You are on page 1of 29

2

Fundamentals of Adhesion

2.1 Introduction
Adhesion is a multidisciplinary science involving various subjects such
as rheology, materials science, organic chemistry, polymer science, and
mechanics. Study of fundamentals of adhesion is essential as it leads to bet-
ter understanding of the factors controlling the performance of the bonded
assemblies [1].
Wood is a complex substrate, and it is hard to understand why some
adhesives work better than other adhesives, especially under stringent
durability tests.
The recent trend in the wood industry is to use smaller-diameter logs
and employ other lignocellulosic raw materials to produce more versatile
and environmentally acceptable green engineered wood products. This in
turn increases the complexity in the choice of adhesives. In order to pro-
vide a scientific basis to make the correct choice of adhesives and their
formulations, a study of the fundamentals of adhesion is essential.
A clear understanding of wood adhesion mechanisms will enable pro-
duction of better adhesive and formulation systems suitable for a wider
array of wood composite materials. The study of the fundamentals of
wood adhesion is essentially distinctive and unique and involves multi-
disciplinary sciences with respect to both the adherend, the adhesives, and
their interactions. The uniqueness of wood as an adherend by virtue of
possessing a hierarchical structure has already been dealt with in detail in
Chapter 1. In this respect, wood differs significantly from other substrates
such as metals, plastics, elastomers, etc. Surface science, rheology, materi-
als science, surface chemistry and surface morphology, organic chemistry,
polymer science and polymer characterization, and solid mechanics and
interaction between polymers and wood—all contribute to the develop-
ment and understanding of the adhesion phenomenon.

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (31–59) © 2019
Scrivener Publishing LLC

31
32 Adhesives for Wood and Lignocellulosic Materials

Further, these studies will enable

• The identification of practical problems and root causes for


adhesion failure and provide practical preventive solution
from the knowledge of the nature of wood–adhesive inter-
action, i.e., a scientific approach to troubleshooting.
• Optimization of the performance of existing adhesives and
to develop new adhesives to meet stringent environmental
regulations.
• The increase in the durability of bonded wood products by
precisely understanding the role of internal and external
stresses to which wood bond lines are subjected.
• The development of new technologies based on the insights
gained from the knowledge of the basic principles, which
can be applied efficiently for bonding difficult to bond
wood species and preservative treated wood. This is so that
the carbon sequestration is possible for prolonged periods
of time to reduce the global warming potential of wood
products.

2.2 Definitions
We should first define the terms adhesive, adhesion cohesion, and other
related terms in order to understand their individual role in determining
the effectiveness of bonding.

2.2.1 Adhesion
Adhesion is defined as the state in which two surfaces are held together by
interfacial forces that may consist of valence forces or interlocking action,
or both. Adhesion is further classified as mechanical adhesion and specific
adhesion. Specific adhesion between two surfaces is caused by the valence
forces of the same type as those that give rise to cohesion, as opposed to
mechanical adhesion in which the adhesive holds the parts together by an
mechanical interlocking.

2.2.2 Cohesion
Cohesion is defined as the internal strength of an adhesive as a result of a
variety of interactions within the adhesive.
Fundamentals of Adhesion 33

2.2.3 Adhesive
ASTM defines an adhesive as a substance capable of holding materials
together by surface attachment.

2.2.4 Adherend
Adhered, also called a substrate, is defined as a body that is held to another
body by an adhesive used interchangeably. Various descriptive adjectives are
applied to the term adhesive to indicate certain characteristics as follows:
(1)  physical form, that is, liquid adhesive, tape adhesive, etc.; (2)  chemical
type, that is, silicate adhesive, resin adhesive, etc.; (3) materials bonded, that is,
paper adhesive, metal–plastic adhesive, can label adhesive, etc.; (4) condition
of use, that is, hot setting adhesive, room temperature setting adhesive, etc.

2.2.5 Bonding
Bonding is the joining of two substrates using an adhesive. According to
DIN EN 923, an adhesive is defined as a non-metallic binder that acts via
adhesion and cohesion. ASTM D907-06 defines an adhesive as “a substance
capable of holding materials together by surface attachment”. A material
attached using adhesive is called an adherend.

2.2.6 Adhesive, Assembly


Adhesive, assembly—an adhesive that can be used for bonding parts
together, such as in the manufacture of a boat, airplane, furniture, and the
like. Note: The term assembly adhesive is commonly used in the wood
industry to distinguish such adhesives (formerly called “joint glues”) from
those used in making plywood (sometimes called “veneer glues”).

2.3 Mechanism of Adhesion


The role of an adhesive for wood is to transfer and distribute loads between
components, thereby increasing the strength and stiffness of wood prod-
ucts [6].
This is achieved through the following three basic types of adhesion:

1. Specific Adhesion—Bonding between the adhesive and the


adherend is due to chemical reaction.
34 Adhesives for Wood and Lignocellulosic Materials

2. Mechanical Adhesion—occurs due to mechanical anchorage.


3. Effective Adhesion—combines specific and mechanical
adhesion for optimum joining strength.

One should distinguish between adhesion and cohesion.


Cohesion as defined earlier is the attraction of molecules and groups
within the adhesive (or other material) that holds the adhesive molecules
together. The combination of adhesion and cohesive strength determines
the bonding effectiveness. An adhesive bond fails if either the adhesive
separates from the substrate (interfacial adhesion failure) or the adhesive
breaks apart (cohesive failure). The adhesive and cohesive strengths of
some adhesives are high enough that the cohesive strength of the substrate
fails before the adhesive bond.

2.3.1 Specific Adhesion


Specific adhesion involves the bond created by chemical means, rather
than mechanical, as a result of the molecular attraction between the sur-
faces in contact. This can be ionic, covalent, or induced by any other inter-
molecular forces (Figure 2.1), as described below:

(a) Coulombic (ionic) or hydrogen bonding


Hydrogen bonds occur in molecules that have H–F, H–O, and
H–N bonds. Basically, this strong intermolecular force is due to
strong dipole–dipole forces.
Besides the above, there can exist non-covalent and non-
electrostatic interactions (apolar interactions) between neu-
tral atoms and molecules [2, 3]. However, they are not as
strong as Coulombic (ionic) or hydrogen-bond interactions.
They are ubiquitous and are always attractive between like
particles.
(b) Apolar interactions
There are three types of intermolecular forces that occur in
chemical compounds. These forces cause molecules or groups of
molecules to be attracted to one another, thus affecting many of
their properties. Collectively known as the van der Waals forces,
these electrodynamic intermolecular forces originated from
three distinct interactions. These are (a) Keesom (permanent–
permanent dipoles) interaction (b) Debye (permanent-induced
dipoles) force, and (c) London dispersion force (fluctuating
Fundamentals of Adhesion 35

+40

Repulsion
+30
+20 Van-der-Waals forces
+10 1 2 3 4 5 6
0
Distance r
–10
[0, 1 nm]
–20
Hydrogen bonds
Attraction

