-lignin) [4,10]. Lignin

offers several functions in the wood structure including fixation of

cellulose and hemicellulose to each other and improving the cell wall

stiffness [4], acting as an adhesive and hold different cells together [11]

and providing a certain degree of protection against microbiological

degradation by making the cell wall more hydrophobic [10]. Lignin also

decreases the influence of water sorption on the stiffness and strength of

tracheids [12]. Extractives are nonstructural compounds present in the wood

structure normally present in quite low concentrations, for softwood

typically 1.5-5 % [13]. On the other hand, they are an important source of

the variety in wood chemical composition. In particular, extractives alter

the surface chemistry of wood [14] due to their natural mobility and

migration to the surface regions, for instance due to kiln drying and also

after woodworking processes. Extractives include a large number of

Page 26: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

6 | INTRODUCTION

individual substances and are classified in different ways. Based on the

type of solvent which the extractives are soluble in, they can be classified

as lipophilic (e.g. terpenes, waxes, triglycerides (fats), steryl esters and

sterols) and hydrophilic (carbohydrates and phenolic compounds).

Lipophilic extractives (i.e. wood resin) are the main source of the wood

surface inactivation, since they can migrate to the surface and make it

more hydrophobic. Additionally, based on the function in the wood

structure, extractives can be divided into pathological resin (resin acids

and monoterpenes) and physiological resin (fats). The main function of

the first type, which is located in resin canals, is to preserve the wood

against microbiological degradation. Conversely, the second type is

located in the parenchyma cells and acts as a supply of reserve

nourishment [15].

The level and types of extractives differ both within and between

species, and also within the same tree [15]. It is understood that

extracting the wood species could lead to increasing rate of swelling [16].

According to Stamm and Loughborough [17] wood extractives can be

classified into two main groups in terms of their location: 1) extractives

located in the capillary structure, and 2) extractives located in the cell

wall structure. Obviously, the first group affects the swelling-specific

density with no influence on overall wood swelling. The second

extractives type can affect the cell wall swelling and as a result cause

dimensional changes [18]. Therefore, not only level and type of

extractives, but also their location, and migration in the wood structure

influence the swelling and wettability properties of wood.

1.3. Wetting

1.3.1. Basic concepts

The term wetting can be defined as “macroscopic manifestations of

molecular interaction between liquids and solids in direct contact at the

interface between them” [19]. The wetting process includes the formation

of a contact angle at the solid/liquid/gaseous interface and the spreading

of the liquid on the solid surface. This spreading continues until

equilibrium contact radius is reached. For a porous material, capillary

forces determined also by wetting may result in wicking of the liquid. It is

important to bear in mind that for swellable materials such as wood, the

Page 27: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 7

swelling process would also occur at the same time as the wetting which is

not considered in the above definition.

A solid is said to swell when it absorbs a liquid under the following

conditions [16]: (i) its dimensions are increased, (ii) its microscopic

homogeneity remains, and (iii) its cohesion is decreased but not

eliminated, and as a result it becomes soft and flexible instead of hard

and brittle. The liquids that wet the solid can be divided into two classes;

swelling liquids and non-swelling liquids. For wood, swelling liquids are

polar liquids that can absorb into the porous wood by both void filling

(capillary uptake) and absorption in the cell wall causing swelling,

whereas non-swelling liquids are non-polar and only fill the voids in the

wood structure.

Wetting may be characterized by measuring wettability parameters

such as equilibrium contact angle, an indicator of the affinity of a liquid

for a solid [20]. In the classical case, when a water droplet is placed on

smooth and non-porous solid surface, the contact angle θ (in an

equilibrium state condition) is defined as the angle between the tangent

to the liquid surface and the liquid/solid interface at the point of

liquid/solid contact. This is typically expressed by Young’s equation [21]:

𝛾𝑆𝐺 = 𝛾𝑆𝐿 + 𝛾𝐿𝐺 cos 𝜃 (1-1)

where 𝛾𝑆𝐺, 𝛾𝑆𝐿 and 𝛾𝐿𝐺 are surface tension of the solid-gas, solid-liquid

and liquid-gas, respectively. At constant temperature and pressure, a

single value of the contact angle (CA) of a liquid on a smooth surface is

expected, since it is a thermodynamic property. Nevertheless, almost all

technical surfaces have a higher so-called advancing than receding

contact angle, mainly due to the surface roughness and surface

heterogeneity. The advancing contact angle can be determined when a

liquid boundary advances over a surface and the receding contact angle

can be determined when a liquid boundary recedes from a previously

wetted surface. The difference between these two contact angles is

defined as the contact angle hysteresis. Several factors have been

suggested for the source of the hysteresis, including surface roughness

[22,23], surface chemical heterogeneity [24,25], molecular scale

topography and rigidity of molecules [26], packing of molecular chains on

the solid surface [27], reorientation of molecules and functional groups

Page 28: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

8 | INTRODUCTION

[28,29], liquid penetration and swelling [30,31] and liquid

sorption/retention on the surface [32].

It is also worth to note that Young’s equation is only valid for an

ideally smooth, homogeneous and clean surface without any liquid

sorption and resulting swelling of the substrate. For complex substrates

such as wood, the surface chemical heterogeneity and surface roughness

cause several difficulties in assessment of wettability.

1.3.2. Wood wettability measurements

The importance of wood wettability

Wood wetting properties are fundamental and key factors for designing

and evaluating of many technical and in-service related processes and

phenomena including adhesion, coating, wetting and waterproofing,

surface modification, weathering, wood modification and liquid

penetration [5,33]. For wood adhesives and glues, it was found that there

is a positive relation between glue-bond quality and wood wetting

properties. The water CA was suggested as being the most important

property controlling penetration of adhesives into wood [34].

Additionally, the work of adhesion is commonly calculated from CA

measurements (Eq. 1-2). For instance, Acda et al. [35] studied the effect

of plasma modification on the adhesion properties of wood surfaces. The

work of adhesion is defined as the work required to separate unit area of

the solid-liquid interface and is expressed as:

𝑊𝑎 = 𝛾𝑆 + 𝛾𝐿 − 𝛾𝑆𝐿 (1-2)

where 𝛾𝑆 and 𝛾𝐿 are the surface free energies of the solid and liquid,

respectively.

In the case of wood coatings, CA is also considered as an important

parameter for determining the coating flowability, penetration and

spreading capacity on the wood surfaces. These properties influence the

coating performance significantly [36]. Changes in the wetting properties

are generally considered as a major indicator for wood weathering tests in

both simulated and outdoor conditions [37]. Furthermore, the CA data of

polar and non-polar liquids are the raw data for determination of surface

free energy of wood surfaces [38]. Generally, surface free energies are

further utilized to investigate e.g. enhancement of adhesion strength in

Page 29: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 9

wood coating systems or control of the hydrophobicity of the wood

surfaces. Surface modification of wood can affect its wettability toward

more hydrophilic or hydrophobic for various applications and purposes

such as self-cleaning, increased resistance against UV light degradation

and weathering, and compatibility with adhesives and coatings [39]. For

these surface modified wood samples, the water CA measurement is the

most direct and important technique to study the modification

performance. Finally, for various wood modification concepts, the

wettability determination was found to be a key technique for direct

investigation of the properties of modified wood samples. Contact angle

measurements have been also utilized in investigations related to wood

modification including chemical modification [40,41], thermal

modification [41-43] and compression densification [44].

Measurement techniques

Generally, there are different techniques for measuring CAs [45]. The

sessile drop method is the most traditional and widely used one [46]. The

technique is based on the direct measurement of the angle by monitoring

of a drop resting on a solid surface [47]. Other techniques have also been

used for determination of CA for wood surfaces including the Wilhelmy

plate method [45, 48,49], inclined plate method [50,51], axisymmetric

drop shape analysis (ADSA) [52], captive bubble method [53], capillary

rise on a vertical plate [45] and column or thin layer wicking [54].

However, the sessile drop and Wilhelmy methods are the most used

techniques for wood surfaces. It is well understood that the sessile drop

method is very useful for homogeneous and smooth surfaces, whereas

wetting phenomena on wood surfaces are much more complex than many

other materials due to the roughness and heterogeneous chemistry.

Additionally, the hygroscopic character of wood and the swelling effects

when using polar probe liquids make the wetting measurement even

more difficult on wood [55]. Therefore, in many reports the Wilhelmy

method was considered as being more accurate than the sessile drop

method, for determination of wetting properties, especially for rough,

heterogeneous and hygroscopic materials such as wood for several

reasons [48, 56-58]. First, in the Wilhelmy method a sensitive

microbalance is utilized for measuring the forces which is free from

operator error. Second, a larger area is covered in the Wilhelmy method

compared with the sessile drop method, which enables a higher

Page 30: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

10 | INTRODUCTION

reproducibility and also solves measurement difficulties related to surface

chemical heterogeneity. In other words, by averaging the wetting

properties of a large area of the wood surface, a more reliable

characterisation of dynamic wetting and sorption can be achieved.

Additionally, by accurate control of advancing and receding rates, more

reliable characterisation of dynamic effects during wetting and dewetting

can be achieved. Finally, for hygroscopic materials it is possible to

measure the level of liquid sorption during the wetting measurement.

Apart from these advantages of the Wilhelmy plate technique, there are

some uncertainties in the literatures about the actual time of selecting the

data for determination of a single value of the CA, in the sessile drop

method [59-65], using anything from 1 s [59] to 60 s [65] after drop

contact as the correct time for CA measurement. Additionally, the shape

of a sessile drop of water on a wood surface is affected by gravity,

contamination from water soluble extractives, local surface anisotropy

and microstructure (e.g. fiber direction). Consequently, a non-spherical

shape is normally formed, leading to errors in determination of dynamic

volume and CA changes [5]. However, sessile drop method is still widely

used to study wetting properties of wood surfaces in spite of these

disadvantages [40, 42, 63-66]. In this thesis, the Wilhelmy plate method

was utilized in Papers I-III, V and VI to investigate wetting properties of

wood.

Challenges in wood wettability measurements

Some of the factors that may technically affect wettability measurements

of wood are shown in Figure 1-3. As can be seen, several factors can

influence the wetting measurement and make it much more complex,

including roughness, polarity, heterogeneity and porosity, wood grain

direction [67], and drying method [68].

Contact angles on wood surfaces normally have large hysteresis as a

result of the chemically heterogeneous and highly rough surfaces [24].

Moreover, the wetting process might be accompanied by liquid sorption

and swelling effects during wettability measurements because of the

hygroscopic nature of the wood [69]. It is understood that some liquids

such as water can easily penetrate into the wood bulk with the resulting

release of heat and swelling during the wetting experiments, thus a non-

equilibrium situation may result in the system [70]. In addition, wood

extractives can contaminate the probe liquids during the wetting

Page 31: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 11

measurement by migration to the wood surface. This contamination may

cause a significant change in surface tension of the probe liquid that if not

corrected for, propagates into an error in the contact angle results [70-

72].

Figure 1-3: Factors that can affect the wetting properties of wood.

Wettability studies on wood using the Wilhelmy plate method

The Wilhelmy method has been found to be a very useful technique for

investigating wettability of wood samples, and it has been utilized to

examine different wetting properties such as CA [41,43,48,49,68], surface

energy [38,73], work of adhesion [41], and dynamic wetting behavior

[74]. For instance [48,72], a modified Wilhelmy method to measure the

contact angle of wood samples and study the wettability characteristics of

wood has been suggested. The contamination effect of probe liquids by

extractives was investigated, and the CA results of extracted and non-

extracted samples were compared [72]. It was found that extractives

could migrate to the probe liquid during measurement and decrease the

surface tension. This contamination problem was less for aged samples

than for fresh ones. In another study [43] the effect of heat treatment on

wetting properties of different wood samples was examined by means of

contact angle measurement and it was found that the wood

hydrophobicity increased significantly by heat treatment. The Wilhelmy

plate method was also utilized to monitor the perimeter changes of

Page 32: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

12 | INTRODUCTION

chemically treated wood samples before and after 24 hours immersion in

water, in order to study the dimensional stability of thin wood veneers

[74]. Based on these findings it was suggested that the Wilhelmy method

is a more precise tool than digital calipers for investigating the

dimensional stability of small, thin or nonrectangular samples such as

wood veneers since it measures perimeter changes on a microscopic level.

The wetting properties of modified wood (acetylated, furfurylated and

thermally modified), fresh and aged, have also been studied using the

Wilhelmy method to predict work of adhesion between modified wood

materials and some thermoplastics and adhesives [41]. The effect of

drying conditions on wettability was investigated by measuring advancing

and receding contact angles of veneers dried according to different

methods [68]. In this study the veneers were immersed into the liquids

for 100 s at a depth of 4 mm to investigate liquid absorption. Surface

energy of wood samples were also determined using this technique

[38,75], and it was demonstrated that the surface energy and acid-base

properties for pine wood determined by the sessile drop method and the

Wilhelmy method were significantly different [73]. In this thesis, the

application of a modified Wilhelmy plate method (multicycle Wilhelmy)

was examined to measure CA (Paper I-III), dynamic advancing and

receding CA (Paper V), dynamic wettability and swelling (Paper I-III, V,

VI) and dimensional stability (Paper II,III) of different wood samples.

1.4. Wood modification

During various in-service uses of wood, e.g. in building material

applications, a change of its moisture content causes problematic swelling

or shrinkage, i.e. dimensional changes. The dimensional stability of wood

can be improved by reducing its hygroscopicity by applying certain wood

modification methods. Wood modification can be defined as”… the action

of a chemical, biological or physical agent upon a material, resulting in a

desired property enhancement during the service life of the modified

wood” [9]. Despite the excellent properties of wood such as a good

strength to weight ratio, aesthetic appearance etc., wood suffers from a

number of disadvantages that restricts it from many potential uses. For

instance, in outdoor conditions wood is sensitive to microbial attack, UV

light and has not sufficient dimensional stability [76]. Furthermore, when

wood is subjected to high moisture contents for longer times it is very

Page 33: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 13

sensitive to decay by fungi attack, consequently limiting this exposure

leads to a longer performance time [77]. In summary, wood is modified in

order to improve decay resistance and dimensional stability, to reduce

water sorption and to improve weathering performance [9]. Due to the

anatomical and chemical variation between wood species, and since

different properties are targeted in wood modification, it is not possible to

design a single wood modification system for all wood species and

applications. There is a range of possible wood modification methods

such as chemical, thermal, impregnation/polymerisation and surface

modification.

