You are on page 1of 29

14

Rheology and Viscoelasticity of Adhesives

14.1 Rheology of Adhesives


Rheology is the science that deals with the deformation and flow of flu-
ids. This is a very important subject for adhesives and adhesion since, for
instance, wood adhesion involves flow, spreading, transfer, and penetra-
tion of adhesives into the wood adherends. Furthermore, the consistency
of the adhesives is an important property to be controlled during the man-
ufacture of adhesives. In addition, the flow properties directly or indirectly
control the reactivity of the adhesives during the curing process. Wood
adhesives consist not only of pure oligomers but also of additives such as
fillers, extenders, surface active agents, etc. These formulated adhesives
exhibit non-Newtonian flow characteristics. Hence, rheological methods
are used to test the flow properties and the reactivity of adhesives during
their routine production as well as during the development of new adhe-
sive systems. The physical property of importance for the study of rheology
is the viscosity. Hence, the theory of viscosity has to be understood first.

14.2 Viscosity—Theory
Viscosity is a measure of “resistance to flow” [1]. Figure 14.1 illustrates this.
Two plates with the area A contain a thin layer of fluid. The bottom
plate does not move, while the top plate is forced to move because of the
force F [N]. The shear stress is defined as

F
[N/ m 2 ]
A

R. N. Kumar and A Pizzi. Adhesives for Wood and Lignocellulosic Materials, (317–345) © 2019
Scrivener Publishing LLC

317
318 Adhesives for Wood and Lignocellulosic Materials

F
A
dux

d
y

Figure 14.1 Viscous flow between two parallel plates.

The upper plate is moving at a velocity dux [m/s]. The shear rate is
defined as the velocity gradient

dux 1
[s ]
dy

If the shear stress and the shear rate are known, the dynamic viscosity
can be calculated as

[P a. S]

1 centipoise [cP] = 0.001 pascal second [Pa·s] or 1 cP = 1 mPa·s (one


millipascal-second). or 1 pascal second is equal to 10 poise.
The dynamic viscosity has the dimensions: ML−1 T−1
If M = 1 g, L= 1 cm, and T = 1 s, then ML−1 T−1 = 1 g cm−1 s−1 = 1 Poise
(g cm−1 s−1)
1 Poise = 100 centipoises.

14.3 Capillary Viscometry


The viscosity of even very dilute polymer solutions is considerably higher
than that of solvents. The increase in viscosity depends on several factors
such as temperature, solvent–polymer interaction, polymer concentration,
and the size of polymer molecules. The dependence of viscosity on poly-
mer size permits one to estimate an average molecular weight from the solu-
tion viscosity. The average molecular weight that is measured is called the
viscosity average molecular weight M v to distinguish it from M n and  M W.
Rheology and Viscoelasticity of Adhesives 319

The relationship between the increase in viscosity and the molecular weight


is given by the Mark–Howink–Sakurada equation:

[η] = K Ma

K and a are empirical constants for a given polymer–solvent and tem-


perature combinations and a increases in the range 0.5 with the degree
of expansion of the molecular coils from their unperturbed dimensions.
Under “theta” conditions, the molecular coils have unperturbed dimen-
sions a = 0.5. The Mark–Howink constants K and a are evaluated from [η]
and molecular weight (M) of calibration samples with narrow molecular
weight distribution. Having determined the constants K and a, the molec-
ular weight of polymers can be determined from the viscosity data. The
viscosity method of molecular weight determination requires very little
investment in apparatus and can be carried out quite rapidly by flow time
measurements with a simple glass viscometer (capillary viscometer) set up
in a constant temperature bath (Figure 14.2) [2].
Capillary viscometers basically are used for Newtonian, incompress-
ible, and wall-adherent liquids. The flow inside the capillary is assumed
to have the following idealized conditions: laminar, incompressible, and
stationary. Moreover, the flow influence at the entry and exit of the cap-
illary is neglected and the viscosity of fluids is assumed to be pressure
independent.

Upper
mark

Lower
mark

Figure 14.2 Ostwald’s viscometer.