–30
–40
Chemical bonds
–50
–60

Figure 2.1 Potential energy diagram for different forces [4].

dipole-induced dipole interaction) [2]. While these three kinds


of interactions have distinct origins, they have in common the
fact that their interaction energies decay rapidly with the sixth
power of the interatomic or molecular distance. See Sections
2.3.1.1 to 2.3.1.3.
The London dispersion force is the weakest, followed in
increasing strength by dipole–dipole forces and then hydro-
gen bonding. Lewis acid–base interactions can also occur (dis-
cussed later) [3].
The mathematical relationships for the various potential
energies are given below:

2.3.1.1 London Dispersion Force

3 2I
V
4 r6

where α is the polarizability, r is the distance, and I is the first ionization


potential.
The negative sign indicates the attractive interaction.
36 Adhesives for Wood and Lignocellulosic Materials

2.3.1.2 Dipole–Dipole Interaction


Dipole–dipole interaction is between polar molecules. A polar molecule
has an electric dipole moment by virtue of the existence of partial charges
on its atoms. Opposite partial charges attract one another, and, if two polar
molecules are oriented so that the opposite partial charges on the mole-
cules are closer together, then there will be a net attraction between the two
molecules with a potential energy V given by

2 2
2 A B 1
V 2 6
3 (4 0 ) r k BT

μi are the dipole moments


ε is the permitivity of the medium
T is the temperature in Kelvin

2.3.1.3 Dipole–Induced-Dipole Interaction


In the dipole–induced-dipole interaction, the presence of the partial charges
of the polar molecule causes a polarization, or distortion, of the electron
distribution of the other molecule. As a result of this distortion, the second
molecule acquires regions of partial positive and negative charge, and thus
it becomes polar. The partial charges so formed behave just like those of a
permanently polar molecule and interact favorably with their counterparts
in the polar molecule that originally induced them. Hence, the two mole-
cules cohere with a potential energy V given by

2 2
V
r6

where μ is the dipole moment of the polar molecule, α is the polarizability


of non-polar molecule, and r is the distance between them.

2.3.1.4 Ion–Dipole Interaction


An ion-dipole force is an attractive force that results from the electrostatic
attraction between an ion and a neutral molecule that has a dipole. It is
most commonly found in solutions. It is especially important for solutions
Fundamentals of Adhesion 37

of ionic compounds in polar liquids. The potential energy of ion–dipole


interaction is given by

q
V
(4 0 )r 2

where q is the charge on the ion.


One should note that in all the above equations describing the inter-
molecular attractions, the denominator contains the factor r6. Thus, the
types of intermolecular interactions described above occur only at very
small distances, of the order of typical atomic bond lengths (the range of
non-bonding interactions is between 0.3 and 0.5 nm). For interactions to
occur, therefore, the two materials must be able to make intimate contact
with each other (i.e., they must be able to approach within a nanometer).
This is possible if the adhesive wets the substrate efficiently. The types of
interactions and the corresponding energies are given in Table 2.1.

Table 2.1 Bond types and typical bond energies [1].


Type of interaction Energy (kJ/mol) Basis of attraction
Bonding
Ionic 400–4000 Cation–anion
Covalent 150–1100 Nuclei–shared electron pair
Metallic 75–1000 Cations–delocalized electrons
Non-Bonding
Ion–dipole 40–600 Ion charge–dipole charge
Polar bond to hydrogen–dipole
Hydrogen bonding 10–40
charge
Dipole–dipole 5–25 Dipole charges
Ion charge–polarizable
Ion–induced dipole 3–15
electrons
Dipole charge–polarizable
Dipole–induced dipole 2–10
electrons
Interaction between polarizable
Dispersion forces 0.1–40
electrons
38 Adhesives for Wood and Lignocellulosic Materials

2.3.1.5 Hydrogen Bonds


This is an important intermolecular interaction specific to molecules con-
taining an oxygen, nitrogen, or fluorine atom that is attached to a hydro-
gen atom. This interaction is the hydrogen bond, an interaction of the
form A−H···B, where A and B are atoms of any of the three elements men-
tioned above and the hydrogen atom lies on a straight line between the
nuclei of A and B (Figure 2.1).

2.3.1.6 Ionic Bonds


Salts like NaCl.

2.3.1.7 Chemical Bonds


The acid–base character of the substrate may influence the reactivity between
adhesive and substrate. A covalent bond involves shared valence electrons
(Figure 2.1).

2.4 Theories of Adhesion


According to Schultz and Nardin (1994), the main adhesion theories are
as follows:

1. Mechanical interlocking
2. Electronic or electrostatic theory
3. Adsorption (thermodynamic) or wetting theory
4. Diffusion theory
5. Chemical (covalent) bonding theory
6. Theory of weak boundary layers and interphases

The adsorption hypothesis, which explains that adhesion is caused by


intermolecular forces such as van der Waals forces, hydrogen bonds, and
electrostatic interactions, is widely considered to be the most applicable
to wood–polymer adhesion [7]. However, in a porous material like wood,
penetration and mechanical interlocking must also play a significant role
in the bonding process.
Marra [5] described adhesive bond formation in wood-based panels as a
dynamic process consisting of flow, transference, penetration, wetting, and
solidification (cure).
Fundamentals of Adhesion 39

The mechanisms outlined above are not mutually exclusive since one or
more of the above mechanisms can occur simultaneously depending on
the specific conditions prevailing during bonding. The hierarchical cellular
characteristics of wood offer such varied conditions.
The mechanical interlocking theory has long been used to explain wood
bonding [6].
The electronic or electrostatic theory has been applied to wood in
finishing and coating operations, although this adhesion bonding
mechanism needs more fundamental research [21]. The adsorption or
wetting theory has been exhaustively studied on wood over the past 40
years [7, 8].
The diffusion theory is appropriate in wood bonding during the produc-
tion of compressed fibrous materials such as hardboard. The thermoplastic
matrix, namely, lignin, can soften beyond its glass-transition temperature
during the thermal conditions employed during hot pressing. Under these
conditions, lignin can diffuse throughout the fibrous mat and react with
the furfural liberated from hemicelluloses (pentosans) and solidify due to
chemical reaction and hence function as an adhesive.
Besides the diffusion and molecular interpenetration of lignin occurring
during wet process in the hardboard production as mentioned above, there
is also the phenomenon of diffusion of monomers/oligomers of synthetic
resin adhesives such as PF or UF into the wood cells followed by subsequent
polymerization. This is an important concept that speaks of monomers that
penetrate at a molecular level for thermosetting adhesives [9].
While discussing on the theories of adhesion in wood, one should keep
in mind the opposite process (debonding). Weak boundary layers have been
identified as the cause for the premature failure of the adhesive bond. In the
case of wood bonding, the theory of weak boundary layers has also been
proposed and studied. The weak boundary layers can be caused as a result of
the mechanical damages occurring during the machining of wood surfaces.
Further, the impact of surface aging the consequent inactivating of the wood
surfaces [10–12] can also be responsible for the weak boundary layer.