1.4.1. Chemical modification

Chemical modification can specifically be defined as [3]: “… any chemical

reaction between some reactive part of the wood cell wall component and

simple single chemical reagent, with or without catalysis that forms a

stable covalent bond.” The hydroxyl groups are mostly used as the

reaction points since they are the most abundant single site for reactivity

in the complex structure of wood. As a result of this reaction, a critical

alternation of the chemical and microstructure characteristics of the

modified wood may be obtained [3]. First, due to the substitution of

hydrophilic hydroxyl group with an added hydrophobic reagent, the

hygroscopic nature of wood is reduced. In other words, as a result of the

chemical modification some of the water in the cell wall would be

substituted with a more hydrophobic species, and even allow some

swelling in the material in order to reduce dimensional change. This leads

to a reduction in further swelling when the modified wood is exposed to

water or moisture, consequently the dimensional stability will be

enhanced [78]. Second, the wood volume increases due to reaction with

chemicals. In addition, biological resistance and thermal properties of

wood might be improved by the chemical modification [3].

Several chemical modification systems have been developed to modify

the properties of wood such as acetylation [79], cross-linking of cellulose

molecules with formaldehyde [80] and furfurylation [81]. Since

acetylated and furfurylated samples were investigated in Paper III and VI,

in the following, they are described in more details.

Page 34: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

14 | INTRODUCTION

Acetylation

Acetylation is the most well studied chemical wood modifications

procedure [9,82]. In the acetylation process an anhydride, usually acetic

anhydride, reacts with OH groups in wood resulting in acetyl groups in

the wood (see Figure 1-4). This reaction results in esterification of

accessible hydroxyl groups in the cell wall [9]. The acetylation is

considered as a single site reaction since one hydroxyl group is replaced

by one acetyl group with no polymerisation [3]. The by-product of the

process, acetic acid, can be converted to the acetic anhydride again. Since

acetyl groups are more hydrophobic than hydroxyl groups, the wettability

reduces resulting in an enhancement in dimensional stability. Moreover,

due to the permanent bulking effect on the wood cell wall, further

swelling by water or moisture is highly decreased.

Figure 1-4: Acetylation of wood with acetic anhydride.

As a result of the acetylation the durability of wood is improved to

class 1, swelling is reduced (i.e. enhancement in dimensional stability) up

to 70-80% compared to unmodified wood, and the hardness and UV-

resistance are also increased.

Furfurylation

The wood furfurylation process is based on the reaction between a bio-

based polar chemical, furfuryl alcohol (FA), with the wood cell wall. FA

can swell the cell wall due to a pressure impregnation process and in situ

polymerize inside the wood structure [81]. As a result a complex polymer

system is formed inside the wood structure. Unlike the acetylation, where

it is well understood that ester bonds form between the acetic anhydride

and the hydroxyl groups [78], there are some uncertainty about the

existence of any chemical bonds between poly furfuryl alcohol and the

Page 35: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

INTRODUCTION | 15

wood polymers in the cell wall for the furfurylation. Based on an ATR-IR

spectroscopy study, no proof for this was found [83]. On the other hand,

Nordstierna et al. suggested that lignin is the main component in wood

that is most likely to have chemical bonds with FA based on an NMR

study on the model compounds similar to lignin [84].

The chemistry of the furfurylation is quite complex and is not fully

understood. When this acid-catalyzed polymerisation occurs inside the

wood structure the reaction is even more complex. However it is

suggested that a highly branched and cross-linked furan polymer is

formed in wood during the furfurylation. These reactions can be

summarized as [85]:

1- Homopolymerisation of FA 2- Copolymerisation of FA and additives or wood extractive

substances 3- Grafting of FA or furan polymer to wood cell wall polymers (most

probably lignin)

Furfurylated wood samples have enhanced properties compared to

unmodified wood such as improved dimensional stability (for high

modification levels the anti-shrink efficiency is as high as 80%), improved

biological resistance and increased hardness as a result of polymer in the

cell lumens [85].

1.4.2. Surface modification

Surface modification of wood is a so-called passive method that improves

wood properties by treatment of thin layers on the wood surfaces [39].

The target properties for surface modification can be very different,

including improving gluing and adhesion of surface coatings, wettability

and weathering resistance, fire retardancy, mechanical properties and

resistance to wood pests. Most commonly, surface modifications are done

either to improve wettability and coating adhesion or to reduce

wettability to have waterproof surface [86]. A well-known approach to

decrease wettability in wood and therefore increase dimensional stability

is to hydrophobise the wood surface by lowering the surface energy and,

in recent cases, by modifying the surface roughness. Different approaches

can be used to create (super)hydrophobic wood surfaces such as sol-gel

method [87], plasma etching and polymerisation [88,89], hydrothermal

treatment [90], covalent grafting [91] and deposition of nanoparticles

Page 36: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

16 | INTRODUCTION

[92]. For example, Wang et al. [87] made superhydrophobic wood

surfaces by deposition of silica nanoparticles followed by modification of

these with POTS (1H, 1H, 2H, 2H- perfluoroalkyltriethoxysilanes). In

another study [90], a cosolvent hydrothermal method was utilized to

deposit zinc oxide nanorod arrays on wood surfaces. In addition, cold

plasma treatment as a simple, dry and low energy consumption method

has been used in many studies to deposit the polymer of a wide range of

monomers on wood surfaces. It is well-known that the plasma

polymerisation method only affects the outer layer of the wood surfaces

and that it is possible to control the wood wettability by choosing

different hydrophobic monomers [92]. Fluorine-containing (e.g. CF4)

and silicone-containing compounds (e.g. HMDSO) are the most used

monomers to be polymerized by plasma in order to create a thin

hydrophobic layer on wood surfaces [88]. In this study, surface

modification of wood - by liquid flame spray deposition followed by

plasma polymerization - was investigated in Paper IV and V.

Page 37: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

EXPERIMENTAL | 17

2. Experimental

2.1. Wilhelmy plate method

The Wilhelmy plate method is the most important technique for this

thesis, and it is used in all papers except in Paper ΙV. The Wilhelmy

method is based on measuring the force F acting on a plate immersed in a

liquid according to [93]:

𝐹(ℎ) = 𝑃𝛾cos𝜃 − 𝜌𝐴ℎ𝑔 (2-1)

where F is the detected force, P the wetted perimeter of the plate, γ the

surface tension of the probe liquid, θ the liquid-solid-air contact angle, ρ

the probe liquid density, A the cross-sectional area of the plate, h the

immersion depth, and g the gravitational constant. A schematic

illustration of the method and a schematic plot of data obtained are

shown in Figure 2-1.

Figure 2-1: Schematic illustration of the Wilhelmy plate method. In the diagram, the detected force F is plotted as a function of immersion depth h when the plate is immersed and withdrawn from the probe liquid. The advancing, receding and final forces are shown by FA, FR, and Ff, respectively.

Page 38: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

18 | EXPERIMENTAL

It should be noted that for a non-hygroscopic material the final force

(Ff) is zero in the plot of Figure 2-1.

Eq. 2-1 is only valid for cases in which no liquid sorption will occur

during the wetting measurements. However, for the porous and

hygroscopic materials such as wood, a time-dependent force term, Fw(t),

can be added to Eq. 2-1 due to the wicking and liquid sorption during the

measurement [48]. As a result, Eq. 2-1 is modified as follows:

𝐹(ℎ, 𝑡) = 𝑃𝛾𝑐𝑜𝑠𝜃 + 𝐹𝑤(𝑡) − 𝜌𝐴ℎ𝑔 (2-2)

2.1.1. Contact angle determination

By doing a single cycle Wilhelmy measurement, the advancing θA and

receding θR contact angles can be determined by linear regression of the

immersion and withdrawal curves to zero depth (h=0), respectively. Eq.

2-2 can be simplified to:

cos𝜃𝐴 =𝐹𝐴

𝑃𝛾 (2-3)

cos𝜃𝑅 =𝐹𝑅−𝐹𝑓

𝑃𝛾 (2-4)

where 𝐹𝐴, 𝐹𝑅 and 𝐹𝑓 (see Figure 2-1) are the advancing, receding and the

final force (which equals the term 𝐹𝑤(𝑡) in Eq. 2-2 at end of the

measurement cycle) respectively. As can be seen, for the calculation of the

contact angles, the surface tension of the liquid and the perimeter of the

sample should be known. In the next section it will explain how to

determine these quantities.

This technique was used to measure the contact angle of unmodified

Scots pine sapwood and heartwood (Papers I and II), acetylated and

furfurylated samples (Paper III) and surface modified samples (Paper V).

2.1.2. Multicycle Wilhelmy plate method

A multicycle Wilhelmy plate method was developed to investigate other

aspects of wetting of wood samples rather than the contact angle. In this

technique, the plate is immersed and withdrawn from a liquid, while the

force is continuously measured by a microbalance. In Paper I the

principle is presented, and in Paper II its application for dynamic

Page 39: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

EXPERIMENTAL | 19

wettability studies of wood surfaces is discussed. The multicycle

Wilhelmy plate method was applied on chemically modified wood in

Paper III. Finally, the use of this technique for measuring dynamic

contact angles was exemplified on surface modified wood samples in

Paper V.

By following a test procedure (a multicycle measurement and some

supplementary tests) several wetting properties of wood could be

determined as follows:

i. Wood veneers were dried in an oven at 104 OC for 1 h. The surface

tension of pure water was determined according to a standard

procedure using the Wilhelmy plate method.

ii. The multicycle Wilhelmy method in a swelling liquid (water) was

performed on the wood veneers, followed by a single cycle immersion

in a wetting-out and non-swelling liquid (n-octane) to determine the

perimeter of the swelled sample. Advancing and receding contact

angles were obtained from the curves for each cycle (see Fig. 2-1 and

Fig. 1 in Paper II). If a surface modified wood sample was

investigated, the change in advancing and receding CA was

monitored.

iii. The wood veneer was dried to the same weight as before the

measurement, and a multicycle Wilhelmy method in a non-swelling

liquid (n-octane) was performed. From the first cycle curve of this

measurement, the perimeter of dried veneer was calculated. Dynamic

sorption of wood samples was determined by linear regression of the

final forces to zero depth in both water and octane curves, in order to

dynamically investigate the swelling and capillary uptake of wood

samples. In addition, in Paper II, a perimeter model was suggested to

study the dynamic change of veneer perimeters due to the water

swelling, and this model was modified in Paper III.

iv. Finally, to quantify the migration of wood extractives to the air-water

interface during the wetting measurement, the surface tension of the

probe liquid (water) was measured.

2.1.3. Other wetting methods

A set of more simplified methods based on the Wilhelmy plate technique

were utilized for measuring dynamic wettability, swelling and void filling

into porous materials (Paper I). These techniques are based on either

Page 40: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

20 | EXPERIMENTAL

partial or complete immersion of the porous substrate into the probe

liquid. Table 2-1 summarizes key aspects of these Wilhelmy techniques.

Table 2-1: Description and application of different measurement techniques based on the Wilhelmy plate method.

Technique Description Properties extracted In Paper

Single cycle Single cycle immersion and withdrawal

Advancing and receding CA for one cycle, initial liquid uptake

I

Imbibition at constant depth

Immersion to 5 mm and partial withdrawal to 2 mm

Initial absorption, dynamic sorption, rate of sorption, distinguishes swelling liquids from non-swelling liquids

I

Immersion to constant depth (without withdrawal)

Immersion to 2 mm depth without any withdrawal

Different kinetic regimes of wetting (spreading, void filling and swelling, imbibition rate, dynamic sorption

I

Full immersion Horizontally immersed to a depth of 9 mm

Initial absorption, dynamic sorption

I

Multicycle Immersion and withdrawal for several cycles

Dynamic advancing and receding CAs, dynamic wetting, swelling and liquid uptake, distinguishes swelling liquids form non-swelling liquids, ratio of swelling/capillary uptake, dynamic dimensional stability, level of extractives dissolution

I, II, III & V

2.2. Sessile drop method

The sessile drop method is based on monitoring the profile of the drop

resting on a solid surface using a high resolution CCD camera [94]. The

captured images are then analyzed using designated software. An optical

goniometer (Dataphysics OCA40 Micro) was utilized for contact angle

experiments as a comparison to contact angle results obtained by the

Wilhelmy plate method. Droplets (4 µl) of water were applied at 9

different spots of wood plates (see Figure 2-2), 7-15 mm apart from each

other to avoid interaction between droplets that might affect the

wettability measurements.

Furthermore, for droplet freezing measurements (Paper IV) this CA

instrument with a Peltier cooling stage, temperature sensors and

temperature logger was utilized to determine the water CA of modified

wood surfaces at low temperatures and in freezing delay experiments.

Page 41: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

EXPERIMENTAL | 21

Figure 2-2: Schematic illustration of sample for contact angle measurement (sessile drop method). The measurements were done on 9 different spots (marked 1 to 9) 7±15 mm apart from each other on a 40×20×5 mm

3 (L×R×T) wood sample.

2.3. X-ray photoelectron spectroscopy

X-ray photoelectron spectroscopy (XPS) is a highly surface sensitive

analytical method based on the photoelectric effect with a typical analysis

depth of 2-10 nm. The XPS method provides quantitative information of

surface chemical composition. Figure 2-3 illustrates the main principle of

the XPS method. A sample is irradiated with X-rays under high vacuum

conditions, resulting in the emission of photoelectrons. The characteristic

binding energies can be determined by analyzing the kinetic energy of

these photoelectrons. Since this characteristic binding energy is affected

by its chemical environment, each element in a chemical functional group

has its characteristic binding energy and it allows assigning the surface

chemical composition.

Figure 2-3: Schematic illustration of the XPS technique. The kinetic energy of photoelectrons obtained by irradiation of the sample with X-rays is analyzed and the corresponding binding energies are determined. The surface chemical composition is assigned by comparing measured binding energies with known values for each element.

Page 42: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

22 | EXPERIMENTAL

In an XPS spectrum, the electron count (intensity) is plotted as a

function of binding energy. The survey and detail XPS spectra offer

quantitative data on the surface elemental composition and different

chemical states of an element, respectively. In this study, XPS was utilized

to determine surface chemical composition of pine sapwood/heartwood

(Paper II) and surface modified wood (Papers IV and V).

2.4. X-ray computed tomography

X-ray computed tomography (XCT) is a nondestructive technique for

generating 2D and 3D images at micrometer resolution of the inside of an

object using an X-ray source. A schematic illustration of the technique is

shown in Figure 2-4. The main elements of the XCT are an X-ray source,

a series of X-ray detectors and a rotating object. The sample is imaged

using the X-ray source and detectors when it slowly rotates along its axis.

The XCT slices (XCT images) can then be used to generate 3D images. In

the present study, the X-ray CT was used to investigate the

microstructural properties of chemically modified wood samples (Paper

VI).

Figure 2-4: Schematic illustration of the X-ray computed tomography technique. The rotating sample is imaged using the X-ray source and detectors.