320 Adhesives for Wood and Lignocellulosic Materials

Capillary viscometers are working based on the Hagen–Poiseuille Law


as their physical basis:

v πR 4ΔP
t 8L

in which V, t, R, ΔP, L, and η are the volume, the time, the radius of the
capillary, the pressure drop, the length of the capillary, and the dynamic
viscosity of fluids, respectively. The volume flow measurement through the
capillary at a given differential pressure is the fundamental measurement
criteria for capillary viscometers. In other words, the viscosity is deter-
mined by measuring the time required for a defined volume of liquid to
flow through a capillary tube due to the hydrostatic pressure of the liquid
column itself.
Two marks (the upper mark and lower mark) before and after a ball-
shaped extension are reference points that enable measurement of the time
of flow of the liquid through the capillary.
ΔP in the above equation = hρg, where h is the average height of the
height of the liquid during measurement, ρ is the density of the polymer
solution, and g is the acceleration due to gravity. Since all the quantities
except η are known, the viscosity η can be calculated.

14.4 Rotational Viscometers


Rotational viscometers function on the principle that the force required
to turn an object in a fluid is a measure of the viscosity of that fluid [3].
A rotating spindle (cylinder or disk) is submerged in the sample fluid.
The liquid offers resistance to the rotational motion of the spindle. The
force (torque) required to overcome this resistance against the rotation
of the spindle at a constant speed is measured. Thus, the rotational vis-
cometers work by measuring the torque on a vertical shaft that rotates
a spindle. The spindle is in the test sample and its rotation is impeded
in proportion to the viscosity of the sample. With precise calibration,
these instruments measure viscosity. Rotational viscometers fall into
two main types:

1. Spring type
2. Servo system
Rheology and Viscoelasticity of Adhesives 321

14.4.1 Spring Type


This system uses a spring to capture the torque exerted by the liquid on the
spindle. These systems use sets of calibrated springs.
A spring viscometer system is based on a pivot and spring concept. A
pivot/spring assembly rotates on the shaft. The spindle is attached to this
assembly. As the spindle rotates, the drag of the fluid against the spindle
causes the spring to deflect. This deflection is proportional to the torque
caused by the viscosity of the sample. This relationship to viscosity is calcu-
lated automatically [4]. The spring system has an advantage in that it offers
measurements of very high accuracy [5].

14.4.2 Servo Systems


The servo system is the second system for the determination of the torque.
A precision servo motor is used to drive the shaft. The spindle is attached
directly to the shaft instead of a pivot, as in the case of the spring system.
The current needed to drive the shaft rotating at the test speed is first deter-
mined. When the spindle is submerged in the liquid and is allowed to rotate
therein, the drag of the liquid exerted on the spindle makes the system use
more current to keep the spindle running at the set speed. This current is
an indirect measure of the torque and it is proportional to the viscosity.
Viscosity is taken as to be directly proportional to the total amount of cur-
rent produced. The servo system is reported to be more advantageous than
the spring system.

14.5 Cone-and-Plate Viscometer


A cone-and-plate viscometer is an instrument for determining the abso-
lute viscosity of liquids in small sample volumes. Its cone-and-plate
geometry is such that complete rheological properties can be precisely
determined [5].

Principle of operation
A cone-and-plate viscometer is a precise torque meter that is driven at dis-
crete rotational speeds. The sample of the liquid is introduced in the space
between the cone and the plate. The torque measuring system consists of
a calibrated spring (e.g., beryllium–copper) that connects the drive mech-
anism to a rotating cone and senses the resistance to rotation caused by
liquid between the cone and a stationary flat plate. The resistance to the
322 Adhesives for Wood and Lignocellulosic Materials

rotation of the cone produces a torque that is proportional to the shear


stress in the fluid. With a specific diameter, angle of cone, torque, and the
angular velocity the shear stresses, shear rates and viscosity can be deter-
mined by using appropriate equations [4, 6].
The major advantage of this measuring system is the constant shear rate
throughout the sample. Also, the sample size is small that results in less
heat generated at high shear rates. Another advantage is the ease of clean-
ing the system.