2.4.1 Mechanical Theory


McBain and Hopkins [13] first proposed the concept of “mechanical
adhesion” in their classical paper “On adhesives and adhesive action”.
According to McBain and Hopkins, there are two kinds of adhesion:
mechanical and specific adhesion. Specific adhesion involved interaction
between the adherend surface and the adhesive. This interaction might be
chemical interaction or adsorption.
40 Adhesives for Wood and Lignocellulosic Materials

Mechanical adhesion occurs “whenever a liquid adhesive penetrates


into the porous adherend and solidifies in situ in the pores”. Examples are
adhesion to wood, unglazed porcelain, pumice, and charcoal.
The surface of a substrate is never truly smooth but consists of a maze
of peaks and valleys. This type of topography allows adhesive to penetrate
and fill these valleys, displace the entrapped air, and secure mechanically in
position inside the substrate similar to the operation of the Velcro.
Porosity and roughness of the substrate increase the total area of contact
between the adhesive and the adherend. Hence, roughening the adherend
surface enhances the mechanical interlocking since total effective area
over which the forces of adhesion can develop increases. The mechanical
adhesion theory does not take into account the intrinsic incompatibility
between the adhesive and the substrate.

2.4.1.1 lllustration of Mechanical Adhesion for Wood


In wood adherends, there is a vast array of void spaces as shown in Figure
2.2. Spontaneous surface wetting and capillary effects allow the flow of
the adhesive resin into the cell lumen, vessels, or other interstices fol-
lowed by subsequent hardening of the resin and resulting in mechanical
interlocking. The resin acts to reinforce the surface/interface layers of
wood cells. An adhesive penetration of approximately 6–10 cell diame-
ters (fewer than 100 μm, maximum) is regarded as necessary for optimal
adhesive bonding. Filling the cell lumen with adhesive provides much
larger mechanical interlocks than are available with surface roughness

Rays
Earlywood
Tracheids

Latewood
Tracheids

Figure 2.2 Various wood elements.


Fundamentals of Adhesion 41

for other substrates. Absorption into the cell wall can provide microme-
chanical interlocks and interpenetrating networks [1].

2.5 Electronic Theory


This theory was mainly promoted by Deryaguin [14–16]. If the adhesive
and the substrate have different electronic band structures, there is likely
to be electron transfer on contact between the two surfaces. This results in
the promotion of a double layer of electrical charges at the adhesive sub-
strate interface. Electrostatic forces are formed at the adhesive–adherent
interface. This accounts for the resistance to separation. This theory gath-
ers support from the fact that electrical discharges have been noticed when
an adhesive is peeled from a substrate. Electrostatic adhesion is regarded
as a dominant factor in biological cell adhesion and particle adhesion. No
application of this theory to wood appears possible.
The electronic theory

• Depends on material properties that allow electron transfer


across the interface
• Requires intimate contact/smooth surfaces
• Interactions are very weak and rather insignificant
• Mechanism is not important for wood substrates

2.6 Diffusion Theory


The diffusion theory was proposed in the early 1960s by Voyutskii [17–19].
It states that the intrinsic adhesion of a resin to a polymeric substrate is due
to mutual diffusion of polymer molecules across their interface.
As a result of this interdiffusion of molecules of the adhesive and adher-
end, their interface disappears. Hence, the diffusion theory is applicable
only when both adhesive and adherend are compatible polymers that pos-
sess sufficient mobility and mutual solubility. Solvents or heat welding of
thermoplastic substances is caused by diffusion.
The prerequisite of the diffusion theory is that the polymers of the
adhesive and of the substrate should possess similar values of solubility
parameters.
Several problems are encountered when an attempt is made to apply
the diffusion theory to wood. Basically, wood is not homogeneous in com-
position. It is a cellular composite of three polymers, namely, cellulose,
42 Adhesives for Wood and Lignocellulosic Materials

hemicelluloses, and lignin. Furthermore, cellulose consists of both crys-


talline and amorphous regions. It is clear then from solubility parameter
concepts that some polymers, the amorphous ones such as hemicelluloses
and lignin and the amorphous portion of cellulose, could, under some
conditions, undergo mutual diffusion with the polymer chains of the syn-
thetic adhesives. The crystalline portion of cellulose is not likely to be
involved.
There is one specific instance in the case of wood adhesion [17] in which
interdiffusion appears to exist and is likely to play a significant role in wood
bonding. This is the production of fiberboard by the wet process in which
no adhesive is added. At high moisture content, high temperature and pres-
sure and long pressing times, the glass transition temperatures of lignin are
exceeded. Thus, the lignin in the fibers is mobilized and the interdiffusion
between lignin polymers on different fibers contributes to the bonding of
the fibers together.

2.7 Adsorption/Covalent Bond Theory


The adsorption theory of adhesion, and the most widely accepted one,
in wood science, which is sometimes also called “specific adhesion”
[20], states that an adhesive will adhere to a substrate because of inter-
molecular and interatomic forces between the atoms and molecules of
the two materials. The interatomic and intermolecular forces referred
to can be any type of either primary or secondary valency forces. van
der Waals forces, hydrogen bond, and electrostatic forces are as much
applicable as the primary valence forces such as ionic, covalent metal-
lic coordination bonds. In the case of wood adhesion, however, there
is an age-old mistaken notion that covalent linkages must be present
to ensure good joint strength. In fact, covalent bonding theory was
invoked to explain the durable wood bonding with thermosetting adhe-
sives. But as mentioned by Gardner [21], it is very likely that covalent
bonds between the wood and adhesive are not necessary for durable
wood adhesive bonds.
Calculations carried out by a number of authors based on the secondary
forces involved indicate that the wood bond strength in tension should be
over 100 MPa. This is considerably higher than the experimental values
obtained in the case of several wood adhesives. This discrepancy could be
due to the presence of voids, defects, and the geometrical features of the
test specimen. Pizzi concludes that these studies indicate that the second-
ary valency forces themselves are adequate to explain the practical results
Fundamentals of Adhesion 43

Table 2.2 Comparison of adhesion interactions relative to length scale.


Category of adhesion
Mechanism Type of interaction Length scale
Mechanical Interlocking or entanglement 0.01–1000 μm
Diffusion Interlocking or entanglement 10 nm–2 mm
Electrostatic Charge 0.1–1.0 μm
Covalent bonding Charge 0.1–0.2 nm
Acid-base interaction Charge 0.1–0.4 nm
Lifshitz van der Waals Charge 0.5–1.0 nm

and it is not necessary to invoke the involvement of covalent bonds [20].


An elaborate discussion on the relative importance of the primary and sec-
ondary valence forces has been furnished by Pizzi based on the adhesive
strengths obtained from wood joints and the common wood adhesives
such as phenolics, amino resins, and isocyanates [20].