2.5. Atomic force microscopy

Atomic force microscopy (AFM) [95] is a scanning probe technique which

is a powerful method for imaging the topography of different surfaces and

for measuring surface forces and friction. The sample is fixed on the top

of the scanner as it is scanned with a sharp tip attached to a cantilever. A

laser is focused on the backside of the cantilever and reflected onto a

quartered photodetector. The tip-sample interaction (or topography),

either attractive or repulsive (negative or positive), is recorded on the

Page 43: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

EXPERIMENTAL | 23

detector based on the cantilever deflection. This deflection response can

be transformed to a voltage output signal that provides information on

topography or surface forces. In this work, Tapping® mode was utilized to

study the topography of surface modified wood samples (Papers IV and

V). In Tapping® mode the tip gently taps on the sample surface when the

cantilever is oscillated at its resonant frequency. A topography image of

the surface is achieved by maintaining constant oscillation amplitude.

2.6. Materials

2.6.1. Wood samples

Table 2-2 summarizes the three types of wood samples used in this study:

unmodified, chemically modified and surface modified wood.

Table 2-2: Description of different wood samples used in this study.

Wood sample Abbreviation Sample description In Paper

Unm

odifie

d

wood

Scots pine --------- Scots pine sapwood

I, II, IV, V

--------- Scots pine heartwood II

Southern yellow pine

SYPctrl15.9 Matched to SYPacet15.9

III, VI SYPctrl22.2 Matched to SYPacet22.2

Maple Maplectrl Control for Maplefurf32

Beech Beechctrl Control for Beechfurf22 III

Chem

ically

mo

difie

d

wood

Acetylated SYP SYPacet15.9 With 15.9% acetyl content

III, VI SYPacet22.2 With 22.2% acetyl content

Furfurylated SYP SYPfurf28 With WPG of 28%

SYPfurf45 With WPG of 45%

Furfurylated maple Maplefurf32 With WPG of 32%

Furfurylated beech Beechfurf22 With WPG of 32% III

Surf

ace m

odifie

d w

ood

(Scots

pin

e)

Nanoparticle-treated using LFS method

Wood-TiO2 Surface treated with TiO2 nanoparticles

IV, V Plasma-treated

Wood-PFH With a plasma polymer layer of PFH

Wood-HMDSO

With a plasma polymer layer of HMDSO

Nanoparticle and plasma-treated

Wood-TiO2-PFH

With TiO2 nanoparticles and PFH plasma polymer

Wood-TiO2-HMDSO

With TiO2 nanoparticles and HMDSO plasma polymer

LFS: Liquid flame spray, PFH: Perfluorohexane, HMDSO: hexamethyldisiloxane.

In addition, for each analyzing technique the samples have been

prepared differently in shape and size as listed in Table 2-3.

Page 44: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

24 | EXPERIMENTAL

Table 2-3: Wood sample shape and size used for different techniques. L, R and T correspond to longitudinal, radial and tangential direction, respectively.

Technique Sample shape Sample size, mm3 (LxRxT)

Wilhelmy method veneer 30x7x1

Sessile drop method plate 40x20x5

XPS veneer 7x7x1

XCT needle 5x1x1

AFM veneer 30x7x1

The wood veneers were mainly used for wetting measurements. Figure

2-5 shows a schematic drawing of a wood veneer. To prevent end-grain

sorption during the wetting experiments, one cross-section of the veneers

were sealed by a polyurethane lacquer layer (see Figure 2-6).

Figure 2-5: Wood veneers with the dimension of 30×7×1 mm3 (LxRxT) were used for

wetting measurements using the Wilhelmy plate method.

Figure 2-6: Cross section SEM images of (a) un-sealed and (b) sealed wood veneers.

2.6.1. Wetting liquids

A swelling liquid (ultrapure (Milli-Q) water) and a non-swelling liquid (n-

octane) have been selected for the wetting measurements. They have

surface tension of 72.0 mNm-1 and 21.4 mNm-1, respectively.

Page 45: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 25

3. Results and discussion

3.1. Comparison of Wilhelmy and sessile drop methods

This chapter starts by comparing CA results obtained by the sessile drop

and Wilhelmy methods. I also discuss why the Wilhelmy method should

be preferred over the sessile drop method for determination of wetting

properties of porous and heterogeneous surfaces such as wood. Table 3-1

presents contact angle results of ten Scots pine sapwood replicates. To

have a similar condition as during the Wilhelmy measurement, the

average CA data for the first 10 s is reported as the initial contact angle

determined by the sessile drop method.

Table 3-1: Advancing contact angles (θA) for ten samples by the Wilhelmy plate technique and initial contact angle results on nine spots on three samples by the

sessile drop method. Average (�̅�) and standard deviations (s) are given for both methods. Averages of all contact angle results are 73±3° and 74±6° for the Wilhelmy plate and sessile drop method, respectively.

Wilhelmy plate technique Sessile drop method

Sample no. θA (o)

Spot no. Initial CA (

o)

1 74 Sample 1 Sample 2 Sample 3

2 69 1 72 71 76

3 73 2 74 62 75

4 74 3 78 86 83

5 75 4 67 70 74

6 69 5 75 82 81

7 74 6 73 73 79

8 78 7 74 69 79

9 74 8 67 75 71

10 72 9 64 82 70

�̅�

s

73

3

�̅�

s

72

5

74

8

76

4

In addition, when a water droplet is put on a wood surface, it forms an

elliptical shape with elongation in the direction of the wood fibers

(Figure 3-1). This further complicates evaluation of the contact angle and

drop volume.

Page 46: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

26 | RESULTS AND DISCUSSION

Figure 3-1: Water droplets on a wood sample, 20 x 40 mm2. Note the elliptical shape of the

droplets that is due to structural and probably also chemical heterogeneity of the wood surface.

Results from measurements of droplet volume and CA as a function of

time for nine spots on a wood plate are illustrated in Figure 3-2. The

results show a large scatter but conclusions can be made from the trends.

First, due to the hygroscopic characteristic of wood, the water droplet

volume and CA decrease with time and the resulting water penetration

can be followed.

Figure 3-2: Water wettability on nine spots on a pine wood sample determined by the sessile drop method; a) decreasing contact angle versus time, and b) remaining droplet volume. The elliptical shape of the water droplets on wood samples is the probable reason that some values are higher than 100% droplet volume. Two kinetic regimes can be observed in contact angle curves.

Two kinetic regimes are observed for the water contact angle curves: 1)

a rapid initial decrease, suggested to be due to spreading, and 2) a slower

decrease due to capillarity. From the preceding discussion it is clear that

water spreads on the surface and penetrates into the wood during the

measurement. First, this brings us to the question which CA value that

should be reported, the initial value or the average value during the first

Page 47: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 27

10 s of measurements. Second, the heterogeneous nature of the wood

surface causes very different wettability behavior on the nine spots

probed on the same sample. For example, in the contact angle curve, one

spot needs about 200 s to reach the contact angle of 20o, whereas for

another spot, it takes 550 s to reach the same contact angle value. A

similar spread in the data is observed in the droplet volume curves.

The results show that liquid spreading and penetration during CA

measurements using the sessile drop method makes interpretation of

wetting data difficult. Based on these findings and according to other

studies [39,48], it can be concluded that since wood is a hygroscopic

material and has a heterogeneous surface, the Wilhelmy plate method is a

more reliable for wettability determination than the sessile drop method. This reverts to the main motivation of this study for developing novel

techniques for studying wetting properties of wood sample based on the

Wilhelmy plate method.

3.2. Developed techniques for wettability study of wood

3.2.1. Immersion techniques

Based on the Wilhelmy method, some new approaches have been

developed for dynamically study of wettability, liquid sorption and

swelling of wood veneers. These techniques are based on partial or

complete immersion of the wood samples into the liquid while the

tensiometer instrument dynamically records the detected force with

different starting times depending on the method:

- Immersion to constant depth without any withdrawal, the veneer

rapidly immersed to a depth of 2 mm,

- Imbibition at constant depth, the veneer rapidly withdrawn to 2-

mm depth after a fast immersion to 5 mm depth,

- Full immersion, the veneer horizontally immersed to a depth of 9

mm.

Figure 3-3 illustrates typical curves obtained with the two first

methods using Scots pine sapwood veneers. The last technique has been

explained in more detail elsewhere [96]. As can be seen in Figure 3-3, a

wood sample behaves differently when immersed in a swelling liquid

(water) compared to when immersed in a non-swelling liquid (octane). In

Page 48: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

28 | RESULTS AND DISCUSSION

octane, due to the initial wicking and void filling, there is a very fast

kinetic regime in the beginning of the measurement (first 30 s). The

liquid uptake was then stopped or very slowly increased (see octane

curves in Figure 3-3). Since water can swell the wood cell wall as well as

filling the voids, the water wetting curves starts with a very fast kinetic

regime associated with spreading and CA reduction to 0o, followed by a

slower regime due to the void filling and ends with the slowest one

associated with water uptake by swelling and diffusion (Figure 3-3a).

Figure 3-3: Wettability for Scots pine sapwood in water and octane based on two different techniques; a) immersion to constant depth (2 mm) without any withdrawal, b) imbibition at constant depth by immersing the samples to 5 mm depth and then withdrawing them immediately to 2 mm depth. The different kinetic regimes are marked with lines and numbers. Three replicates for each wetting liquid, normalized by the sample perimeter. The dotted lines indicate the standard deviation of the results.

In order to keep the CA constant during the wetting experiment and to

have simpler wetting curves, another technique (imbibition at constant

depth) was developed, in which the wood veneer was first immersed to a

deeper depth (e.g. 5 mm) and it was then withdrawn to 2 mm depth. Due

to this initial immersion the CA between water and the wood surface is

zero, consequently the wetting measurement is not affected by any

changes in CA (Figure 3-3b). The results indicate that 60 min is needed to

observe all kinetic regimes and shorter tests used in other studies [68]

were not sufficient to evaluate all aspects of the wetting properties of

wood.

Page 49: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 29

By interpretation of the imbibition curves of these two techniques,

different wetting properties of porous material can be investigated. First,

in general, the dynamic wetting properties of the material can be studied.

Second, initial liquid uptake can be obtained from the first kinetic regime

(Figure 3-3a) or from the y-intercept (Figure 3-3b). Third, the rate of the

sorption and swelling can be determined form the slope of the curves in

the different kinetic regimes. In addition, by performing wetting

experiments in swelling and non-swelling liquids, we are able to evaluate

capillarity and swelling separately. Finally, from the shape of the wetting

curves, it is possible to determine whether an unknown liquid is swelling

the wood cell wall or not.

3.2.2. Multicycle Wilhelmy method

The single-cycle Wilhelmy approach was extended to a multicycle

Wilhelmy method, in which the sample is immersed and withdrawn into

the liquid for several cycles instead of one cycle. This repeated

immersion/withdrawal process allows investigating the dynamic wetting

properties of wood such as CA and its changes over time/cycle, dynamic

absorption, swelling and dimensional stability. The use of this technique

was examined on untreated (Papers I and II), chemically modified (Paper

III) and surface modified wood samples (Paper V). Figure 3-4 shows a

typical multicycle Wilhelmy curve for each wood samples in comparison

to a very smooth silica surface. It is seen that the shape of the multicycle

curves are completely different for a rough and hygroscopic material such

as wood than for a completely wetted surface such as silica. Additionally,

different types of wood samples demonstrate different wetting curves

mainly due to the changes in the surface and the bulk chemical

composition as well as changes in the microstructure after modification.

For untreated wood (Figure 3-4a), the final force (water uptake) increases

after each cycle, and the rate of this increase diminishes with the number

of cycles. This is consistent with the observation from imbibition

techniques in which the water wetting process starts with a fast initial

wicking and absorption followed by a slower swelling process.

Furthermore, due to the chemical heterogeneity and high roughness of

the wood surface, the first advancing curve is uneven. This irregular

shape of the curve is less pronounced for the receding curves since the

wood surface is already wetted during the first immersion cycle.

Page 50: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

30 | RESULTS AND DISCUSSION

Figure 3-4: Typical multicycle Wilhelmy curves for: (a) Scots pine sapwood, (b) bare silica, (c) furfurylated SYP with WPG of 48% and (d) surface modified Scots pine with TiO2 nanoparticles and HMDSO plasma polymer layer. The specimen were immersed and withdrawn from the water for 20 cycles while the instrument detected the force normalized with the sample perimeter (F/P) as a function of immersion depth (h). The final force (liquid uptake) of each cycle was calculated by extrapolation of the final force values to zero depth.

The same observation can be made for the chemically modified wood

(Figure 3-4c). However, the liquid sorption increases less with time

compared to the untreated sample and the final absorption is significantly

lower. This is due to the swollen cell walls and impregnation of wood

voids with chemical reagents during the modification process. In the case

of the smooth surface (Figure 3-4b), the force curves are straight due the

homogeneity and high smoothness. In addition, after 20 cycles the final

force is still almost zero indicating that the sample did not absorb any

water during the wetting experiment. As seen in Figure 3-4d, both

immersion and withdrawal curves shifted to lower values for the surface

modified wood samples, indicating higher advancing and receding CAs.

Page 51: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 31

Moreover, the liquid sorption (see final force values) is significantly

reduced as a result of forming a hydrophobic layer on the wood surface.

3.3. Applications of multicycle Wilhelmy plate method

3.3.1. Dynamic contact angle

The advancing and receding CAs can be determined by extrapolating the

immersion and withdrawal curves, respectively, to zero depth (see Figure

3-4). For each cycle, the difference between immersion and withdrawal

curves demonstrates the CA hysteresis, CAH. For untreated and

chemically modified wood samples, the receding CA of the first cycle is

zero as a result of water wetting the wood surface during the first

immersion cycle. Thus, CAH for these samples is equal to the value of the

first advancing CA. I note that for hygroscopic wood samples, only the

initial (the first cycle) advancing CA can be determined. However, for

surface modified wood (Figure 3-4d), since the wood surface has been

covered with a hydrophobic layer, non-zero values of the advancing and

receding CAs were obtained for subsequent cycles. In this case, the forced

wetting durability of the modified surfaces can be investigated by

studying the changes in CA after performing repeated wetting cycles. This

is another application of the multicycle Wilhelmy plate method for

hydrophobized wood surfaces to examine dynamic CA that provides more

information about wetting stability than a single CA measurement. In

general, by modifying the Eq. 2-3 and 2-4, the advancing and receding

CAs for each cycle (n) can be calculated as:

cos 𝜃𝐴,𝑛 =𝐹𝐴,𝑛−𝐹𝑓,𝑛−1

𝑃𝛾 (3-1)

cos 𝜃𝑅,𝑛 =𝐹𝑅,𝑛−𝐹𝑓,𝑛

𝑃𝛾 (3-2)

where 𝜃𝐴,𝑛 and 𝜃𝑅,𝑛 are the advancing and receding CA at cycle n,

respectively, 𝐹𝑓,𝑛−1 and 𝐹𝑓,𝑛 the final forces after cycle n-1 and n,

respectively, due to the sorption.

In the next part, the application of this technique for CA

determination will be discussed in details for two examples, chemically

modified wood (initial CA) and surface modified wood (dynamic CA).