14.6 Parallel Plate Viscometer


The parallel plate viscometer consists of two identical, coaxial and parallel
circular plates. The shear rate and shear stress and hence the viscosity are
calculated from the radius of a circle with the center of the plate, radius of
the plate, the gap between two plates, torque, and the angular velocity by
employing the appropriate equations [5, 7].
Parallel plate systems are referred to by the diameter of the upper plate.
For instance, a PP40 is a 40-mm-diameter plate. The lower plate is either
larger than or the same size as the upper plate.
The major application of this measuring system is to study the curing
behavior of different wood adhesives. Some work on the curing of phenol–
formaldehyde (PF) resins has been reported [8].

14.7 Concentric Cylinder Viscometer


The concentric cylinder viscometer consists of two concentric cylinders
with different radii. The rheological behavior of samples filled in the gap
between the two cylinders is determined. If the gap is small enough com-
pared to the diameters of the cylinders, shear rate is approximately con-
stant across the sample in the gap. The shear rate, shear stress, and viscosity
can be determined from the radius of the outer cylinder, the radius of the
inner cylinder, the angular velocity, effective immersed length of the sam-
ple, and the torque from the appropriate equations [5].

14.8 Ford Cup Viscosity


The Ford viscosity cup (Figure 14.3) is a simple device for the determina-
tion of time of flow of a known volume of liquid through an orifice located
Rheology and Viscoelasticity of Adhesives 323

Figure 14.3 Ford cup.

at the bottom. This device is employed to check continuously the time of


flow of the resin as the reaction proceeds in the reaction vessel. This time of
flow indicates the extent of polymerization of the resin in the kettle. Under
ideal conditions, this rate of flow would be proportional to the kinematic
viscosity (expressed in stokes and centistokes) that is dependent on the
specific gravity of the draining liquid. However, the conditions in a simple
flow cup are seldom ideal for making true measurements of viscosity.
The Ford cup is used mainly to determine the time of flow of a resin
during manufacture and to determine the completion of the reaction as
well as maintain the consistency of the quality of the resin. The tempera-
ture of the cup and the liquid is maintained, as ambient temperature makes
a significant difference to viscosity and thus flow rate.
The original Ford cup was based on imperial (US) measurement of the
aperture. Many other types of flow cups are being used, depending on the
industry or region, such as Din Cup 4  mm; standard DIN 53211 (can-
celled) ISO Cup 2, 3, 4, 5, 6, and 8 mm; standard ISO 2431 AFNOR Cup
2, 5, 4, 6, and 8 mm; and standard NF T30-014 ASTM Cup 1, 2, 3, 4, and 5
standard ASTM D1200.

14.9 Gardner–Holt Tubes


Gardner bubble viscometer tubes are used to quickly determine kinematic
viscosity of known liquids such as resins and varnishes. The bubble vis-
cometer tubes are also described as Gardner–Holt tubes. Repeated read-
ings may be taken easily once the temperature has been controlled. The
principle of operation is based on the time required for an air bubble to
rise in the adhesive from one end to the other end of each tube, which is
directly proportional to the viscosity of the liquid—the faster the bubble
324 Adhesives for Wood and Lignocellulosic Materials

rises, the lower the viscosity. A complete set of Gardner–Holt tubes covers
in general ranges from 0.05 to 1000 stokes.

14.10 Newtonian and Non-Newtonian Fluids


Two types of fluids exist: Newtonian fluids and non-Newtonian fluids.
They are governed by the power law

n
dux
K [N/ m 2 ]
dy

The factor K [Pa sn] is the consistency index or power law coefficient
and n is the power law exponent. For Newtonian flow, n is equal to 1 and
K equals the viscosity. For shear-thinning fluids, n is less than 1 while it is
larger than 1 for shear-thickening fluids.
Newtonian fluids have a constant viscosity independent of shear rate.
Hence, for Newtonian fluids, changing the force (stress) applied to the fluid
will not change their viscosity. The viscosity remains constant as the force
applied changes.
Non-Newtonian fluids instead have viscosities that change according
to the magnitude of the force (stress) that is applied upon the fluid. The
viscosity changes as the force applied changes.