2.8 Adhesion Interactions as a Function of Length


Scale
It is useful to know the scale of lengths over which the adhesion inter-
actions as described above do occur (Table 2.2) [21]. It is apparent
from Table 2.2 that the adhesive interactions relying on interlocking or
entanglement can occur over larger lengths than the adhesion involv-
ing charge interactions. Most of the charge interactions occur at the
molecular level or the nano-length scale. Electrostatic interactions are
the exception to this generalization. For the purpose of adhesive inter-
actions, they are considered to operate from a nano- to a micron-length
scale.

2.9 Wetting of the Substrate by the Adhesive


It is evident from the previous section that various interactions operate
effectively only when the molecules of the adhesives come as close as pos-
sible to those of the substrate in order that such a proximity will lead to
44 Adhesives for Wood and Lignocellulosic Materials

α
α

(a) (b)

Wetting Dewetting
Adhesion Forces > Cohesive Forces Adhesion Forces < Cohesive Forces
Spreading of the liquid on the The liquid pulls itself together into the
surface of the solid. shape of a droplet.
Contact angle θ: 0 < θ < π/2 Contact angle θ: π/2 < θ < 0

Figure 2.3 Wetting phenomenon.

θ = 0° θ < 90° θ > 90°

Spreading Wetting Dewetting

Figure 2.4 Wetting, spreading, and dewetting for different contact angles.

maximum mutual interaction. Such closeness is possible only when the


adhesives wet the substrate.
Wetting is the ability of liquids to form interfaces with solid surfaces.
To determine the degree of wetting, the contact angle (θ) that is formed
between the liquid and the solid surface is measured. The smaller the con-
tact angle and the smaller the surface tension, the greater the degree of
wetting (Figure 2.3).
For maximum adhesion, the adhesive must completely cover the sub-
strate, i.e., spreading is necessary. The contact angle is a good indicator of
adhesive behavior. This is illustrated in Figure 2.4.

2.10 Equilibrium Contact Angle


In 1805, Thomas Young provided the first good approach for describing
wettability, spreading, and their relationship to the contact angle.
A drop of adhesive on a surface will come to equilibrium under the
action of three forces as shown in Figure 2.5.
Fundamentals of Adhesion 45

γLV

vapor

liquid θ
γSL
γSV
solid

γLV cosθ = γSV – γSL (Young equation)

γLV = liquid-vapor interfacial tension or surface tension


γSV = solid-vapor interfacial tension, not true surface energy
γSL = solid-liquid interfacial tension
θ = contact angle (angle liquid makes with solid surface)

Figure 2.5 Equilibrium contact angle based on balance of forces.

Considering the component of γLV along the X-axis, we can write the
following force balance:

LV cos SL SV

SV SL
Or cos (2.1)
LV

Thus, when θ = 0, the liquid spreads spontaneously on the substrate; in


other words, when cos θ is high (i.e., as it approaches 1), there is sponta-
neous spreading.
From Equation 2.1, it is clear that wetting will be favored when the sur-
face tension of the liquid is low.
Since the tendency of the liquid to wet and spread spontaneously increases
as the contact angle decreases, the contact angle is a useful inverse measure
of wetting or the cosine of the contact angle is a direct measure of wetting.

2.11 Thermodynamic Work of Adhesion


Perhaps the most convenient way of interpreting the wettability of a low
energy solid is the formulation of the work of adhesion, WA, defined by
46 Adhesives for Wood and Lignocellulosic Materials

Dupré and Dupré [22] as the work required to separate a unit area of the
solid–liquid interface.
Consider the wetting of a solid substrate (S) by a liquid (adhesive) “L”. A
solid–liquid interface is formed as a result according to the following equation:

S + L = SL (2.2)

If γS, γL, and γSL are the surface free energies of solid substrate, liquid
(adhesive), and the interphase, then the free energy change of the process
(ΔGA) can be written as

ΔGA = γSL – (γS + γL) (2.3)

The work of adhesion WA = −ΔG can be written as

WA = −ΔGA = (gs + γL) − γSL (2.4)

This is the thermodynamic work of adhesion or the work needed to


separate unit area of the solid–liquid interface.
Assuming γLV = γL and γSV = γs from Equations 2.1 and 2.4, we get

WA = γL (1 + cos θ) (2.5)

This is known as Young–Dupre’s equation. Thus, if the contact angle,


θ, of a well-defined probe liquid against a solid is measured, the work of
adhesion can be determined.
Thus, the thermodynamic work of adhesion (W) is, by definition, the
free energy change per unit area required to separate to infinity two surfaces
initially in contact with a result of creating two new surfaces at the inter-
face between two materials, for example, an adhesive and an adherend.
It is related to the intermolecular forces that operate
at the interface between two materials, for example, an adhesive and an
adherend.
It is related to the intermolecular forces that operate
at the interface between two materials, for example, an adhesive and
an adherend.
It is related to the intermolecular forces that operate
at the interface between two materials, for example,
It is related to the intermolecular forces that operate
at the interface between two materials, for example,
Fundamentals of Adhesion 47

Fowkes [23] proposed that both reversible work of adhesion (W) and
the surface free energy (γ) had additive components and can be parti-
tioned into individual components. Accordingly, several equations were
proposed based on this important approach. This pioneering development
of Professor Frederick M. Fowkes regarding the acid–base theory in adhe-
sion have attracted the attention of several laboratories. A Festschrift in his
honor on the occasion of his 75th birthday was published in 1991.
The approach is described below:

(1) Partitioning of surface free energies into components


The principle of partitioning is based on the assumption that
the surface free energy is determined by various interfacial
interactions. These interactions in turn depend on the basic
properties of the interacting liquid and that of the solid–liquid
interface (SL) [23, 24].

d p h i ab o
s s s s s s s
(2.6)

d p h
where s , s , s , is , ab
s , o
s

are the dispersion, polar, hydrogen (related to hydrogen bonds),


induction, and acid–base components, respectively, while o
refers to all remaining interactions.
(2) Mode of combinations of the individual energy components
According to Fowkes, the dispersion component of the sur-
face free energy is connected with the London interactions.
The remaining van der Waals interactions, i.e., the Keesom and
Debye ones, have been considered by Fowkes as a part of the
induction interactions.
Fowkes investigated mainly two-phase systems containing a
substance (solid or liquid) in which the dispersion interactions
appear only. Considering just such systems, Fowkes determined
the SFE corresponding to the solid–liquid interface as follows:
For two-phase systems comprising of a solid and liquid, in
which only dispersion interactions occur, namely, between
d d
s, 1,
Fowkes employed geometric mean as the mode of com-
bination of these energy components to give the following
equation:

d d d d 0.5 (2.7)
SL s L 2( s L )
48 Adhesives for Wood and Lignocellulosic Materials

Fowkes [25] modified Equation 2.7 by changing from geometric mean to


arithmetic mean to arrive at the following equation:
d d d d (2.8)
SL s L ( s L )

Owens and Wendt [26] significantly changed the Fowkes idea while
assuming that the sum of all the components occurring on the right-hand
side of Equation 2.11, namely
p h i ab o
s , s , s , s , s ,
p
except that γd can be considered as associated with the polar interaction s ,
Consequently, the following equation was obtained:

d d p p
SL s L 2 s L 2 s L (2.9)

Wu [27, 28] accepted the idea by Owens and Wendt to divide the SFE into
two parts, but used the harmonic means of the interfacial interactions instead
of the geometric means in Equation 2.9 and derived the following equation:

d d p p
s L s L
SL s L 4 d d p p
(2.10)
s L s L

van Oss, Chaudhury, and Good proposed the latest concept of partition
in which surface energy is partitioned into two components [29, 30]:

(1) Long range interactions London, Keesom, and Debye


called the Lifshitz–van der Waals component (γLW)
(2) The short-range interactions (acid–base), called the acid–
base component (γAB) = 2(γ+ γ–)0.5 where γ+ and γ– mean
the acidic and basic constituents, respectively, which are
associated with the acid–base interactions.