Page 52: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

32 | RESULTS AND DISCUSSION

Contact angle of chemically modified wood

Figure 3-5 reveals the initial 𝜃𝐴 results for a series of chemically modified

wood samples, for freshly cut and thermally treated veneers. As

explained, only the advancing CA of the first cycle can be obtained for

these samples since wood surface would be completely wet after

performing an immersion cycle. As a result of the migration of the

hydrophobic extractives to the wood surface during thermal treatment

[48,96], thermally treated veneers generally showed higher CA compared

to freshly cut ones for both acetylated and furfurylated samples. This

difference is smaller for acetylated samples than for unmodified ones.

This might be related to that acetylated wood holds less extractives or

that the extractive migration is lower on the acetylated samples since

their surfaces are already hydrophobic.

Figure 3-5: Advancing water CA of acetylated, furfurylated wood and their corresponding unmodified samples with two different types of sample preparation; thermally treated at 104

OC for 1 h and freshly cut veneers. The results are based on four replicate

measurements.

Additionally, for fresh veneers, acetylated samples showed higher CA

than unmodified samples which is consistent with other studies [41,76].

Nevertheless, for thermally treated samples, no differences were observed

Page 53: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 33

between acetylated and unmodified samples. The sample containing

more furfuryl alcohol (SYPfurf45) reveals a more hydrophobic surface

compared to the sample at a low level of furfuryl alcohol (SYPfurf28).

However, in general, no significant change in CA has been observed for

furfurylated samples. Consequently, no certain trend has been observed

in other studies on CA measurements of the furfurylated samples [41,97].

Contact angle of surface modified wood

The use of the multicycle Wilhelmy method for investigating the dynamic

CA is exemplified by studies of contact angle changes on surface

modified wood samples after 20 cycles. The surface modification has

been done by deposition of metal oxide nanoparticles using the liquid

flame spray, LFS, method followed by plasma polymerisation. Figure 3-6

presents the results of 𝜽𝑨 and 𝜽𝑹 of these samples for the first cycle and

after 20 cycles. As a result of the higher roughness, the wood surface has

a higher CA than the smooth silica surface carrying a similar

hydrophobic layer. Furthermore, among wood surfaces, micro-scale

rough hydrophobic samples (wood-PFH and wood-HMDSO) show a very

high CAH compared to the multi-scale rough surfaces (wood with

nanoparticles and plasma polymer layer). Overall, creating a

hydrophobic layer on the highly rough wood surface using LFS and

plasma polymerisation leads to a superhydrophobic surface with high

𝜽𝑨 and 𝜽𝑹 and comparatively low CAH due to the multi-scale roughness

and the non-polar molecules on the surface. Combining plasma

polymerisation with LFS enhances roughness in the micro- and nano

scale, resulting in higher CA than those obtained with only plasma

polymerisation [88,89,98].

The forced wetting durability of the surfaces is examined by studying

the CA changes during multicycle Wilhelmy experiment. As can be seen

in Figure 3-6, there is no change in 𝜽𝑨 but a low reduction in 𝜽𝑹 for the

micro-nano scale rough surfaces (e.g. wood-TiO2-HMDSO) showing that

they withstand the forced wetting better than a wood modified with only

the plasma polymer.

Page 54: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

34 | RESULTS AND DISCUSSION

Figure 3-6: Advancing and receding water CA for the first cycle and cycle 20 for different surface modified samples: (a) Plasma polymer of HMDSO and (b) plasma polymer of PFH.

3.3.2. Dynamic swelling and liquid sorption

The multicycle Wilhelmy plate technique can also be utilized to

investigate the dynamic wetting, swelling and liquid uptake of the

materials. Complete withdrawal of the plate from the liquid allows

determination of the final force and consequently liquid uptake for each

cycle (see Figures 2-1 and 3-4) [96,99]. By performing several cycles it

was possible to dynamically examine the wetting, sorption and swelling

properties of the wood samples. This can be expressed as the percentage

of change in liquid mass uptake at cycle n, ln, according to:

𝑙𝑛(%) =𝐹𝑓,𝑛

𝑚𝑖𝑔× 100 (3-3)

where 𝐹𝑓,𝑛 is the final force after cycle n and mi the initial mass of the oven

dried wood veneer. Note that this is the same term as the “liquid mass” in

Paper II (Eq. 6). By plotting ln as a function of cycle/time, the dynamic

sorption could be studied. In addition, by performing multicycle

Wilhelmy measurements in non-swelling and swelling liquids, the

capillarity and swelling contributions of the liquid uptake could be

investigated. Octane as a non-swelling liquid fills only the voids of the

wood, whereas water can also swell the cell wall. This difference can be

utilized to distinguish between the capillary uptake and the swelling of

the liquid sorption [96]. A wetting term, ∆𝑙20 was defined to calculate the

swelling contribution [99]:

Page 55: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 35

∆𝑙20(%) = 𝑙𝑤,20 − 𝑙𝑜𝑐𝑡,10𝜌𝑤

𝜌𝑜𝑐𝑡 (3-4)

where 𝑙𝑤,20 is the final water uptake at cycle 20, loct,10 the final octane

uptake at cycle 10, ρw and 𝜌𝑜𝑐𝑡 the density of water and octane,

respectively.

Unmodified wood

Scots pine sapwood and heartwood were selected as unmodified wood

samples to study dynamic swelling properties. Water and octane uptake

as a function of cycle/time is presented in Figure 3-7. The shape of the

water curves are similar to those obtained with the imbibition techniques

(Figure 3-3) and starts with a very fast water wetting on the surface,

followed by the void filling and ends with a time-dependent liquid

diffusion in the cell wall resulting in swelling. From the results, it is clear

that pine heartwood is significantly more hydrophobic than pine

sapwood. This is mainly demonstrated by the lower sorption rate and

water uptake level after 20 cycles. Additionally, according to the CA

measurements, the sapwood has lower CA than heartwood (65±7° vs.

83±3°), causing faster wetting and void filling for pine sapwood. The

results are consistent with other research findings [100,101].

Figure 3-7: Liquid mass uptake for unmodified wood samples (pine sapwood and heartwood) in water and octane.

Page 56: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

36 | RESULTS AND DISCUSSION

For both water and octane wetting, the initial liquid uptake dominated

by void filling, occurs approximately with the same rate. However, the

additional water uptake occurring at longer times is dominated by the

slower process of swelling into the cell walls. For pine heartwood, this cell

wall swelling is significantly slower than sapwood and it might be

concluded that a 20-cycle test was not sufficient to distinguish all wetting

kinetic regimes. This is mainly illustrated by that after 20-cycle Wilhelmy

in water and 10-cycle Wilhelmy test in octane, the liquid uptake are in a

same level for both liquids (see water and octane curves in Figure 3-7).

These different wetting properties will be related to surface chemical

composition of the pine sapwood/heartwood later in section 3.4.1.

Chemically and surface modified wood

Figure 3-8 gives dynamic water sorption for the chemically and surface

modified wood samples. As can be seen in Figure 3.8a, the water sorption

is strongly affected by chemical modification. A significant reduction in

water uptake is shown as a result of acetylation and furfurylation. This

enhancement in water resistance can be related to the cell walls swollen

by the modifying chemicals, resulting in less accessible hydroxyl groups.

Additionally, in the furfurylation process, furan polymer fills wood voids

and decreases the porosity and as a result reduces the capillary uptake.

Figure 3-8: Dynamic water uptake for (a) chemically modified wood and (b) surface modified wood. Chemically modified samples were acetylated and furfurylated (at two levels of modification) southern yellow pine (SYP). Surface modification was done on Scots pine sapwood and wetting properties of the pine sapwood (bare wood) is also presented in figure b.

Page 57: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 37

For the surface modified wood samples (Figure 3-8b), the water

uptake is significantly reduced for the LFS and plasma polymer treated

wood compared to the control sample. The same trend has also been

observed for the PFH-treated samples. Comparing chemically and surface

modified samples (Figure 3-8a,b) clearly demonstrates that

hydrophobised wood samples are significantly more water-repellent than

acetylated and furfurylated samples, at least during the time scale

investigated (about 1 h).

As stated above, combining the water and octane uptake results makes

it possible to distinguish between capillary liquid uptake and swelling.

The final water (20 cycles) and octane (10 cycles) sorption results are

given in Figure 3-9a for chemically modified wood samples.

Figure 3-9: (a) Final water and octane uptake based on 20-cycle and 10-cycle measurements, respectively, and (b) swelling part of the liquid sorption (Eq. 3-4) for acetylated and furfurylated samples. The control sample for SYPfurf28 and SYPfurf28 is SYPctrl15.9.

For all acetylated and furfurylated samples, a major reduction in water

uptake can be observed compared to the control samples. However, for

acetylated samples, the octane uptake remains close to unchanged

(SYPacet22.2) or slightly reduced (SYPacet15.9) after the acetylation. This

demonstrates that although the acetylation causes swelling on the wood

cell walls, it has a weak effect on the wood porosity, and hence a minor

effect on the capillary uptake (void filling). Conversely, the furfurylation

significantly decreases the octane uptake which might be attributed to the

reduction in porosity.

Page 58: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

38 | RESULTS AND DISCUSSION

Although the furfurylation process starts by formation of a linear

polymer, crosslinking reactions become more favorable as the

polymerisation progresses and this eventually results in a highly

branched and cross-linked polymer [85]. It is suggested that these cross-

linked poly (furfuryl alcohol) chains fill the voids and partly block the

liquid flow paths, whereby reducing both the rate and extend of liquid

uptake.

Acetylation mainly reduces swelling by water, while furfurylation

reduces both capillary liquid uptake and swelling. This difference in

capillary uptake was the main motivation to perform a microstructural

study of these samples in order to relate the wetting properties to

morphological characteristics (see section 3.4.2).

The extent of water uptake due to the swelling expressed by ∆𝑙20 (Eq.

3-4), is also presented in Figure 3-9b. As can be seen, the swelling level is

greatly affected by the chemical modification. However, for these levels of

modification and measurement time scale, the furfurylated samples

demonstrates lower water uptake due to swelling compared to the

acetylated ones. The furfurylated southern yellow pine with WPG of 45%

(SYPfurf45) shows lower water and octane uptakes and swelling level than

other samples.

3.3.3. Dimensional stability

Dimensional stability of the wood veneers can be studied by monitoring

the sample dimension during the wettability experiment. However, only

the initial and final (after wetting measurement) perimeters can be

directly determined and no dynamic response can be investigated. Thus, a

model for calculating the sample perimeter at intermediate cycles was

suggested based on a linear combination of the measured force and final

change in sample perimeter [96,99]:

𝑃𝑛 = 𝑃𝑛−1 +(𝐹𝑓,𝑛−𝐹𝑓,𝑛−1)∆𝑃

∆𝐹𝑓

(3-5)

where 𝑃𝑛 and 𝑃𝑛−1 are the veneer perimeter after cycle n and (n-1),

respectively, 𝐹𝑓,𝑛 and 𝐹𝑓,𝑛−1 the final force for cycle n and (n-1),

respectively, ∆𝑃 is the total perimeter change during the measurement. It

is worth noting that ∆𝐹𝑓 was first defined as total final force during the

Page 59: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 39

measurement [96]. However, it was realized that it is not possible to

completely distinguish between "capillary uptake" and "swelling", water

especially in the initial cycles. Thus, the contribution of the wicking and

capillary uptake should also be considered. The model was modified by

using the approximation that void filling completely dominates the

sorption process during the first cycle while swelling is the dominating

sorption process during the following cycles [99]. Thus, ∆𝐹𝑓 was defined

as the final force difference between last cycle and first cycle (∆𝐹𝑓=𝐹𝑓,20 −

𝐹𝑓,1). The results demonstrate that this overestimation in the initial

cycles could be almost solved by separating the two dominating effects of

void filling and the swelling [99].

The model was validated by studying the apparent perimeter changes

of pine sapwood and heartwood veneers as a function of cycle number

(see Figure 3-10) [96]. The results are based on octane immersion

experiments and on the calculations using the modified model (Eq. 3-5).

As shown, compared to the experimental results, the model provides a

reasonable approximation for determining the perimeter changes during

repeated cycling. It can be concluded that this model demonstrates a

relation between the liquid uptake and the perimeter changes in swellable

materials.

Figure 3-10: A comparison of the model and experimental (octane immersion) of apparent perimeter change for pine sapwood and heartwood veneers during multi-cycle Wilhelmy plate experiments. All model data (filled symbols) based on ten measurements, were calculated using the modified perimeter model (Eq. 3-5) with the exception of the final perimeter (P20) which was determined by a single immersion in octane after 20 cycles in water. Empty symbols refer to the octane immersion data (5 replicates) of intermediate cycles (C5, C10, C15). The standard deviation of the octane immersion and the model are illustrated by error bars and lines, respectively.

Page 60: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

40 | RESULTS AND DISCUSSION

In addition, the pine sapwood samples showed an increased perimeter

change compared to the heartwood veneers (8% vs. 4%) after a 20-cycle

measurement, which is in line with the water uptake where pine sapwood

absorbed about twice as much water as heartwood.

This technique was utilized for a study of dimensional stability of

chemically modified wood samples. Figure 3-11 illustrates the perimeter

changes of acetylated and furfurylated (at two levels of modification) SYP

during a 20-cycle Wilhelmy experiment. The perimeters were calculated

from the model with the exception of the initial and final perimeters

which were obtained from octane immersion. A significant difference can

be observed between dimensional stability of unmodified and chemically

modified wood. Based on these results, SYPacet15.9 and SYPfurf45

demonstrate a perimeter change reduction of 70% and 93%, respectively,

compared to the unmodified sample. This suggests that wood cell wall is

less prone to swelling after the chemical modification as a result of the

formation of chemical bonds between the modifying chemicals and the

cell wall. Additionally, furfurylated sample with higher WPG (SYPfurf45)

are clearly more dimensionally stable than at low WPG (SYPfurf28). The

values of the reduction in perimeter change obtained from this method

were found to be consistent with the anti-shrinkage efficiency (ASE)

values commonly used in investigations [85,102,103].

Figure 3-11: Apparent perimeter change of acetylated and furfurylated SYP samples during multicycle Wilhelmy experiments. All data were calculated using the perimeter model (Eq. 3-5) with the exception of the final perimeter (P20) that was determined by a single immersion in octane after 20 cycles in water.

Page 61: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 41

3.3.4. Dissolution of extractives

Recent studies have demonstrated that the surface tension of the probe

liquid (especially water) is changed during wood wettability

measurements [71,72] ,and this has been suggested to be due to migration

of extractives from the wood surface to the air-water interface. It is worth

to mention that, a surface tension reduction can be observed even after a

single-cycle Wilhelmy experiment [72] and should carefully be considered

for any further calculation and measurements. However, in this study,

since the instrument measures the real detected force and this force was

then utilized for liquid uptake calculations, the surface tension change

will not influence the wettability results.