14.10.1 Types of Non-Newtonian Fluid Behavior


The viscosity of non-Newtonian fluids is dependent on shear rate or shear
rate history. The behavior of different types of non-Newtonian fluids is
given in Figure 14.4.
In a Newtonian fluid, the relation between shear stress and the shear
rate is linear, passing through the origin, the constant of proportionality
being the coefficient of viscosity.
For a Bingham Plastic fluid (see Figure 14.4), stress can be applied, but
it will not flow until a certain value, the yield stress, is reached. Beyond
this point, the flow rate increases steadily with increasing shear stress. The
physical reason for this behavior is that the liquid contains particles (such
as clay) or large molecules (such as polymers) that have some kind of inter-
action, creating a weak solid structure, formerly known as a false body,
and a certain amount of stress is required to break this structure. Once
the structure has been broken, the particles move with the liquid under
Rheology and Viscoelasticity of Adhesives 325

Bingham Plastic

Pseudoplastic Fluid

Shear Stress, t
Dilatant Fluid

Newtonian Fluid

Shear Rate, k

Figure 14.4 Non-Newtoninan fluids.

viscous forces. If the stress is removed, the particles associate again. It is


that causes the thixotropic behavior of many adhesives.

(a) Pseudoplastic fluids


They are shear-thinning fluids. The viscosity of these fluids
decreases as the shear rate increases.
(b) Dilatant fluids
Called shear thickening fluids in which the apparent viscosity
increases as the shear rate increases.
(c) Rheopectic fluids
Rheopecty or rheopexy is the rare property of some non-
Newtonian fluids to show a time-dependent increase in vis-
cosity (time-dependent viscosity); the longer the fluid under-
goes shearing force, the higher its viscosity. Rheopectic fluids,
such as some lubricants, thicken or solidify when shaken.
(d) Thixotropic fluids
Thixotropy is a time-dependent shear-thinning property. Certain
gels or fluids that are thick, or viscous, under static conditions
will flow (become thin, less viscous) over time when shaken,
agitated, sheared, or otherwise stressed (time-dependent vis-
cosity). Constant shear stress can be applied by shaking or
mixing.

14.11 Viscoelasticity of Adhesives


Materials respond in specific manners when they are subjected to a force
(stress). An ideal elastic solid follows Hooke’s law, which states that the
326 Adhesives for Wood and Lignocellulosic Materials

stress/strain ratio is constant and is called the Young’s modulus. In con-


trast, an ideal viscous liquid follows Newton’s law, which states that the
stress is proportional to the rate of strain. But there are a number of mate-
rials encountered in normal life that are not ideal. They exhibit both elas-
tic and viscous responses to an applied force. They are called viscoelastic
materials. Polymeric adhesives for instance exhibit such unique viscoelas-
tic responses when subjected to a stress.
Dynamic mechanical analysis (DMA) is the most common technique
for determining viscoelastic properties and their dependence on time and
temperature.

14.11.1 Phenomenological Models for Viscoelastic Materials


Models to describe the behavior of viscoelastic materials are constructed
based on coupled or uncoupled spring and dashpot elements. The spring
represents an ideal elastic material while the dashpot portrays an ideal
Newtonian liquid. Attempts have been made to account for the response
of real materials in terms of suitable combinations of spring and dashspot.
Thus, a viscoelastic material can both store the energy corresponding
to the elastic component and dissipate energy corresponding to the vis-
cous component. In a viscoelastic material, therefore, part of the energy
is stored while the remaining is dissipated through viscous flow. Some of
these models are described below.