Combining van Oss Chaudhury’s concept with Young’s equation


(Equation 2.5), we obtain the Young–Fowkes–van Oss–Good acid–base
equation referred to as the acid–base approach:

0.5 0.5 0.5


LH LH
(1 cos ) L 2 S L S L S L (2.11)
Fundamentals of Adhesion 49

Table 2.3 Physical properties and surface free energy components of test liquids
used at 20°C [31].
Surface free energy (mJm–2)
(LW–AB approach)
Density Viscosity LW + AB
Liquid (kg/m3) (mPas) 1v 1v 1v 1v γ1v
Water 1000 1.00 21.8 25.5 25.5 51.0 72.8
Formamide 799 1.02 39 2.28 39.6 19 58
Ethylene glycol 1109 19.9 29 1.92 47 19 48
Diethylene glycol 1130 26.8 44.7 – – – –
Diodomethane 3325 2.8 50.8 0 0 0 50.8
1-Bromo- 1483 – 44.4 0 0 0 44.4
naphthalene

This equation contains three unknowns and, therefore, we need contact


angle data of three liquids. Table 2.3 contains the physical properties and
surface free energy components of test liquids normally used. One of them
must be non-polar and the other two should be polar. The Lifshitz–van der
Waals component of the surface free energy is obtained from the contact
angle of the apolar test liquid, e.g., diiodomethane.
The acid–base interaction and its relevance to adhesives and adhesive
bonding have been reviewed in detail by Chehimi et al. [32].

2.12 Spreading Coefficient


For spreading, another parameter, the spreading coefficient γSV − γSL − γLV
appears to be important in classifying liquids that have a tendency to form
good films on a given substrate. In general, the larger and more positive
value of the spreading coefficient (S), the more energy is gained by interca-
lating a liquid film between a solid and air. Thus,

S > 0, spontaneous spreading


S < 0, not spontaneous spreading

Though the condition S > 0 is necessary for a liquid to spread sponta-


neously on a solid, it is insufficient to describe the final state of the film.
50 Adhesives for Wood and Lignocellulosic Materials

1.0
Dimethyl sulfoxide y = –0.0108x + 1.3778
R2 = 0.8871
0.9

0.8
Cos θ

Ethylene glycol
0.7
Glycerol Water

0.6

0.5
30 35 40 45 50 55 60 65 70 75 80

γc = 34.98 mJ/m2 Surface tension of liquid – mJ/m2

Figure 2.6 Zisman’s plot (https://www.researchgate.net/publication/279532584_Analysis_of_


the_Results_of_Surface_Free_Energy_Measurement_of_Ti6Al4V_by_Different_Methods/
figures?lo=1).

2.13 Zisman’s Rectilinear Relationship—Zisman’s


Plots and Critical Surface Tension of a Solid
Fox and Zisman [33] introduced the well-known concept of critical surface
tension, γc. In the method adopted by Zisman, the contact angles θ for a
series of organic homologous liquids were measured on a solid. A plot of
cos θ vs. surface tension of the liquids gave a straight line (Figure 2.6). The
point of intersection of the straight line and the line through cos θ = 1 is the
“critical surface tension” γc of the solid. All liquids whose surface tension is
less than the critical surface tension will wet the substrate.

2.14 Effect of Surface Roughness on Contact Angle


The above equation is true for smooth, contamination-free surface. However,
the real solid surface is not smooth and the roughness of the surface has a pro-
found effect on the wetting and adhesion. The innumerable small hills, valleys,
and crevices on the solid surface entrap and occlude air or vapor within them.
Even if θ = 0, it is not possible under real conditions to ensure that an intimate
contact between the adhesive and adherend is established. Surface roughness
plays therefore an important role in the wettability of a solid surface.
The impact of roughness on the contact angle is given by the Wenzel
equation (Equation 2.12)

Cos θW = r Cos θY (2.12)


Fundamentals of Adhesion 51

The Wenzel equation (Equation 2.12) relates the contact angle θw of


a liquid measured on a rough surface having a roughness ratio, r, with
the contact angle of the same liquid measured on a smooth surface, θY.
The roughness ratio is the ratio of the true surface area of a rough sur-
face to the surface area of the smooth surface. This ratio r will always be
larger than one. Wenzel’s relation also shows that surface roughness will
decrease the contact angle for a water droplet on a hydrophilic surface or
increase the contact angle for a water droplet on a hydrophobic surface.
The advancing and receding contact angles can throw light on the mag-
nitude of roughness of the wood surface. The difference between advancing
and receding contact angles is the contact angle hysteresis. The magnitude of
contact angle hysteresis is dependent on roughness, topography, morphol-
ogy, and chemical homogeneity of the solid surface [31]. Good [34] sug-
gested that the advancing contact angle represents hydrophobic areas on the
surface, while the receding contact angle characterizes hydrophilic areas.

2.15 Weak Boundary Layer Theory


The weak boundary layer theory explains the loss of adhesion as a failure in
an intermediate molecular layer between adhesive and adherent [35]. This
molecular layer consists of low-molecular-weight impurities of various ori-
gins including water. This theory has never been verified for wood, but it is
known that low-molecular-weight extractives can easily migrate to the sur-
face and might reduce adhesion due to weak boundary layer (see Chapter 9).
Bonding of wood is described as a chain of several links, comprising
wood, wood surface and its boundary layer, interphase of wood and adhesive
and interface between wood and adhesive, and the adhesive bond line itself.
One useful method for understanding the abovementioned links that
control the adhesive strength as well as weakness is the chain link analogy
proposed by Marra [5]. Different areas of the substrate and adhesive are
likened to a series of chain links, with the weakest link being the site of
fracture. This is depicted in Figure 2.7. These links are as follows:
Link 1, adhesive film; links 2 and 3, intra-adhesive boundary layers; links 4
and 5, adhesive–adherend interface (in this region, the weak boundary layer
exist); links 6 and 7, adherend subsurface; links 8 and 9, adherend proper.
Another cause for weak link is due to the stresses caused at the bond
line due to swelling and shrinkage due to moisture changes. If the adhesive
bond has to be durable, it has to adapt itself to the dimensional changes
and the consequent strain due to swelling and shrinkage accompanying the
changes in moisture conditions. Two distinct classes of wood adhesives have
52 Adhesives for Wood and Lignocellulosic Materials