Figure 3-12: Surface tension reduction of the water during and after the multicycle Wilhelmy experiment for (a) pine sapwood sample as a function of cycle number and (b) different chemically modified wood veneers. The measured surface tension

reduction, ∆𝛾𝑚𝑒𝑎𝑠, is defined as the surface tension deference of pure water and water after performing the multicycle Wilhelmy experiment (𝛾 − 𝛾𝑓).

The surface tension reduction of the water was examined during

(Figure 3-12a) and after (Figure 3-12b) the multicycle Wilhelmy plate

measurement on pine sapwood and chemically modified wood samples.

Use of the pine sapwood samples results in a surface tension reduction

which increases with increasing number of cycles, reaching a plateau level

after 10-15 cycles. This is suggested to be related to the migration of

surface active extractives to the air-water interface. Conversely, for the

pine heartwood samples (not shown in the figure) the surface tension of

Page 62: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

42 | RESULTS AND DISCUSSION

the water was almost constant during a 20-cycle measurement,

demonstrating that pine heartwood contains a type of extractives which

are less prone to migrate. The results indicate the complication of

analyzing wettability on complex materials such as wood. It should be

noted that these changes in surface tension of the probe liquid have been

overlooked in many studies and but can provide valuable information

about the liquid-solid properties.

For the chemically modified samples and the corresponding unmodified

control samples, the final surface tension reduction after 20 cycles are

presented in Figure 3-12b. The surface tension reduction is much higher

for SYP and maple than for pine sapwood and beech. In addition, a higher

surface tension reduction can generally be observed for unmodified

samples compared to the modified ones. This can be related to that the

majority of extractives were removed during the modification process.

Moreover, after the chemical modification the wood cell walls are likely

thicker and more hydrophobic which might lead to extractives being less

mobile.

3.4. Factors affecting wood wetting properties

3.4.1. Surface chemical composition

Since x-ray photoelectron spectroscopy XPS is a highly surface

sensitive technique, it is known as a powerful method to determine

chemical composition of the wood surfaces and to examine surface

chemical changes [40,104-107]. Knowing the surface chemistry is crucial

for understanding the liquid-surface interaction. Thus, wetting properties

have often been related to the surface chemical composition obtained by

XPS [40,49,106,108,109]. It is very informative to analyze wood surfaces

in terms of the O/C atomic ratio and the relative amount of different

carbon types (C1-C4) and to determine the level of cellulose, lignin and

extractives [107,110]. Increasing the O/C atomic ratio and decreasing the

C1 component (C–C and C–H bonds) indicate an increase of the

cellulose/hemicellulose level and thus a reduction of lignin and extractive

levels on wood surfaces [110].

In the present work, wettability results (mainly CA) were related to the

surface chemical composition for Scots pine sapwood and heartwood as

unmodified wood surfaces and for surface modified wood samples.

Page 63: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 43

Scots pine sapwood and heartwood

According to the wettability results (see section 3.3.2), pine sapwood

showed lower contact angle, higher liquid uptake and higher swelling

than pine heartwood. Table 3-2 provides the surface chemical

composition in atomic % of pine sapwood and heartwood (measured) and

of cellulose, hemicellulose, lignin and a typical extractive (literature).

Cellulose has a polysaccharide structure with no C1 carbon. Conversely,

extractives are generally devoid of C2 carbon and mainly contain aliphatic

carbons (C1-carbon). Finally, lignin contains both C1 and C2 carbons at

approximately the same level. Thus C1 carbon is the most abundant in

extractives and the least abundant in cellulose and hemicellulose.

Consequently, a higher C1 level (and a lower O/C ratio) demonstrates an

enhanced level of lignin and extractives on the surface. These differences

could be utilized to study the level of wood components (cellulose, lignin

and extractives) on pine sapwood and heartwood by means of XPS

analysis.

Table 3-2: Surface chemical composition of Scots pine sapwood and heartwood based on atomic percent, different carbon types (C1- C4) and O/C ratio. Similar data are presented for wood components based on literature data. Chemical composition of Scots pine is 40% cellulose, 28.5% hemicellulose, 27.7 lignin and 3.5 extractives [13].

C1: C-C, C=C, C-H at 285.0 eV C2: C-O, C-O-C at 286.7 eV; C3: C=O, O-C-O at 288.2 eV C4: O-C=O, C(=O)OH at 289.4 eV

As seen in Table 3-2, the pine sapwood surface is more oxidized (higher

O/C ratio) than pine heartwood. This demonstrates that extractives are

more abundant on the surface of pine heartwood than pine sapwood,

which has been suggested before [6,112]. Moreover, it may be concluded

that pine heartwood contains more lignin than pine sapwood, since C2

and C3-carbons are more abundant on sapwood (see Table 3-2).

Sample O1s C1s C1 C2 C3 C4 O/C

Cellulose[111]

45.4 54.6 --- 83 17 --- 0.83

Hemicellulose (Arabinoglucuronoxylan)

[111]

44.8 55.2 --- 78 19 3 0.81

Lignin[111]

24.8 75.2 49 49 2 --- 0.33

Extractives (oleic acid)[111]

9.9 90.1 94 --- --- 6 0.11

Sapwood1 26 73 52±2 35±1 8±2 5±2 0.36

Sapwood2 28 71 43±2 43±2 10±3 4±2 0.39

Heartwood1 18 81 69±2 24±1 5±2 3±2 0.23

Heartwood2 17 83 68±3 23±1 5±3 4±3 0.21

Page 64: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

44 | RESULTS AND DISCUSSION

Now, when the surface chemical composition of the pine sapwood and

heartwood has been determined, it can be related to their wetting

characteristics. For this purpose, it is informative to plot the contact

angles as a function of the O/C ratio. The experimental results of this

work (pine sapwood and heartwood) were combined with the literature

results (cellulose and lignin) and illustrated in Figure 3-13. It clearly

demonstrates that the O/C ratio is higher for surfaces with lower contact

angle. According to literature, the water CA of isolated cellulose [113-115]

and lignin [116,117] are 20-30° and 55-77°, respectively. In addition, most

extractives are hydrophobic. Therefore, pine sapwood as a wood species

containing more cellulose and less lignin and extractives, is expected to

be more hydrophilic with lower water CA than pine heartwood, in line

with wetting results.

Figure 3-13: Contact angle as a function of O/C atomic ratio for Scots pine samples (sapwood and heartwood) and for lignin and cellulose. CA of cellulose [113-115] and lignin [116,117] were calculated as an average based on what have been reported. O/C ratio data for cellulose and lignin were also based on literature data [111].

As stated in section 3.3.4, although extractives are more abundant in

pine heartwood than in pine sapwood (according to XPS), they are less

prone to migrate to the surface and change the water surface tension

(from extractives dissolution). This shows that not only the level of

extractives but also the type of extractives are different in pine sapwood

and heartwood. In addition, during the conditioning in vacuum for XPS

analysis and the conditioning for wettability measurements, some

extractive migration occurs, especially for pine sapwood since it contains

Page 65: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 45

extractives which are more mobile. In other words, if it was possible to do

XPS analysis without any conditioning in vacuum, and the wettability

measurements were done without any conditioning, the pine sapwood

and heartwood would be even more clearly distinguished in terms of

hydrophobicity and surface chemistry.

As described there are three different kinetic regimes in wood wetting

curves (see Figure 3-3). The first and the fastest one which is associated

with wetting and spreading of the liquid on the wood surface, is mainly

affected by surface chemical composition and surface topography. The

other slower kinetic regimes (wicking and swelling) specify the other

wetting properties such as swelling and dimensional stability of wood. It

is worth noting that these wetting properties would mainly be affected by

the structure and chemical composition of wood bulk. This will be

discussed in the next section for pine sapwood and heartwood.

The surface chemical composition of the pine sapwood/heartwood

interface was also investigated by detailed XPS spot analysis (most signal

from area with a diameter of 55 µm) at different spots across the interface

between pine sapwood and heartwood. Figure 3-14 gives the results (not

included in the appended papers to this thesis) of the O/C atomic ratio,

relative amount of C1 and C2 as a function of distance from the interface

(as defined by ocular inspection) of pine sapwood/heartwood. In all

figures, distinctive changes in these parameters can be observed in the

interface area, showing that the surface chemical composition is

significantly different between sapwood and heartwood areas.

However, there is no strict border defining the sapwood/heartwood

interface but instead an interphase area with the approximate width of 2

mm. There are some exceptional spots in heartwood area in terms of

chemical composition. One possible explanation for these exceptions

could be that the XPS analysis has been done in earlywood (EW) or

latewood (LW) areas as they are chemically different. As can also be seen

in Figure 3-14, the chemical composition of pure areas of sapwood or

heartwood is consistent with the chemical composition of all wood

samples (Sapwood2, Heartwood1,2) presented in Table 3-2. For all these

analysis spots, pine sapwood show higher O/C and C2 and lower C1

compared to the pine heartwood. Considering the quantitative level of C1

and C2, the sample sapwood2 may be a better representative of the

sapwood area in terms of chemical composition than sapwood1.

Page 66: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

46 | RESULTS AND DISCUSSION

Figure 3-14: The O/C atomic ratio, relative amount of C1 (aliphatic carbon) and C2 as a function of distance from the interface of Scots pine sapwood/heartwood. The results are based on the XPS spot analysis. The vertical dotted line indicates the interface defined by eye prior to XPS analyses.

Surface modified wood samples

The wood surfaces modified by LFS and plasma polymerisation were also

chemically analyzed using XPS. Table 3-3 presents the results of total

amount of C and of different carbon types and as well as Si/C and F/C

ratios for HMDSO-treated and PFH-treated surfaces, respectively.

Comparing bare wood and surface modified wood reveals that the

surfaces were significantly changed in terms of chemical composition due

to the surface treatment. This explains the changes in hydrophobicity of

the surface modified samples (see section 3.3.1). The Si/C atomic ratio is

similar for HMDSO-treated surfaces of silica and wood. Furthermore,

additional information can be obtained by interpreting the chemical

shifts of the carbon (C1s). Theoretically there is only unoxidized C1-

carbon in branched and cross-linked siloxane polymers. Nevertheless, a

low amount of oxidized carbons (C2-C4) was observed for both Si-

Page 67: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 47

HMDSO and wood-TiO2-HMDSO surfaces, which could be related to

reactions with oxygen or water vapor during the plasma polymerisation

process.

Table 3-3: The total amount of carbon (in atomic%) divided into different types of carbon, as well as Si/C and F/C ratios for bare wood and surface modified silica and wood samples.

C1 (285.0 eV): C-C, C=C, C-H C2 (286.3-5 eV): C-O, C-O-C C3 (287.3-288.0 eV): O-C-O, C=O, C*-CF C4 (288.6-7 eV): O-C=O, C(=O)OH C5 (289.9 eV): CF C6 (292.0 eV): CF2 C7 (294.0 eV): CF3

For the PFH-treated samples, the wood substrate showed a lower F/C

atomic ratio than the silica substrate. In addition, the level of (C6+C7)

which shows carbon bound to fluorine, are more abundant on silica than

on wood. The chemical composition of silica-PFH is closer to what has

been found in another study [118]. This could be explained by TiO2 acting

as a catalyst for PFH degradation, consequently forming more C1-

carbons, or, migration of extractives to the wood surface either during the

sample preparation [96] or during XPS conditioning [119]. Since

extractives are mainly hydrocarbons, the level of C1-carbons would

become larger.

3.4.2. Microstructure

Unmodified pine sapwood and heartwood

Pine sapwood and heartwood have surfaces with different hydrophobic

properties, mainly related to their surface chemical compositions.

Sapwood and heartwood samples demonstrate different sorption,

swelling and dimensional stability properties, most likely related to the

detailed morphological differences. Figure 3-15 illustrates x-ray

computerized microtomography (XCT) scan images of sapwood and

heartwood as well as an image of the sapwood/heartwood interfacial

zone. In both pure heartwood area and the area in the interface closer to

the heartwood, the tracheids were filled more with wood resins especially

Sample C tot Si/C F/C C1 C2+C3+C4 C6+C7

Bare wood (pine sapwood)

72±1 0 0 34±4 38±3 0

Si-HMDSO 66±1 0.26±0.02 64±1 2±1 0

Wood-TiO2-HMDSO 59±1 0.29±0.01 55±1 4±0 0

Si-PFH 32±1 1.9±0.1 2±2 n/a 17±1

Wood-TiO2-PFH 44±1 1.1±0.1 12±4 n/a 13±2

Page 68: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

48 | RESULTS AND DISCUSSION

in the LW parts. This demonstrates that heartwood contains more

extractives located in the wood cell wall, consequently the rate and

amount of the swelling is lower for heartwood than sapwood. More

aspiration of bordered pits in heartwood could be another reason for the

lower rate of swelling and sorption.

Figure 3-15: X-ray computed microtomography images of transverse sections of (a) pine sapwood, (b) pine sapwood/heartwood interface and (c) pine heartwood. The scale bars correspond to 200 µm.

Chemically modified wood

The EW/LW proportion of the wood sample could affect the wetting

properties. The SYPacet22.2 and SYPctrl22.2 wood veneers have higher EW

proportion and fewer annual rings compared to the SYPacet15.9 and

SYPctrl15.9 samples. Figure 3-9 shows that SYPacet22.2 and SYPctrl22.2

absorbed more liquid (especially water) than those samples with less EW

proportion. This is in line with a study [76] where it was demonstrated

that acetylated SYP latewood has significantly lower liquid uptake than

EW samples. On the other hand, it was demonstrated that for LW the

preferential pathway for liquid uptake is in the longitudinal and

transverse directions, while in the radial direction the growth rings

restrict liquid sorption [120]. According to results in the present study,

the total amount of liquid uptake is in overall lower in LW. For the

chemically modified samples, an interesting finding was that acetylated

and furfurylated samples differed in their sorption properties, in which

acetylation mainly decreased the water swelling, whereas furfurylation

deceased both capillary uptake and water swelling (see section 3.3.2).

Therefore, the XCT technique was used to investigate the microstructural

properties of these samples qualitatively and quantitatively.

2D microstructural images of the chemically modified EW samples are

given in Figure 3-16. No visible microstructural differences can be seen

Page 69: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 49

for acetylated and unmodified SYP earlywood (Figure 3-16a,b). The shape

and size of tracheids are similar and no significant changes in cell wall

thicknesses can be observed. No visible effect on the wood ultrastructure

has also been reported for acetylated samples studied with SEM and TEM

[121]. This could be related to the fact that during the acetylation, in

contrast to furfurylation, no polymerisation or cross-linking occurs and

actually acetylation is a single site reaction.

Fig 3-16: X-ray computed microtomography images of transverse sections of (a) unmodified SYP earlywood (SYPctrl15.9) (b) acetylated SYP (SYPctrl15.9) and (c) furfurylated SYP (SYPfurf28). The lower images (d-e) are XCT images of furfurylated SYP (SYPfurf45): (d) transverse, (e) tangential, (f) radial. All samples were from the earlywood area. The scale bars correspond to 100 µm.