14.11.1.1 Maxwell Element (Elastic Deformation + Flow)


A Maxwell element (spring and dashpot in series) (Figure 14.5) symbol-
izes a material that can respond elastically to stress but can also undergo
viscous flow. The two contributions to the strain are additive in this model:
ε  = εelas + εvisc
The force under constant stretch decays exponentially with time in the
Maxwell model. The relaxation time is given by

Kv/Ke

where Ke is the linear spring constant (ratio of force and displacement, in


N/m) and Kv is the linear dashpot constant (ratio of force and velocity, in
Ns/m). Thus, the Maxwell element predicts stress relaxation experiments
but fails to describe a creep experiment. Other viscoelastic models may be
envisioned, and it must be recognized that neither a Maxwell element nor
Rheology and Viscoelasticity of Adhesives 327

I.O

F
X-axis: Time s; Y-axix: σ(t)/σ(0)

Figure 14.5 Maxwell element.

γ∞

F
X-axix: Time (s) and Y-axis: Y-axis: Strain (γ)

Figure 14.6 Voigt element.

any complex combination of springs and dashpots succeeds in portraying


the overall response of a viscoelastic materials.

14.11.1.2 Voigt Element (Spring and Dashpot in Parallel)


A Voigt element symbolizes a material whose elastic response is not
instantaneous but retarded by a viscous resistance. The Voigt model cap-
tures the creep behavior of viscoelastic solids (Figure 14.6).

14.11.1.3 Maxwell–Voigt Mixed Model (Figure 14.7)


G , the storage modulus, represents the elastic deformation of a material
and is a measure of the hardness of the adhesive at a given temperature.
A typical hot melt adhesive has a G at 25°C, varying from 107 to 108 Pa. A
good PSA has a value of 104 to 105 at room temperature.
328 Adhesives for Wood and Lignocellulosic Materials

X-axis: Time; Y-axis: Strain

Figure 14.7 Maxwell–Voigt mixed model.

G , the loss modulus, is associated with energy absorption mechanisms


and correlates to the viscous deformations in a material; i.e., it represents
in a sense the flexibility of an adhesive.

14.12 Dynamic Mechanical Analysis


Adhesives for wood such as urea–formaldehyde (UF), melamine–formal-
dehyde (MF), PF, epoxies, PU, etc. are essentially thermosetting in nature.
They basically exhibit viscoelastic properties. DMA is the most common
technique employed for studying such viscoelastic behavior and its depen-
dence on time and temperature.
DMA and torsional braid analysis (TBA) have shown potential for
detecting thermosetting events [9–12]. For DMA and TBA, glass fiber
braids are typically impregnated with liquid PF, and the cure phenomenon
is detected from changes in storage and loss moduli. Palmese and Gillham
demonstrated that thermoset gelation and vitrification events are accom-
panied by a damping peak as well as a rapid increase in storage modulus
[13]. Likewise, vitrification has been detected in several cure studies of
neat PF resins [12].
In the case of thermosetting resins, as the reaction proceeds from the
liquid state to a gel, there is an increase in molecular weight and viscos-
ity with time initially followed by an accelerated (or abrupt) increase in
viscosity as shown in Figure 14.8. Macroscopically, the thermoset can be
characterized by an increase in its viscosity η approaching infinite vis-
cosity, implying that all the chains become linked to one another at the
gel point through a cross-linked network structure of infinite molecular
Rheology and Viscoelasticity of Adhesives 329

0 Conversion 100%

η0 Ge

Newtonian Network at Hookean


liquid the gel point solid

Figure 14.8 Macroscopic development of rheological and mechanical properties during


network formation.

weight. Figure 14.8 illustrates the macroscopic progress from uncured to


fully cured thermoset [14].
The uncured thermoset, often a mixture of monomers and oligomers, is a
Newtonian liquid. As cure progresses, the viscosity increases with increasing
molecular weight, which can be monitored by rheological measurements. As
the viscosity approaches infinity at the gel point, steady shear measurements
reach their limits. However, oscillatory or dynamic rheology and dynamic
mechanical measurements can characterize material in the gelation region.
Reaction continues beyond the gel point to complete the network forma-
tion. Macroscopically, physical properties such as modulus build to levels
characteristic of a fully developed network. It should be noted that while
DMA may be applied to uncured materials below their gel point, the sam-
ples require a support such as metal shim, glass fabric, or wire mesh [15].
DMA is ideally suited to characterize supported samples from pre-
gelation to the completion of cure, as well as fully cured thermosets.
DMA measures the deformation of a polymeric material as a response
to a small stress applied in a time-varying periodic, e.g., sinusoidal fashion.
The strain response to the sinusoidal stress depends on the property of
the polymeric material. For elastic solids subjected to infinitesimal strain,
Hooke’s law states a linear stress–strain relationship. Hence, under sinusoi-
dal loading of elastic solids, the strain is in-phase with the stress [10, 16].
330 Adhesives for Wood and Lignocellulosic Materials