Figure 2.7 Different links in adhesive bonding.

different ways to distribute this strain: (a) rigid in situ polymerized adhe-
sives relieves this stress in many cases by distributing the strain through the
wood interphase region. (b) The other class, the more flexible pre-polymerized
adhesives, can distribute the strain through the adhesive interphase. Failure to
adequately perform either of these strain distribution processes can lead to
high strains and subsequent failure zones. As wood dries, it shrinks back
to near its original dimensions. These failure zones can expand and become
more visible as delamination areas [36].
Mechanically weak wood surfaces can be another source of weak
boundary layer. The causes of this are many [5, 35]. One cause is physical
crushing of the surface, especially by abrasive planing or by too high of
a bonding pressure. This occurs when more pressure is applied than the
thin-walled earlywood cells can withstand. The strength of these cells is
reduced due to deformation and fracture of the cell walls. A second cause
is sanding dust or other dust accumulation on surfaces. A third cause is
tearing of the surface that occurs during planing and sanding. A fourth
cause, associated with high-density wood species, is cells becoming sepa-
rated from one another due to the force of planning [36].

2.16 Measurement of the Wetting Parameters


for Wood Substrate
Unlike other substrates, wood exhibits complex anatomical features that
make it heterogeneous and porous. It is basically hygroscopic. It contains
extractives.
Fundamentals of Adhesion 53

As a result of these unique physicochemical characteristics, wetting


measurements on wood is difficult. Direct measurement of contact angle
of adhesives on wood surface may not be reproducible and hence not sat-
isfactory. For example, a drop of water on a wood surface will, in most
cases, quickly change its size and shape over time, which will thus lead to a
change in the contact angle.
New methods for determining the wetting parameters of wood are nec-
essary. A detailed and critical review of the various methods to determine
the wetting parameters have been reported by Magnus Wålinder. The read-
ers may refer to the same [37]. A brief summary from the above review is
given below.
“One way to address some of the difficulties in wood wetting measure-
ments may be to apply the Wilhelmy (1863) method [38]. In contrast to
direct measurement of contact angles, as in the drop method, the Wilhelmy
method involves determining the force acting on a specimen when it is
immersed in and withdrawn from a liquid. An apparent contact angle can
then be estimated from an analysis of the recorded force”.
“Other promising techniques for estimating the surface energetics of
wood may be the Axisymmetric Drop Shape Analysis-contact diameter
(ADSA-CD) technique. Contact angle measurements determined as con-
stant wetting rate angle values (cwra) and also a capillary rise technique
(column wicking) applied to wood”.
“Inverse gas chromatography (IGC) is a useful technique for determin-
ing surface energetics of particle surfaces. By using appropriate gas probes,
IGC can provide information on the surface thermodynamic character-
istics of particles including surface free energy, acid-base interactions,
enthalpy, and entropy. IGC has been applied to many materials such as
polymers, wood pulp and wood particles”.
Some spectroscopic methods, namely, X-ray photoelectron spectros-
copy (XPS) and FT Raman, have also been recommended by the author.
Both IGC and wicking methods rely on wood powder, which will give
different results compared to measurements on solid wood.

2.16.1 Some Results on Surface Energy of Wood


As discussed in Chapter 1, wood is a hierarchical cellular material and is
therefore anisotropic in nature. Further wood surfaces are topographically
different in radial, tangential, and transverse sections [31]. The wood sur-
face consists of earlywood and latewood. The water contact angle of early-
wood is often different from that of latewood. At the microscale, the wood
surface consists primarily of lumen surfaces and cross-sectional walls.
54 Adhesives for Wood and Lignocellulosic Materials

Thus, wetting of wood by the adhesives is therefore complex. Freeman was


the first to report on the wettability of wood [39].
Gray carried out an extensive investigation on the wettability of 20
species of wood [7]. Gray was first to determine the surface free energy
of wood. Sessile drop method was used to determine the contact angle,
and the critical surface tension (γc) was obtained by the Zisman method.
Species used by Gray were Western Hemlock (Tsuga heterophylla),
Douglas-fir (Pseudotsuga taxifolia), Afrormosia (Afrormosia elata),
Parana pine (Araucaria angustifolia), European redwood (Pinus sylves-
tris), English oak (Quercus robur), and Beech (Fagus sylvatica). The val-
ues of critical surface tensions ranged from 34.5 to 81.0 mJ/m2. One of
the important findings was that freshly sanded surfaces were approxi-
mately 20 mJ/m2 higher in surface energy than un-sanded, aged surfaces.
It was also shown that surface contamination occurs rapidly on freshly
cleaned surfaces.
Herczeg reported on the wettability of Douglas-fir wood [40]. The crit-
ical surface tensions, γc, were found to be between 44 and 50 dynes/cm for
summerwood and springwood, respectively. The surface-free energy and
the maximum work of adhesion were also reported. It was also reported
that aging increased contact angle, showing that wood wettability was
reduced. Chen reported that removal of extractives from some tropical
woods improved wettability [41]. Hse measured the wettability of south-
ern pine veneers by measuring the contact angles formed with 36 phenol-
formaldehyde resins [42]. The contact angle of resins on earlywood was
less than that on latewood, apparently because earlywood surfaces were
rougher. Also, the contact angle was positively correlated with the glue bond
quality as tested by wet shear strength, percent of wood failure, and per-
cent of delamination. Nguyen and Johns found that the surface free energy
of Douglas-fir and redwood decreased with aging time [43]. Extractives
and aging had significant influence on the surface energy. The surface free
energy of Douglas-fir was 48.0 mJ/m2, and after extraction, it increased
to 58.9 mJ/m2. These results emphasize the influence of extractives on the
wood surface energy.
Kalnins et al. [44] employed a video-type technique to measure the
dynamic contact angle of distilled water as test liquid on wood with mea-
surements conducted at the elapsed time of 3 to 5 s.
Gardner et al. found dynamic contact angle measurements to be useful
for determining the effect of wood processing and environmental condi-
tioning on surface energetic [45].
Kajita and Skaar evaluated the wettabilities of the surfaces of some
American softwoods species (using cosine 0 as the index of wettability) [46].
Fundamentals of Adhesion 55