Unlike the acetylation samples, a significant change in microstructural

properties was detected for furfurylated wood samples (Figure 3-16c-f).

As seen in Figure 3-16 a, the majority of the tracheids in the transverse

cross-section of unmodified SYP EW are rectangular, pentagonal and

hexagonal shape. However, after a low level of furfurylation (WPG=28%)

(Figure 3-16c) the shape of these tracheids became more circular or

Page 70: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

50 | RESULTS AND DISCUSSION

elliptical. In addition, it is clearly revealed that some tracheids were

completely filled with furan polymers and in some others the furan

polymer chains were stacked inside the tracheids. This was also observed

in tangential and radial images (not shown). Furthermore, the cell walls

of furfurylated samples are clearly thicker than for unmodified ones due

to swelling by furfuryl alcohol molecules. When a high level of

furfurylation (WPG=45%) was applied, even more tracheids were filled

with furan polymers and the cell walls seemed to be slightly thicker

compared to SYPfurf28 (Figure 3-16d). This is consistent with the finding of

Wålinder et al. [122] that more polymer-filled lumen were found in the

high-WPG furfurylated samples compared to the low-WPG ones.

Additionally, as illustrated in Figure 3-16e,f, the tracheids are completely

or partly filled due to the chemical modification. Furan polymer deposits

can be found inside LW tracheids (not shown, see Paper VI) as well,

where some of the tracheids are completely filled.

The XCT images of unmodified and furfurylated maple are presented

in Figure 3-17 as an example of a chemically modified hardwood sample.

The images clearly show that a significant microstructural changes

occurred after furfurylation. An extensive furan polymer deposit could be

distinguished within vessels, fibers and rays. The rays were mostly filled,

while some of the vessels were filled and most of the fibers remained

empty. This indicates that during the furfurylation process, the furfuryl

alcohol molecules selectively impregnate the wood void spaces.

Figure 3-17: X-ray computed microtomography images of transverse sections of (a) furfurylated maple (Maplefurf32) and (b) unmodified maple (Maplectrl). The scale bar corresponds to 100 µm.

Page 71: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 51

It is possible to visualize the 3D reconstruction using XCT which gives

the opportunity to illustrate the wood microstructure in details without

any further sample preparation. The advantage of the 3D reconstruction

will be explained by an example of a furfurylated SYP sample (see Figure

3-18). The deposition of furan polymers can clearly be observed on the

wall of the tracheids (Figure 3-18a,b). Additionally, as seen in Figure 3-

18c, visualization of the 3D reconstruction helps to understand whether a

tracheid was partially or completely filled by the furan polymers. This is a

clear advantage of 3D XCT over 2D techniques (e.g. SEM) which offers

the opportunity to quantitatively investigate the microstructure of these

chemically modified wood samples. Using softwares, analytical properties

of these samples can be determined such as total porosity and porosity of

different wood structural components.

Figure 3-18: Volume rendering of furfurylated SYP earlywood with WPG of 28%: (a) with randomly colour labeled tracheid lumens, (b) a zoomed view of the deposits on the tracheid cell wall a tracheid inside and (c) randomly colur labeled tracheid lumens of furfurylated SYP with WPG of 45%.

Total porosity and cell wall thickness results of different samples are

presented in Table 3-4. For the acetylated samples (in both EW and LW

areas), the total porosity is similar (SYPacet22.2) or slightly lower

(SYPacet15.9) than the corresponding control samples. Conversely, the

Page 72: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

52 | RESULTS AND DISCUSSION

results show that furfurylation significantly reduces the porosity of the

sample in both EW and LW areas. This can be related to the wetting

results (see section 3.3.2) where no changes in octane uptake were

observed for acetylated.

Table 3-4: Total porosity and cell wall thickness (mean±standard deviation) determined from software analysis of XCT scans of different acetylated and furfurylated samples. The control sample for SYPfurf28 and SYPfurf45 is SYPctrl15.9.

Sample

Total porosity (%) Cell wall

thickness (µm)

Earlywood Latewood Earlywood Latewood

SYPacet15.9 73.8±1.8 24.1±5.2 5.7±2.1 14.3±6.9

SYPctrl15.9 76.1±1.8 29.3±1.0 5.3±2.0 14.6±5.2

SYPacet22.2 76.6±0.6 26.4±0.8 5.5±2.2 16.1±5.4

SYPctrl22.2 75.3±0.9 27.5±4.4 5.5±2.1 13.9±5.8

SYPfurf28 66.2±3.8 16.7±3.5 8.4±4.5 19.6±7.2

SYPfurf45 61.1±2.7 18.0±1.0 9.2±5.0 19.2±8.3

Maplefurf32 38.9±1.3 n/a n/a

Maplectrl 51.3±2.7 n/a n/a

Figure 3-19: Octane uptake as a function of total porosity for different modified and unmodified wood samples.

Figure 3-19 demonstrates that the total porosity and octane uptake

(capillary uptake) is related for acetylated and furfurylated

softwood/hardwood samples. The total porosity of each specimen was

calculated based on the EW/LW proportion of that specimen (see Paper

VI). According to this relation, a wood sample of higher porosity has

higher wicking and capillary uptake. For instance, an acetylated sample

with approximately same porosity as its control sample (SYPacet22.2),

absorbed almost the same amount of octane in wetting measurements as

Page 73: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 53

the corresponding control sample, while a high level furfurylated sample

(SYPfurf45) showed a significant lower level of porosity and octane uptake

compared to the unmodified sample (SYPctrl15.9).

The porosity of the microstructural components of the wood can also

be determined using the XCT technique. As an example, the total porosity

and porosity of the vessels, fibers and rays are presented in Figure 3-20

for a furfurylated hardwood sample (Maplefurf32) and its corresponding

control sample (Maplectrl32). This detailed microstructural study

demonstrates that the furfurylation impregnation mostly occurred in rays

and partly in some vessels while the wood fibers were less affected by the

modification.

The cell wall thickness (CWT) of acetylated samples are similar to that

of unmodified samples, whereas CWT of furfurylated sample are much

thicker than the control samples in both EW and LW areas (Table 3-4).

These values are mean values of the cell wall thickness of 50-100

tracheids. More details can be obtained by considering the distribution of

the CWT given in Figure 3-21 for acetylated and furfurylated SYP in EW

and LW areas. The CWT slightly shifts to higher values for acetylated

samples (Figure 3-21a,c) compared to the control samples, while this

difference in CWT is significantly greater for furfurylated samples

especially in LW areas (Figure 3-21d). Additionally, the wood cell walls

are thicker for a high level furfurylated sample (SYPfurf45) than a low level

one (SYPfurf28) in both EW and LW areas.

Figure 3-20: Total porosity determined from XCT scans for different microstructural components of hardwood samples: Furfurylated maple and unmodified maple.

Page 74: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

54 | RESULTS AND DISCUSSION

Figure 3-21: The distribution of the cell wall thickness (CWT), for (a) acetylated SYP earlywood, (b) furfurylated SYP earlywood, (c) acetylated SYP latewood and (d) furfurylated SYP latewood.

3.4.3. Temperature effects of wetting properties

Although interactions between water and hydrophobic wood surfaces

have been studied previously, only a limited number of reports has

explored the impact of low temperatures on the interaction of water

droplets with hydrophobic surfaces. It should be noted that wetting

properties of wood at low temperatures could be critical for the

performance of wood-based materials in cold climates.

The effect of temperature on the wetting properties was investigated

by performing CA measurements using the sessile drop method on

surface modified silica and wood samples during a freeze-thaw cycle.

Figure 3-22 shows the static water CA on these modified samples during a

Page 75: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

RESULTS AND DISCUSSION | 55

freeze-thaw cycle starting at room temperature, cooling down to -7 °C and

finally increasing to room temperature again.

Figure 3-22: Static contact angles of (a) Si-PFH, (b) Si-HMDSO (c) wood-PFH, (d) wood-HMDSO, (e) wood-TiO2-PFH, and (e) wood-TiO2-HMDSO. The results are based on 3 measurements in a freeze-thaw cycle. The CA measurements were done in contact with ambient air with about 40% relative humidity at 23 °C.

The wetting properties of these surface modified wood samples are

highly affected by the temperature. For the flat surface modified silica, a

small wetting hysteresis (changes in CA after one freeze-thaw cycle) can

be observed especially in the cooling part where CA slightly decreases

with decreasing temperature. The water contact angle at room

Page 76: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

56 | RESULTS AND DISCUSSION

temperature at the beginning of the experiment, initial CA, is higher for a

rough modified wood surface than for the smooth modified silica surface.

As shown in section 3.4.1, the chemistry of HMDSO plasma polymer layer

is very similar on the wood and silica surfaces, hence the difference in

wetting properties can clearly be related to the larger roughness of the

surface modified wood samples. The larger roughness of the wood surface

may also result in a larger freeze-thaw induced wetting hysteresis

(compare the initial and final CA in the beginning and at the end of the

freeze-thaw cycles in Figure 3-22). Additionally, the multi-scale rough

wood surfaces (e.g. wood-TiO2-HMDSO) have higher initial CA and

smaller freeze-thaw induced wetting hysteresis than a micro-scale rough

surface (e.g. wood-HMDSO) in line with the wetting results measured by

the Wilhelmy plate method (see section 3.3.1). This smaller hysteresis is

related to the multi-scale roughness introduced by the LFS treatment. As

a conclusion, the LFS treatment enhances both the initial hydrophobicity

and the hydrophobicity after a freeze-thaw cycle.

As a result of water vapor condensation, the CA of plasma polymer

treated wood surfaces begins to decrease at a temperature of about 10 oC.

This decrease continues at subzero temperatures, mainly due to frost

formation on the surfaces that facilitates a partial transition from the

Cassie-Baxter state to the Wenzel state [123].

Page 77: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

CONCLUSIONS | 57

4. Conclusions

This thesis focuses on wetting phenomena of wood with emphasis on

modified wood. For this purpose novel wetting measurement techniques

were proposed to investigate dynamic wetting properties of porous and

swellable solids. The Wilhelmy plate method was chosen as the main

technique to assess wettability and methods were successfully utilized for

measuring dynamic wettability, void filling, and liquid sorption into

wood. Combining the wetting of a swelling liquid (water) and of a non-

swelling liquid (octane) makes it possible to study both swelling and

capillary uptake (void filling) properties of a wood sample. Based on

results of wetting for water three kinetic regimes were found, starting

with a very fast one mainly associated with spreading, followed by a

slower regime due to the capillary uptake and ending with the slowest one

due to the swelling. Since the Wilhelmy plate method measures an

average contact angle of a larger area than the sessile drop method, the

wetting properties measured by the Wilhelmy plate are less affected by

the chemical and structural surface heterogeneity of wood.

A multicycle Wilhelmy plate method was developed based on repeated

immersion and withdrawal in the probe liquid. The application of this

technique was examined on both unmodified and modified wood

samples. It was concluded that by performing this technique and

supplementary tests, several wetting properties can be measured

including dynamic advancing and receding contact angles, dynamic

wetting and swelling, ratio of swelling/capillary uptake, dynamic

dimensional stability and the level of extractives dissolution.

It was found that the wetting properties of wood are affected by many

factors such as surface chemical composition, surface topography,

microstructure and temperature. The results of wetting showed lower

contact angle, higher liquid uptake, higher swelling, less dimensional

stability and higher water contamination for pine sapwood than pine

heartwood. This observation was in line with results from X-ray

Page 78: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

58 | CONCLUSIONS

photoelectron spectroscopy where a higher level of extractives and lignin

was found in heartwood than in sapwood. Wetting of chemically modified

wood samples revealed that acetylation mostly impacts on swelling, while

furfurylation decreases both capillary uptake and swelling. X-ray

computed tomography was found to be a suitable technique for

visualization and quantitative determination of anatomical characteristics

(CWT, porosity, tracheid size) for chemically modified wood. Based on

this microstructural study, it was found that no changes in porosity

occurred for acetylated samples compared to the unmodified one which is

consistent with wetting results. Finally, the effect of surface topography

on wetting properties was examined on hydrophobised wood surfaces by

combining liquid flame spray and plasma polymerisation methods. It was

concluded that the micro-nano rough surface created by the combination

of both methods gave higher CA, lower CA hysteresis and increased

forced wetting durability compared to the micro-scale rough wood

surfaces treated by plasma polymerisation alone.

Page 79: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

FUTURE WORK | 59

5. Future work

The main technique developed (multicycle Wilhelmy plate) in this thesis

work was utilized for investigating wood veneers as an example of a

swellable porous substrate. However, it was shown that the wetting

properties of wood veneers were highly affected by variation in replicates

due to sample preparation and sample size. This was a main reason for

scattered results, especially for evaluation with the suggested perimeter

model. In order to generalize the application of this method for porous

materials, a swellable material of less variation (e.g. a swellable polymer)

can be used to examine the accuracy of the multicycle Wilhelmy plate

method, and particularly for evaluation of the perimeter model.

Additionally, all multicycle experiments were performed with a limited

number of cycles (20 cycles). In many applications wood/water

interaction is needed to be determined in order to examine the wood

performance. Longer wetting measurements are recommended to

investigate the dynamic wetting properties of wood. For instance, for

surface modified samples a longer wetting measurement (e.g. 200 cycles)

can be done to evaluate dynamic durability of hydrophobised layers.

In wetting curves for wood samples, three different kinetic regimes

were observed. To achieve better knowledge of this sorption process,

research can be done to provide a kinetic model for these wetting

regimes. Such a model should give further insight into the sorption

process. This can be done by using various wood species with different

wetting properties such as initial contact angle and spreading time,

capillary uptake and swelling level.

One suggestion to achieve better understanding of the impact of

surface chemical composition on wetting properties can be to use

different wood samples with different chemistry (level of cellulose, lignin

and extractives) and study their wetting properties in terms of capillary

uptake and swelling.

Page 80: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

60 | FUTURE WORK

The 3D x-ray computed tomography was found to be an interesting

technique for microstructural investigation of modified wood and

particularly to relate the wetting results to anatomical features. To gain

more understanding of how chemical modification is effective to

minimize the wood hygroscopicity, a real-time test is recommended. This

can be done by scanning the unmodified and modified wood samples

when they are exposed to water or octane at the same time as performing

the wetting measurements. The liquid movement inside the wood sample

and the level of swelling in the wood cell walls can then be investigated.

Additionally, in order to investigate the effect of swelling on unmodified

and modified wood samples, a humidity chamber can be used to have

different moisture content when samples are scanned. For instance, the

microstructural properties of vapor saturated samples can be compared

to those for the samples at atmospheric conditions, and consequently it

would be possible to study how the modification reduces the level of

swelling in cell walls.