For a viscous liquid on the other hand, Newton’s law describes a linear
relationship between stress and rate of strain under infinitesimal strain
rate. This implies that viscous liquids respond to a sinusoidal stress with
a strain that is 90° out of phase with the stress.
Thus in the case of an ideal (100%) elastic polymer, the strain will be
in-phase with the stress wave. On the other hand, for an ideal (100%) vis-
cous material, the strain curve will be 90° out of phase (i.e., a cos wave)
with the input sinusoidal stress. This is shown in Figure 14.9.
The dynamic storage modulus (E ), the loss modulus (E ), and the
mechanical damping or internal friction (tan δ = E /E ) are obtained by
this technique. Tan δ gives the ratio of the amount of energy dissipated as
heat to the energy stored during the deformation, and is often the parame-
ter chosen to relate dynamic data to molecular or structural motion in the
sample (Figures 14.9 and 14.10).

δ = 0° δ = 90°

Shear
stress

Shear
strain

100% elastic 100% viscous

Figure 14.9 Response of elastic and viscous materials to sinusoidal stresses.

γ0 strain σ0
stress
Stress σ(t)
Strain γ(t)

0.0 0.0

–γ0 δ –σ0

0 90 180 270 360


ωt (degrees)

Figure 14.10 Strain response to sinusoidal stress.


Rheology and Viscoelasticity of Adhesives 331

In a DMA experiment, two independent properties, storage (E ) and


loss moduli (E ), are measured to provide a complete description of visco-
elastic properties of a polymeric adhesive at a given temperature.
The relationship between the storage and loss modulus is shown below:

E* = E + iE (14.1)

tan δ = E /E (14.2)

Thus, the ratio of loss and storage moduli (Equation 14.2) defines
another useful parameter in DMTA, called tan δ.
E , the storage modulus represents the elastic deformation of a material
and is a measure of the hardness of the adhesive at a given temperature.
A typical hot melt adhesive has a G at 25oC, varying from 107 to 108 Pa.
A good PSA has a value of 104 to 105 at room temperature. E , the loss
modulus, is associated with energy absorption mechanisms and correlates
to the viscous deformations in a material; i.e., it represents in a sense the
flexibility of an adhesive [17].
The loss tangent (tan δ), can be used to define the gel point. The latter
occurs at the point where E´ crosses E and where tan δ equals 1, i.e., as the
crossover point of the storage modulus and the loss modulus (E = E ). The
crossover point is generally accepted as the gel point of gel temperature of
the cure reaction [18].

E”(t) E*(t)

δ(t)

E’(t)

Figure 14.11 Graphical representation of storage and loss moduli (Argand diagram).

E” ~ energy loss in internal


motion

E’ ~ elastic response

Figure 14.12 Significance of elastic response and energy loss.


332 Adhesives for Wood and Lignocellulosic Materials

The physical significance of E and E is illustrated diagrammatically in


Figures 14.11 and 14.12.
That is, if there is no energy loss E , then the ball would have reached the
original position from which it is dropped.
Thus, in a dynamic mechanical test, an oscillating strain (sinusoidal
or other waveform) is applied to a sample and the resulting stress devel-
oped in the sample is measured. The output signals are analyzed, and,
using established mathematical methods, the rheological parameters are
computed.