Also, the earlywood wet more easily than did the latewood (earlywood has
a greater roughness factor and a greater porosity). The wetting angles var-
ied from 68° (eastern red cedar) to 14° (Alaska cedar). The greater wettabil-
ity of sapwood compared with heartwood was attributed to the extractive
content of the heartwood.
Cosine of the advancing contact angle was employed as the measure of
wettability. The wettability, pH, and specific gravity were closely related to
glue-bond quality of resorcinol-phenolic and urea formaldehyde-bonded
adhesive joints [47].
Shi and Gardner developed a wetting model to describe to quantify the
adhesive penetration and spreading during the adhesive wetting process
[8]. Sapwood and heartwood of southern pine and Douglas-fir were stud-
ied. Two resin systems, polymeric diphenylmethane diisocyanate (PMDI)
and phenol formaldehyde (PF), were evaluated. It was learned from this
study that the wetting model could accurately describe the dynamic adhe-
sive wetting process on wood surfaces. Shen et al. presented a systematic
study of surface free energy and acid–base properties of pine (P. sylvestris
L.); for evaluation of the data, the Lifshitz–van der Waals/acid–base (LW–
AB) approach was applied [48].
Nussbaum observed a decrease of wettability as a function of time
on wood surface after sawing due to the migration of wood extractives
to the surface [49]. Gindl et al. compared the applicability of different
approaches to determine the surface free energy of wood and found the
LW–AB approach to be the most effective among the generally accepted
models [50].
de Meijer et al. [51] employed contact angle measurements to calculate
the Lifshitz−van der Waals, acid−base, and total surface free energies of
wood species spruce (Picea abies) and meranti (Shorea spp.). These species
were characterized by low surface energy with a dominant Lifshitz−van
der Waals component. The authors report that thermodynamic equilib-
rium conditions as assumed by Young’s equation are generally not fulfilled
with wood surfaces because of chemical heterogeneity, surface roughness,
and the adsorption of the test solvent.
An exhaustive review of wettability of wood has been published by
Piao et al. [31]. The review also includes calculation of surface tension of
wood, Zisman’s critical surface tension, Owens–Wendt’s geometric mean,
and Wu’s harmonic mean; Young–Fowkes–van Oss–Good acid–base
approaches and the inverse gas chromatography method have been dis-
cussed in detail. The review also deals with variables that affect the wet-
tability and surface energy of wood. Detailed overview of literature data
obtained on wood surfaces was presented by de Meijer et al. [51].
56 Adhesives for Wood and Lignocellulosic Materials

2.17 Covalent Bond Formation


Covalent bonds between an adhesive and wood are believed to improve
bond durability. Although covalent bonds—chemical bonds between the
adhesive and wood—seem plausible with some adhesives, they have never
been unambiguously detected in an adhesive bondline and no evidence
exists that they contribute to the strength of adhesive bonds To determine
whether an adhesive forms covalent bonds to wood, it must have the fol-
lowing characteristics: (1) be highly reactive with wood polymer hydrox-
yls, (2) be capable of permeating the cell wall, (3) exhibit strong wettability
to wood, and (4) be amenable to study using a monofunctional model
compound. Ideally, the reaction between the monofunctional model com-
pound and wood will produce distinct chemical shift differences between
unreacted and reacted wood polymers in solution-state nuclear magnetic
resonance (NMR) spectroscopy.
While there is no doubt that adhesive-to-wood covalent bonds can
form under specific experimental conditions, the conditions employed in
studies reporting such bond formation generally have not corresponded
to the conditions commonly used in the bonding of wood, in particular
hot-pressing of wood panels [52]. For instance, covalent bonds between
wood and a synthetic adhesive can form at temperatures higher than
120°C maintained for 2 h [52]. But extensive covalent bonding appears
unlikely in the core of a particleboard, which is able to reach only 115°
to 120°C for no more than 1 to 1.5 min when pressed at 200°C for 3 min.
Allan and Neogi found in the case of phenol-formaldehyde bonding of
wood at 120°C for 2 h that only one adhesive-to-substrate covalent bond
was formed for approximately every 1200 cross-links within the resin itself
[53]. This was also the case for isocyanate binders, for which the original
misconception of a predominance of covalent bonding between adhesive
and substrate that has been used to explain the high strength of the panels
obtained in that manner was disproved [54, 55].

References
1. Frihart, C.R., Wood adhesion and adhesives, in: Handbook of Wood Chemistry
and Wood Composites, 2nd edn., R. Rowell (Ed.), pp. 255–319, CRC Press,
Boca Raton, Florida, 2013.
2. Israelachvili, J.N., Intermolecular and Surface Forces, 2nd edn., Academic
Press, London, 1991.
Fundamentals of Adhesion 57

3. van Oss, C.J., Interfacial Forces in Aqueous Media, Marcel Dekker, New York,
1994.
4. Arif Butt, M., Arshad Chautai, A., Ahmad, J.A.R., Theory of adhesion and its
practical implications. J. Faculty Eng. Technol., 21–45, 2007.
5. Marra, G.G., Technology of Wood Bonding: Principles in Practice, Van
Nostrand, New York, 1992.
6. Browne, F.L. and Brouse, D., Nature of adhesion between glue and wood. Ind.
Eng. Chem., 21, 80–84, 1929.
7. Gray, V.R., The wettability of wood. For. Prod. J., 12, 452–461, 1962.
8. Shi, S.Q. and Gardner, D.J., Dynamic adhesive wettability of wood. Wood
Fiber Sci., 33, 56–68, 2001.
9. Marcinko, J.J., Phanopoulos, C., Beachey, P., Wood Adhesives 2000, Proceedings
No. 7252, pp. 111–121, Forest Products Society, Madison, Wisconsin, 2001.
10. Christiansen, A.W., How overdrying wood reduces its bonding to phenol–
formaldehyde adhesives: A critical review of the literature. Part I. Physical
responses. Wood Fiber Sci., 22, 4, 441–459, 1990.
11. Christiansen, A.W., How overdrying wood reduces its bonding to phenol–
formaldehyde adhesives: A critical review of the literature. Part II. Chemical
reactions. Wood Fiber Sci., 23, 1, 69–84, 1991.
12. Stehr, M., Adhesion to machined and laser ablated wood surfaces, Dissertation,
KTH Stockholm, Sweden, 1999.
13. McBain, J.W. and Hopkins, D.G., On adhesives and adhesive action. J. Phys.
Chem., 29, 2, 188–204, 1925.
14. Deryaguin, B.V., Problems of Adhesion. Vestnik Akademie, 8, 70, 1955.
15. Deryaguin, B.V., Krotova, N.A., Karassev, V.V., Kirillova, Y.M., Aleinikova,
I.N., Proceedings of the Second International Congress on Surface Activity—
III, Butterworths, London, 1957.
16. Deryaguin, B. and Smiliga, V.P., Adhesion Fundamentals and Practice,
MacLaren and Sons, London, 1969.
17. Voyutskii, S.S., Adhesives Age, 5, 4, 30, 1962; Voyutskii, S.S., Vakula, V.L.,
The role of diffusion phenomena in polymer-to-polymer adhesion. J. Appl.
Polym. Sci., 7, 2, 475–491, 1963.
18. Voyutskii, S.S., Autohesion and Adhesion of High Polymers, Wiley Interscience,
New York, 1963.
19. Voyutskii, S.S., Markin, S., Yu, I., Gorchakova, v.M., Gul, V.E., Adhesion of
polymers to metals. Adhesives Age, 8, 24, 1965.
20. Pizzi, A., A brief, non-mathematical review of adhesion theories as regards
their applicability to wood. Holzforsch Holzververt, 44, 1, 6–11, 1992.
21. Gardner, D.J., Adhesion mechanisms of durable wood adhesive bonds.
BLIO009-Stokke September 13, 21, 254–266, 2005.
22. Dupré, A. and Dupré, P., Théorie mécanique de la chaleur, Gauthier-Villars,
Paris, 1869.
23. Fowkes, F.M., Attractive forces at interfaces. Ind. Eng. Chem., 56, 12, 40–52, 1964.
58 Adhesives for Wood and Lignocellulosic Materials