Page 81: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

ACKNOWLEDGEMENTS | 61

6. Acknowledgements

Many people were involved in my work and I am really glad to have their support during my

doctoral study.

First of all, I would like to thank my supervisors, without your guidance, encouragement, knowledge

and wisdom this project would probably not have been completed. I was very proud of having a

multidisciplinary supervision team from surface chemistry to wood science. My main supervisor,

Prof. Agne Swerin, thanks for giving me the opportunity to carry out this work, thanks for your

continuous support and guidance, both scientifically and personally. My co-supervisor, Prof. Per

Claesson, thanks for all your inspiring ideas and motivation, and for your talent in research. My co-

supervisor, Prof. Magnus Wålinder, thanks for introducing me to the wood science and for all

fantastic scientific discussions.

I am very thankful to the main financer, Nils and Dorthi Troëdsson Foundation. I also acknowledge

Swedish Research Council FORMAS within the EnWoBio project (2014-172) and RISE Research

Institutes of Sweden AB for the financial support of the last year of the PhD work. Financial support

of internship program at Ghent University by an Integrative European Research Infrastructure

project (Trees4Future) is gratefully acknowledged.

I would like to thank all co-authors, Dr. Mikko Tuominen, Golrokh Heydari, Prof. Joris Van Acker,

Prof. Jyrki Makela, Janne Haapanen, for the valuable scientific collaboration. Special thanks go to

Dr. Jan Van den Bulcke, for performing a very effective and interesting research together. I really

missed those intensive days of scanning at the basement, as well as Belgian beers and chocolates.

I would like to thank all my colleges at SP for making such a nice place to work. I would like to

dedicate a special thanks to a few colleagues who have helped me during this research. To Mikael

Sundin, Rodrigo Robinson, Karin Hallstensson, and Anders Svensk for endless support especially

for lab work. To Marie Ernstsson, Johan Andersson and Kenth Johansson for helping me with

analysis and interpretation. To Magnus Eriksson for helping me in writing the abstract of the thesis

in Swedish. To current and past PhD candidates, especially to Asaf, Lukas, Maria, Marine and

Petra. To my office mate, Lina, for the valuable discussions when we both were writing thesis and

for your comments. My sincere thanks also go to my immediate manager, Maria Eklund, for all her

support and advice during these years. I acknowledge Tommy Sebring, Dr. Mats Westin and Dr.

Pia Larsson Brelid at SP Trä for supplying the wood samples.

I would like to thank my colleagues in division of surface and corrosion science and division of

building material at KTH.

Page 82: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

62 | ACKNOWLEDGEMENTS

During these years I have been lucky to have many wonderful friends. Thanks all to keep me always

happy during this journey. Alireza, Eric, Payam, Shaghayegh, Ali, Babak T., Shadi, Esben,

Abolfazl, Ehsan, Babak Gh., Zahra and Neda, you have a special place in my memory.

Dearest Ghazal and Dariush are warmly thanked for being so close, kind and supportive and make

my personal life much more fun and pleasant.

Last but not the least, I would like to thank my family: my wonderful parents in law, who always

inspired me to be a better person and supported me in difficult days, also my brothers, Peyman and

Mazda, for supporting me in life.

.زندگی به من مادر مهربانم پدر عزیزم از همه زحمت های بی دریغتان متشکرم وسپاسگذارم از نشان دادن مسیر درستMy love, Golrokh, thanks for your always positivity, for your continued support and love during

these PhD years and also throughout our life together, you’ve helped me more than anyone else. I

love you!

Golsa, my sweetest, you cannot imagine how much energy I have gotten with your warm hugs after

a long working day. Words can’t describe my feeling for you. I love your smart eyes!

Page 83: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

REFRENCES | 63

References

[1] Kollmann, F.F.P., W.A. Côté, Jr., 1984. Principles of Wood Science and Technology,

Vol. I: Solid Wood. Reprint. Springer-Verlag, Berlin.

[2] Hon D.N.-S., N. Shiraishi (Eds)., 1991. Wood and cellulosic chemistry. Marcel Dekker,

New York.

[3] Rowell, R.M. (Ed.), 2012. Handbook of wood chemistry and wood composites, Second

edition. Taylor & Francis, Boca Raton, Florida.

[4] Ek, M., G. Gellerstedt, G. Henriksson., 2009. Wood Chemistry and Biotechnology.

Walter de Gruyter.

[5] Petrič, M., and P. Oven, Determination of Wettability of Wood and Its Significance in

Wood Science and Technology: A Critical Review. Rev Adhes Adhes, 2015. 3:121-

187.

[6] Back, E.L., The locations and morphology of resin components in the wood. In: Pitch

Control, Wood Resins and Deresination, Tappi Press Atlanta, 2000. p 43.

[7] Nuopponen, M., S. Willför, A.S. Jääskeläinen, T. Vuorinen, A UV resonance Raman

(UVRR) spectroscopic study on the extractable compounds in Scots pine (Pinus

sylvestris) wood: Part II. Hydrophilic compounds. Spectrochimica Acta Part A: Mol

Biomol Spec, 2004. 60, 2963-2968.

[8] Pereira, H., Graca, J., J.C. Ridrigues, 2003. Wood chemistry in relation to quality. in:

WoodQuality and Its Biological Basis, J.R. Barnett and G. Jeronimidis, (Eds.), pp.

53–86, Blackwell Publishing Ltd, Oxford.

[9] Hill, C., 2006. Chapter 2: Modifying the properties of wood. In: Wood modification:

Chemical, thermal and other processes, p. 19-44. John Wiley and Sons Ltd,

Chichester, England.

[10] Sjöström, E., 1993. Chapter 4 – LIGNIN. In: Wood Chemistry (Second Edition),

Sjöström E., Ed. Academic Press: San Diego, pp 71-98.

[11] Haygreen, J.G., J.L. Bowyer, 1996. Forest Products and Wood Science - An

Introduction, 3rd edition, Chapter 3: Composition and structure of wood cells. p. 45-

63, Iowa State University Press, Ames, United Stated.

[12] Mark, R.E., 1967. Cell wall mechanics of tracheids. Yale Univ. Press, New Haven,

London.

[13] Sjöström, E., 1993. Appendix. In Wood Chemistry (Second Edition), Sjöström E., Ed.

Academic Press: San Diego, pp 249.

[14] Källbom, S., 2015. Surface characterisation of thermally modified spruce wood and

influence of water vapor soprtion. Licentiate thesis, Stockholm.

[15] Sjöström, E., 1993. Chapter 5 - EXTRACTIVES. In: Wood Chemistry (Second Edition),

Sjöström E., Ed. Academic Press: San Diego, pp 90-108.

[16] Mantanis, G.I., R.A. Young, R.M. Rowell, Swelling of wood. Wood Sci Technol, 1994.

28, 119-134.

[17] Stamm, A.J., W.K. Loughborough, Variation in shrinking and swelling of wood. T Am

Soc Mech Eng, 1942. 64, 379-386.

[18] Mantanis, G.I., R.A. Young, R.M. Rowell, Swelling of wood. Part III. Effect of

temperature and extractives on rate and maximum swelling. Holzforschung, 1995.

49, 239-248.

[19] Berg, J.C., 1993. Chapter 2. Role of acid-base interactions in wetting and related

phenomena. In: Wettability. Ed. J.C. Berg. Marcel Dekker, New York, pp. 75–148.

Page 84: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

64 | REFRENCES

[20] Collett, B., A review of surface and interfacial adhesion in wood science and related

fields. Wood Sci Technol, 1972. 6: 1-42.

[21] Young, T., An essay on the cohesion of fluids. Phil Trans Roy Soc. London, 1805. 95:

65–87.

[22] Wenzel, R.N., Resistance of solid surfaces to wetting by water. Ind Eng. Chem., 1936.

28: 988.

[23] Oliver, J.F., C. Huh, S.G. Mason, An experimental study of some effects of solid

surface roughness on wetting. Colloid Surface, 1980. 1: 79–104.

[24] Good, R.J., 1993. Contact angle, wetting, and adhesion: a critical review. In: Contact

Angle, Wettability and Adhesion. Ed. K.L. Mittal. VSP, Utrecht, The Netherlands, pp.

3–36.

[25] Schwartz, L.W., S. Garoff, Contact angle hysteresis on heterogeneous surfaces.

Langmuir, 1985. 1: 219–230.

[26] Fadeev, A.Y., T.J. McCarthy, Trialkylsilane monolayers covalently attached to silicon

surfaces: wettability studies indicating that molecular topography contributes to

contact angle hysteresis. Langmuir, 1999. 15: 3759–3766.

[27] Neumann, A.W., Contact angles and their temperature dependence: thermodynamic

status, measurement, interpretation and application. Adv Colloid Interfac, 1974. 4:

105–191.

[28] Holly, F.J., M.F. Refojo., Wettability of hydrogels I. Poly (2-hydroxyethyl methacrylate).

J Biomed Mater Res, 1975. 9: 315–326.

[29] Yasuda, T., M. Miyama, H. Yasuda, Effect of water immersion on surface configuration

of an ethylene–vinyl alcohol copolymer. Langmuir, 1994. 10: 583–585.

[30] Timmons, C.O., W.A. Zisman. Effect of liquid structure on contact angle hysteresis. J

Colloid Interf Sci, 1966. 22: 165–171.

[31] Sedev, R.V., J.G. Petrov, A.W. Neumann, Effect of swelling of a polymer surface on

advancing and receding contact angles. J Colloid Interf Sci, 1996. 180: 36–42.

[32] Lam, C.N.C., R. Wu, D. Li, M.L. Hair, A.W. Neumann, Study of the advancing and

receding contact angles: liquid sorption as a cause of contact angle hystersis. Adv

Colloid Interfac, 2002. 96: 169–191.

[33] Gray, V.R., The wettability of wood. Forest Prod J, 1962. 12: 452–461.

[34] Cheng, E., X. Sun, Effects of wood-surface roughness, adhesive viscosity and

processing pressure on adhesion strength of protein adhesive. J Adhes Sci Technol,

2006. 20: 997–1017.

[35] Acda, M.N., E.E. Devera, R.J. Cabangon, H.J. Ramos, Effects of plasma modification

on adhesion properties of wood. Int J Adhes Adhes, 2012. 32: 70–75.

[36] Saha, S., D. Kocaefe, C. Krause, Y. Boluk, A. Pichette, Enhancing exterior durability of

heat-treated jack pine by photo-stabilization by acrylic polyurethane coating using

bark extract. Part 2: Wetting characteristics and fluorescence microscopy analysis.

Prog Org Coat, 2013. 158: 504–512.

[37] De Vetter, L., J. Van den Bulcke, J. Van Acker, Impact of organosilicon treatments on

the wood-water relationship of solid wood. Holzforscung, 2010. 64: 463–468.

[38] Wålinder, M.E.P., Study of Lewis acid-base properties of wood by contact angle

analysis. Holzforschung, 2002. 56: 363−371.

[39] Petrič, M., Surface modification of wood. Rev Adhes Adhes, 2013. 1: 216-247.

Page 85: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

REFRENCES | 65

[40] Mohammed-Ziegler, I., Z. Hórvölgyi, A. Tóth, W. Forsling, A. Holmgren, Wettability and

spectroscopic characterization of silylated wood samples. Polym Advan Technol,

2006. 17: 932-939.

[41] Bryne, L.E., M.E.P. Wålinder., Ageing of modified wood. Part 1: Wetting properties of

acetylated, furfurylated, and thermally modified wood. Holzforschung, 2010. 64:

295–304.

[42] Herrera, R., X. Erdocia, R. Llano-Ponte, J. Labidi, Characterization of hydrothermally

treated wood in relation to changes on its chemical composition and physical

properties. J Anal Appl Pyrol, 2014. 107: 256–266.

[43] Pétrissans, M., P. Gérardin, I. El Bakali, M. Serraj, Wettability of heat-treated wood.

Holzforschung, 2003. 57: 301−307.

[44] Jennings, J.D., A. Zink–Sharp, C.E. Frazier, F.A. Kamke, Properties of compression

densified wood, Part II: Surface energy. J Adhes Sci Technol, 2006. 20: 335–344.

[45] Johnson, R.E., Jr. and Robert H. Dettre, 1993. Chapter 1. Wetting of low-energy

surfaces. In: Wettability. Ed. J.C. Berg. Marcel Dekker, New York, pp. 1–74.

[46] Neumann, A.W., R.J. Good, 1979. Chapter 2. Techniques of measuring contact angles.

In: Surface and Colloid Science. Volume II. Eds. R.J. Good and R.R Stromberg.

Plenum Press, NewYork, p. 47-51.

[47] Scheikl, M., M. Dunky, Measurement of dynamic and static contact angles on wood for

the determination of its surface tension and the penetration of liquids into the wood

surface. Holzforschung, 1998. 52: 89-94.

[48] Wålinder, M.E.P., G. Ström, Measurement of wood wettability by the Wilhelmy method.

Part 2. Determination of apparent contact angles. Holzforschung, 2001. 55: 33-41.

[49] Gardner, D.J., N.C. Generalla, D.W. Gunnells, M.P. Wolcott, Dynamic wettability of

wood. Langmuir, 1991. 7: 2498-2502.

[50] Chen, C.M., Effect of extractive removal on adhesion and wettability of some tropical

woods. Forest Prod J, 1970. 20: 36–41.

[51] Nahum, T., H. Dodiuk, A. Dotan, S. Kenig, J.P. Lellouche, Superhydrophobic durable

coating based on UV-photoreactive silica nanoparticles. J Appl Polym Sci, 2014.

131: DOI: 10.1002/app.41122 (1–8).

[52] Rodriguez-Valverde, M.A., M.A. Cabreizo-Vilchez, P.Rosales-Lopez, A. Paez-Duenas,

R. Hidalgo Alvarez, Contact angle measurements on two (wood and stone) nonideal

surfaces. Colloid Surface A, 2002. 206: 485–495.

[53] Mazurek, A., S.J. Pogorzelski, K. Boniewicz-Szmyt, Adsorption of natural surfactants

present in sea waters at surfaces of minerals: Contact angle measurements.

Oceanologia, 2009. 51: 377–403.

[54] van Oss C.J., R.F. Giese, Z. Li, K. Murphy, J. Norris, M.K. Chaudhury, R.J. Good,

Determination of contact angles and pore sizes of porous media by column and thin

layer wicking. J Adhes Sci Technol, 1992. 6: 413–428.

[55] Sinn, G., J. Sandak, T. Ramananantoandro, Properties of wood surfaces-

Characterisation and measurement. A review. COST Action E35 2004-2008: Wood

machining - Micromechanics and fracture. Holzforschung, 2009. 63: 196-203.

[56] Gaonkar, A.G., R.D. Neuman, The effect of wettability of Wilhelmy plate and du Nouey

ring on interfacial tension measurements in solvent extraction systems. J Colloid

Interfac, 1984. 98: 112–119.