14.13 TTT and CHT Diagrams


Temperature–time–transformation (TTT) and continuous heating trans-
formation (CHT) cure diagrams provide a complete cure characterization
of thermosetting systems in that they define the time and temperature
required for achieving specific viscoelastic properties. TTT and CHT
diagrams were developed for the first time by Gilham for epoxy resins
[19, 20]. These diagrams define zones of curing, vitrification, and degrada-
tion of resins such as epoxies in their pure state and not in contact with any
resin-interacting substrate. Gilham studied and modeled the times to gela-
tion and to vitrification for the isothermal cure of an amine‐cured epoxy,
measuring on macroscopic and molecular levels by dynamic mechanical
spectrometry (TBA and Rheometrics dynamic spectrometer), infrared
spectroscopy, and gel fraction experiments. The relationships between the
extents of conversion at gelation and at vitrification and the isothermal
cure temperature formed the basis of a theoretical model of the TTT cure
diagram (Figure 14.13), in which the times to gelation and to vitrifica-
tion during isothermal cure versus temperature are predicted. The model
demonstrates that the “S” shape of the vitrification curve depends on the
reaction kinetics, as well as on the physical parameters of the system, i.e.,
the glass transition temperatures of the uncured resin (Tg0), the fully cured
resin (Tg∞), and the gel (gelTg).
In the case of the CHT diagram, Wisanrakkit and Gilham [20] found
that the overall reaction kinetics of a high‐Tg tetrafunctional aromatic
diamine/difunctional epoxy system, which can satisfactorily describe the
rate of the reaction in both kinetically and diffusion‐controlled regimes,
can be determined from isothermal conversion/Tg data by differential scan-
ning calorimetry. They used the mathematical expression of the kinetics,
together with the unique one‐to‐one relationship between Tg and chemical
conversion, to calculate the material’s Tg vs. time under heating at constant
Rheology and Viscoelasticity of Adhesives 333

Degradation

τgo
Devitrification
Ce
ll
ed
ru
bb
er

Celled glass

Culation curve
τcure(°C)

Vitrification curve
Liquid

τgo

Non grilled class

In time

Figure 14.13 Details of a TTT diagram of an epoxy resin adhesive on a non-interacting


glass fiber substrate.

rates. For a heating scan from below Tg0 (the glass transition temperature
of the unreacted material), initial devitrification corresponds to the reac-
tion temperature (Tcure) first passing through the Tg of the reacting mate-
rial; vitrification corresponds to Tg becoming equal to the increasing Tcure
after initial devitrification; and finally, upper devitrification corresponds
to Tg eventually falling below the rising Tcure. The results of the calcula-
tion correlated well with the available experimental data of the dynamic
mechanical behavior of the resin during temperature scans at constant
rates obtained by the TBA.
Although these works were a considerable innovation, they could not be
transferred to resins such as wood adhesives containing water as a solvent
and especially in contact with a lignocellulosic substrate interacting with
both the water and the resin. The diagrams had then to be extensively mod-
ified for the case of water-carried wood adhesives in contact with wood
to take into account the fundamental influence of these two parameters,
water and the absorbent substrate. Furthermore, TTT and CHT diagrams
334 Adhesives for Wood and Lignocellulosic Materials