24. Fowkes, F.M., Role of acid–base interfacial bonding in adhesion. J. Adhes.


Sci. Technol., 1, 7–27, 1987.
25. Fowkes, F.M., Calculation of work of adhesion by pair potential summation.
J. Colloid Interface Sci., 28, 493–505, 1968.
26. Owens, D.K. and Wendt, R.C., Estimation of the surface free energy of
polymers. J. Appl. Polym. Sci., 13, 1741–1747, 1969.
27. Wu, S., Calculation of interfacial tension in polymer system. J. Polym. Sci. C,
34, 19–30, 1971.
28. Wu, S., Polar and nonpolar interactions in adhesion. J. Adhes., 5, 39–55,
1973.
29. van Oss, C.J., Good, R.J., Chaudhury, M.K., The role of van der Waals forces
and hydrogen bonds in “hydrophobic interactions” between biopolymers
and low energy surfaces. J. Colloid Interface Sci., 111, 378–390, 1986.
30. van Oss, C.J., Chaudhury, M.K., Good, R.J., Interfacial Lifshitz–van der
Waals and polar interactions in macroscopic systems. Chem. Rev., 88, 927–
940, 1988.
31. Piao, C., Winandy, J.E., Shupe, T.F., From hydrophilicity to hydrophobicity:
A critical review: Part I. Wettability and surface behaviour. Wood Fiber Sci.,
42, 4, 490–510, 2010.
32. Chehimi, M.M., Azioune, A., Cabet-Deliry, E., Acid–base interactions:
Relevance to adhesion and adhesive bonding, in: Handbook of Adhesive
Technology, 2nd edn., A. Pizzi and K.L. Mittal (Eds.), Marcel Dekker,
New York, 2003.
33. Fox, H.W. and Zisman, W.A., The spreading of liquids on low energy sur-
faces. J. Colloid Sci., 5, 6, 499–595, 1950.
34. Good, R.J., Contact angles and the surface free energy of solids. Page 1, in:
Surface and Colloid Science, vol. VII, R.J. Good and R.R. Stromberg (Eds.),
Plenum Press, New York, 1979.
35. Stehr, M. and Johansson, I., Weak boundary layers on wood surfaces.
J. Adhes. Sci. Technol., 14, 1211–1224, 2000.
36. Frihart, C.R. and Hunt, C.G., Adhesives with wood materials: Bond forma-
tion and performance, in: Wood Handbook: Wood as an Engineering Material,
Centennial edn., General Technical Report FPL–GTR–190. pp. 10.1–10.24,
U.S. Dept. of Agriculture, Forest Service, Forest Products Laboratory, USDA,
2010.
37. Walinder, M., Wetting Phenomena on Wood: Factors Influencing Measurements
of Wood Wettability, Dissertation, KTH Royal Institute of Technology,
Stockholm, 2000.
38. Wilhelmy, L., Ueber die Abhängigkeit der Capillaritäts-Constanten des
Alkohols von Substanz und Gestalt des benetzten festen Körpers. Ann.
Physique Chimie, Band CXIX, 5, 6, 12, 1863.
39. Freeman, H., Properties of wood and adhesion. For. Prod. J., 9, 451–458,
1959.
40. Herczeg, A., Wettability of wood. For. Prod. J., 15, 499–505, 1965.
Fundamentals of Adhesion 59

41. Chen, C., Effect of extractive removal on adhesion and wettability of some
tropical woods. For. Prod. J., 20, 1, 36–40, 1970.
42. Hse, C.-Y., Wettability of southern pine veneer by phenol–formaldehyde
wood adhesives. For. Prod. J., 22, 1, 51–56, 1972.
43. Nguyen, T. and Johns, W.E., The effects of aging and extraction on the surface
free energy of Douglas fir and redwood. Wood Sci. Technol., 13, 1, 29–40,
1979.
44. Kalnins, M.A., Katzenberger, C., Schmieding, S.A., Brooks, J.K., Contact
angle measurement on wood using videotape technique. J. Colloid Interface
Sci., 125, 344–346, 1988.
45. Gardner, D.J., Generalla, N.C., Gunnels, D.W., Wolcott, M.P., Dynamic
wettability of wood. Langmuir, 7, 2498–2502, 1991.
46. Kajita, H. and Skaar, C., Wettability of the surfaces of some American soft-
woods species. Mokuzai Gakk., 38, 516–521, 1992.
47. Mantanis, G.I. and Young, R.A., Wetting of wood, Wood Sci. Technol., 31,
339–353, 1997.
48. Shen, Q., Nylund, J., Rosenholm, J.B., Estimation of the surface energy and
acid–base properties of wood by means of wetting method. Holzforschung,
52, 521–529, 1998.
49. Nussbaum, R.M., Natural surface inactivation of Scots pine and Norway
spruce evaluated by contact angle measurements. Holz Roh-Werkst., 57,
419–424, 1999.
50. Gindl, M., Sinn, G., Gindl, W., Reiterer, A., Tschegg, S., A comparison of
different methods to calculate the surface free energy of wood using contact
angle measurements. Colloids Surf. A Physicochem. Eng. Asp., 181, 279–287,
2001.
51. de Meijer, M., Haemers, S., Cobben, W., Militz, H., Surface energy determi-
nations of wood: Comparison of methods and wood species. Langmuir, 16,
9352–9359, 2000.
52. Hubbe, M.A., Pizzi, A., Zhang, H., Halis, R., Critical links governing per-
formance of self-binding and natural binders for hot-pressed reconstituted
lignocellulosic board products: A review. Bioresources, 13, 1, 1–67, 2018.
53. Allan, G.C. and Neogi, A.N., Fiber surface modification, Part VIII: The
mechanism of adhesion of phenol–formaldehyde resins to cellulosic and lig-
nocellulosic substrates. J. Adhes., 3, 1, 13–18, 1971.
54. Pizzi, A. and Owens, N.A., Interface covalent bonding vs. wood-induced
catalytic autocondensation of diisocyanate wood adhesives. Holzforschung,
49, 269–272, 1995.
55. Wandler, S.L. and Frazier, C.E., The effects of cure temperature and time
on the isocyanate–wood adhesive bondline by 15N CP/MAS NMR. Int. J.
Adhes. Adhes., 16, 3, 179–186, 1996.

You might also like