[57] Seebergh, J.E., J.C. Berg, A comparison of force and optical techniques for the

measurement of dynamic contact angles. Chem Eng Sci, 1992. 47: 4468–4470.

Page 86: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

66 | REFRENCES

[58] Lander, L.M., L.M. Siewierski, W.J. Brittain, E.A. Vogler, A systematic comparison of

contact angle methods. Langmuir, 1993. 9: 2237–2239.

[59] Nguyen, T., W.E. Johns, Polar and dispersion force contributions to the total surface

free energy of wood. Wood Sci Technol, 1978. 12: 75–86.

[60] Hse, C.-Y., Wettability of Southern pine veneer by phenol formaldehyde wood

adhesives. Forest Prod J, 1972. 22: 51–56.

[61] Nussbaum, R.M., Natural surface inactivation of Scots pine and Norway spruce

evaluated by contact angle measurements. Wettability of Sitka spruce timber stored

in a Scottish sawmill yard. Holz Roh Werkst, 1999. 57: 419–424.

[62] Esteves, B., A. Velez Marquez, I. Domingos, H. Pereira, Influence of steam heating

on the properties of pine (Pinus pinaster) and eucalypt (Eucalyptus globulus) wood.

Wood Sci Technol, 2007. 41: 193–207.

[63] Huang, X., D. Kocaefe, Y. Kocaefe, Y. Boluk, A. Pichette, Changes in wettability of

heat-treated wood due to artificial weathering. Wood Sci Technol, 2012. 46: 1215–

1237.

[64] Kutnar, A., L. Rautkari, K. Laine, M, Hughes, Thermodynamic characteristics of

surface densified solid Scots pine wood. Eur J Wood Prod, 2012. 70: 727–734.

[65] Ayrilmis, N., Effect of fire retardants on surface roughness and wettability of wood

plastic composite panels. Bioresources, 2011. 6: 3178-3187.

[66] Lu, J.Z., Q. Wu, Surface characterization of chemically modified wood: dynamic

wettability. Wood Fiber Sci, 2006. 38: 497-511.

[67] Shupe, T.E., C.Y. Hse, E.T. Choong, L.H. Groom, Effect of wood grain and veneer side

on loblolly pine. Forest Prod J, 1998. 48: 95-97.

[68] Wang, S., Y. Zhang, C. Xing, Effect of drying method on the surface wettability of wood

strands. Holz als Roh- und Werkstoff, 2007. 65: 437-442.

[69] Liptáková, E., J. Kúdela, Analysis of the wood wetting process. Holzforschung, 1994.

48: 139–144.

[70] Wålinder, M.E.P., D.J. Gardner. Factors influencing contact angle measurements on

wood particles by column wicking. J Adhes Sci Technol, 1999. 13: 1363–1374.

[71] Nussbaum, R.M., M. Sterley, The effect of wood extractive content on glue adhesion

and surface wettability of wood. Wood and Fiber Science, 2002. 34: 57-71.

[72] Wålinder, M.E.P., I. Johansson, Measurement of wood wettability by the Wilhelmy

method. Part 1. Contamination of probe liquids by extractives. Holzforschung, 2001.

55: 21-32.

[73] Shen, Q., J. Nylund, J.B. Rosenholm, Estimation of the surface energy and acid-base

properties of wood by means of wetting method. Holzforschung, 1998. 52: 521−529.

[74] Son, J., D.J. Gardner, Dimensional stability measurements of thin wood veneers using

the wilhelmy plate technique. Wood Fiber Sci, 2004. 36: 98-106.

[75] de Meijer, M., SHaemers, W. Cobben, H. Militz, Surface Energy Determinations of

Wood:  Comparison of Methods and Wood Species. Langmuir, 2000. 16: 9352-9359.

[76] Wålinder, M.E.P., P.L. Brelid, K. Segerholm, C.J. Long II, J.P. Dickerson, Wettability of

acetylated Southern yellow pine. In. Wood Prod J, 2013. 4: 197-203.

[77] Fredriksson, M., O. Lindgren, End grain water absorption and redistribution in slow-

grown and fast-grown Norway spruce (Picea abies (L.) Karst.) heartwood and

sapwood. Wood Mater Sci Eng, 2013. 8: 245-252.

Page 87: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

REFRENCES | 67

[78] Rowell, R.M., Simonson, R., Hess, S., Plackett, D.V., Cronshaw, D., Dunningham, E.,

Acetyl distribution in acetylated whole wood and reactivity ofisolated wood cell-wall

components to acetic anhydride. Wood Fiber Sci, 1994. 26: 11-18.

[79] Rowell, R.M., R. Simonson, A.M. Tillman, A process for improving dimensional stability

and biological resistance of lignocellulosic materials. 1985. European Patent

0213252.

[80] Stevens, M., J. Schalck, J.V. Raemdonck, Chemical modification of wood by vapour-

phase treatment with formaldehyde and sulfur dioxide. Int J Wood Pres, 1979. 1: 57-

68.

[81] Lande, S., M. Westin, M. Schneider, Development of modified wood products based on

furan chemistry. Mol Cryst Liq Cryst, 2008. 484: 1/[367]-12/[378].

[82] Rowell, R.M., Acetylation of wood: Journey from analytical technique to commercial

reality. Forest Prod J, 2006. 56: 4-12.

[83] Venås T., L. Thygesen, S. Barsberg, Chemical reactions involved in furfurylation of

solid wood—an investigation by ATRIR spectroscopy. Int Res Group Wood Pre.,

2006. Document No. IRG/WP/40347. Stockholm.

[84] Nordstierna, L., S. Lande, M. Westin, O. Karlsson, I. Fúró, Towards novel wood-based

materials: Chemical bonds between lignin-like model molecules and poly(furfuryl

alcohol) studied by NMR. Holzforschung, 2008. 62: 709-713.

[85] Lande, S., M. Eikenes, M. Westin, M. Schneider, Furfurylation of Wood: Chemistry,

Properties, and Commercialization. In: Development of Commercial Wood

Preservatives. Eds. Schultz, T.P., Militz, H., Freeman, M.H., Goodell, B., Nicholas,

D.D. ACS Symp Ser No 982, 2008. pp. 337-355.

[86] Podgorski, L., B. Chevet, L. Onic, A. Merlin, Modification of wood wettability by plasma

and corona treatments. Int J Adhes Adhes, 2000. 20: 103-111.

[87] Wang, S., C. Liu, G. Liu, M. Zhang, J. Li, C. Wang, Fabrication of superhydrophobic

wood surface by a sol–gel process. Appl Surf Sci, 2011. 258: 806-810.

[88] Denes, A.R., M.A. Tshabalala, R. Rowell, F. Denes, R.A. Young,

Hexamethyldisiloxane-plasma coating of wood surfaces for creating water repellent

characteristics. Holzforschung, 1999. 53: 318-326.

[89] Podgorski, L., C. Bousta, F. Schambourg, J. Maguin, B. Chevet, Surface modification of

wood by plasma polymerisation. Pigm Resin Tech, 2002. 31: 33-40.

[90] Fu, Y., H. Yu, Q. Sun, G. Li, Y. Liu, Testing of the superhydrophobicity of a zinc oxide

nanorod array coating on wood surface prepared by hydrothermal treatment.

Holzforschung, 2012. 66: 739-744.

[91] Mamiński, M.Ł., K. Mierzejewska, P. Borysiuk, P. Parzuchowski, P. Boruszewski,

Surface properties of octadecanol-grafted pine veneers. Int J Adhes Adhes, 2009.

29: 781-784.

[92] Shupe, T., C. Piao, C. Lucas, The termiticidal properties of superhydrophobic wood

surfaces treated with ZnO nanorods. Eur J Wood Prod, 2012. 70: 531-535.

[93] Wilhelmy J., Über die Abhängigkeit der Kapillaritäts-Konstanten des Alkohols von

Substanz und Gestalt des Benetzten Festen Körpers. Ann Physik, 1863. 119: 177–

217.

[94] Kwok, D.Y., A.W. Neumann, Contact angle measurement and contact angle

interpretation. Advan Colloid Interfac, 1999. 81L 167-249.

[95] Binning, G., C.F. Quete, C. Gerber, Atomic force microscope, 1986. 56L 930-933.

Page 88: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

68 | REFRENCES

[96] Sedighi Moghaddam, M., M.E.P. Wålinder, P.M. Claesson, A. Swerin, Multicycle

Wilhelmy Plate Method for Wetting Properties, Swelling and Liquid Sorption of

Wood. Langmuir, 2013. 29: 12145-12153.

[97] Bastani, A., S. Adamopoulos, H. Militz, Water uptake and wetting behaviour of

furfurylated, N-methylol melamine modified and heat-treated wood. Eur J Wood

Prod, 2015. 73: 1-8.

[98] Zanini, S., C. Riccardi, M. Orlandi, V. Fornara, M.P. Colombini, D. Donato, S. Legnaioli,

V. Palleschi, Wood coated with plasma-polymer for water repellence. Wood Sci

Technol, 2008. 42: 149-160.

[99] Sedighi Moghaddam, M., M.E.P. Wålinder, P.M. Claesson, A. Swerin, Wettability and

swelling of acetylated and furfurylated wood analyzed by multicycle Wilhelmy plate

method. Holzforschung, online February 2015. doi:10.1515/hf-2014-0196.

[100] Metsä-Kortelainen, S., H. Viitanen, Wettability of sapwood and heartwood of thermally

modified Norway spruce and Scots pine. Eur J Wood Prod, 2012. 70: 135-139.

[101] Metsä-Kortelainen, S., T. Antikainen, P. Viitaniemi, The water absorption of sapwood

and heartwood of Scots pine and Norway spruce heat-treated at 170 °C, 190 °C,

210 °C and 230 °C. Holz als Roh- und Werkstoff, 2006. 64: 192-197.

[102] Baysal, E., S.K. Ozaki, M.S. Yalinkilic, Dimensional stabilization of wood treated with

furfuryl alcohol catalysed by borates. Wood Sci Technol, 2004. 38: 405-415.

[103] Epmeier, H., M. Westin, A. Rapp, Differently modified wood: comparison of some

selected properties. Scand J For Res, 2004. 19(sup5): 31-37.

[104] Inari, G.N., M. Petrissans, J. Lambert, J.J. Ehrhardt, P. Gérardin, XPS characterization

of wood chemical composition after heat-treatment. Surf Interface Anal, 2006. 38:

1336-1342.

[105] Popescu, C.-M., C.-M. Tibirna, C. Vasile, XPS characterization of naturally aged

wood. Appl Surf Sci, 2009. 256: 1355-1360.

[106] Bryne, L.E., J. Lausmaa, M. Ernstsson, F. Englund, M.E.P. Wålinder, Ageing of

modified wood. Part 2: Determination of surface composition of acetylated,

furfurylated, and thermally modified wood by XPS and ToF-SIMS. Holzforschung,

2010. 64: 305-313.

[107] Rautkari, L., T. Hänninen, L.S. Johansson, M. Hughes, A study by X-ray photoelectron

spectroscopy (XPS) of the chemistry of the surface of Scots pine (Pinus sylvestris L.)

modified by friction. Holzforschung, 2012. 66: 93-96.

[108] Gindl, M., A. Reiterer, G. Sinn, S.E. Stanzl-Tschegg, Effects of surface ageing on

wettability, surface chemistry, and adhesion of wood. Holz als Roh- und Werkstoff,

2004. 62: 273-280.

[109] Englund, F., L.E. Bryne, M. Ernstsson, J. Lausmaa, M.E.P. Wålinder, Spectroscopic

studies of surface chemical composition and wettability of modified wood. Wood

Mater Scie Eng, 2009. 4: 80-85.

[110] Gustafsson, J., L. Ciovica, J. Peltonen, The ultrastructure of spruce kraft pulps studied

by atomic force microscopy (AFM) and X-ray photoelectron spectroscopy (XPS).

Polymer, 2003. 44: 661-670.

[111] Laine, J., P. Stenius, G. Carlsson, G. Ström, Surface characterization of unbleached

kraft pulps by means of ESCA. Cellulose 1994, 1: 145-160.

[112] Taylor, A., B. Gartner, J. Morrell, Heartwood Formation and Natural Durability—A

Review. Wood Fiber Sci, 2002. 34: 587-611.

[113] Luner, P., M. Sandell, The wetting of cellulose and wood hemicelluloses. J Polym Sci

Pol Sym, 1969. 28: 115-142.

Page 89: MAZIAR SEDIGHI MOGHADDAM863723/...spelar vätbarheten av trä och växelverkan mellan trä och vätskor en viktig roll när det gäller att utforma och utvärdera olika processer och

REFRENCES | 69

[114] Hodgson, K., J. Berg, Dynamic Wettability Properties of Single Wood Pulp Fibers and

Their Relationship to Absorbency. Wood Fiber Sci, 1988. 20: 3-17.

[115] Holmberg, M., J. Berg, S. Stemme, L. Ödberg, J. Rasmusson, P.M. Claesson, Surface

Force Studies of Langmuir–Blodgett Cellulose Films. J Colloid Interfac, 1997. 186:

369-381.

[116] Lee, S.B, P. Luner, The wetting and interfaeial properties of lignins. Tappi J, 1972. 55:

116-121.

[117] Laschimke, R., Investigation of the wetting behaviour of natural lignin - a contribution

to the cohesion theory of water transport in plants. Thermochim Acta, 1989. 151: 35-

56.

[118] Pykönen, M., K. Johansson, M. Dubreuil, D. Vangeneugden, G. Ström, P. Fardim, M.

Toivakka, Evaluation of plasma-deposited hydrophobic coatings on pigment-coated

paper for reduced dampening water absorption. J Adhes Sci Technol, 2010. 24: 511-

537.

[119] Nguila Inari, G., M. Pétrissans, S. Dumarcay, J. Lambert, J.J. Ehrhardt, M. Šernek, P.

Gérardin, Limitation of XPS for analysis of wood species containing high amounts of

lipophilic extractives. Wood Sci Tech, 2011. 45: 369-382.

[120] Sedighi-Gilani, M., M. Griffa, D. Mannes, E. Lehmann, J. Carmeliet, D. Derome,

Visualization and quantification of liquid water transport in softwood by means of

neutron radiography. Int J Heat Mass Tran, 2012. 55: 6211-6221.

[121] Sander, C., E.P.J. Beckers, H. Militz, W. van Veenendaal, Analysis of acetylated wood

by electron microscopy. Wood Sci Technol, 2003. 37: 39-46.

[122] Wålinder M.E.P., A. Omidvar, J. Seltman, K. Segerholm, Micromorphological studies

of modified wood using a surface preparation technique based on ultraviolet laser

ablation. Wood Mater Sci Eng, 2009. 4: 46-51.

[123] Murakami, D., H. Jinnai, A. Takahara, Wetting transition from the Cassie–Baxter state

to the Wenzel state on textured polymer surfaces. Langmuir, 2014. 30: 2061-2067.