were difficult and very time-consuming to build with the approach of


Gilham. It is the use of thermomechanical analysis (TMA) techniques that
instead allowed for the first time the determination of these diagrams in an
easier and faster way, just on wood adhesive resins in contact with wood
[9, 21–28].
The use of the equation f = km/αE, correlating number of degrees of
freedom m of polymer segments between cross-linking nodes in polycon-
densation networks to the energy of interaction polymer segment/polymer
segment, both within the same polymer and at different polymer interfaces,
through measures of deflection f in bending by dynamic TMA, yielded
a number of consequences of interest in the field of polycondensation-
hardened networks and of their process of hardening and, among others,
opened the way to an easier determination of TTT and CHT diagrams for
wood adhesives on lignocellulosic substrates [9, 21].
Thus, lignocellulosic substrates such as wood were found to have a
marked modifying influence on a well‐defined region of CHT diagrams
during hardening of PF and UF adhesive resins. This was ascribed to more
complex resin phase transitions due to resin/substrate interactions peculiar
to these substrates. The chemical and physical mechanisms of the interac-
tions of the resin and substrate causing such CHT diagram modifications
were presented and discussed. The Di Benedetto equation [29] describing
the glass transition temperature Tg of the system as a function of the resin
degree of conversion p was then slightly modified to take into account the
modified TTT and CHT diagram [21]. The modified diagrams can be used
to good effect to describe the behavior of polycondensation resins when
used as wood adhesives during their curing directly into the wood joint.
Lignocellulosic substrates such as wood were found to have a marked
modifying influence on both lower-temperature and higher-temperature
zones of TTT and CHT diagrams during hardening of phenol–resorcinol–
formaldehyde (PRF) and melamime–urea–formaldehyde (MUF) polycon-
densates. Although the modifying influence of the substrate on the higher-
temperature zone of CHT diagrams presented the same trend of what
was found first for PF and UF polycondensates, marked differences from
what was reported in the literature were recorded for TTT diagrams of all
these polycondensates as well for the lower-temperature zones of the CHT
diagrams on lignocellulosic substrates [25]. The chemical and physical
mechanisms of the interactions of the resins, the substrate, and the water
carrier causing such marked variations were presented and discussed.
Although in the higher-temperature zones both substrate and water car-
rier play an important role, in the lower-temperature zone, the presence of
water appears to be the dominant factor causing the observed variations.
Rheology and Viscoelasticity of Adhesives 335

The  generalized modified CHT and TTT diagrams characteristic of the


behavior of these water‐borne polycondensates on lignocellulosic sub-
strates can be used to describe the behavior and complex changes of phase
the formaldehyde‐based polycondensation resins undergo when used as
wood adhesives during their curing directly in the wood joint. The results
also show that diagrams obtained with pure resins cannot be used to pre-
dict the behavior of the polycondensate when this is markedly modified by
the presence of interacting solvents and substrates.
Furthermore, the modifying influence of some of the most import-
ant manufacturing parameters of the resins on a CHT diagram were also
explored with some clear trends being shown [27]. In the case of melamine–
urea–formaldehyde (MUF) resins for wood adhesives, the molar ratio
(M + U):F appeared to be the dominant parameter influencing the relative
position of gel and vitrification curves of the CHT diagram in relation to
each other. The ratio of melamine to urea did not appear to have any effect
on the relative position of the curves, lacking any clear trend, at least at the
higher (M + U) molar ratio of 1:1.9 used for this series of resins. Thus, the
influence of resin manufacturing parameters on CHT curing diagrams was
studied in combination with the modifications introduced by the substrate.
TTT and CHT curing diagrams were also built for tannin‐based adhe-
sives by TMA by following the in situ hardening directly in a wood joint
[30]. The curve trends observed were similar to those previously observed
for synthetic polycondensation resins on lignocellulosic substrates. Of the
parameters that most influence the relative position of vitrification and
gel curves on the diagrams (i.e., where the influence has been quantified),
chief among them is the reactivity of the tannin with formaldehyde and any
factor influencing it: thus, the inherent higher reactivity of the A‐ring of
the tannin (such as in procyanidins versus prorobinetinidins) and the pH
of the tannin solution. The percentage formaldehyde hardener has some
influence in CHT diagrams, especially for the slower‐reacting tannins, but
practically no influence in TTT diagrams within the 4–10% formaldehyde
range used. As in the case of synthetic polycondensation adhesive resins,
regression equations relating the internal bond strength of a wood par-
ticleboard, prepared under controlled conditions, with the inverse of the
minimum deflection, obtained by constant heating rate TMA of a wood
joint during resin cure, have been obtained for the two types of tannins
of lower reactivity (profisetinidins and prorobinetinidins) but not for the
faster‐reacting procyanidin and prodelphinidin tannins.
Different trends to those reported for TTT and CHT diagrams of
epoxy resins reported in the literature occur in the higher- and lower-
temperature zones of the diagrams of water-borne formaldehyde-based
336 Adhesives for Wood and Lignocellulosic Materials

resins hardening on wood. CHT and TTT diagrams have been reported for
PF, UF, MUF, PRF, and tannin–formaldehyde thermosetting resins [24, 25].
The experimental TTT diagrams shown in Figure 14.14 show, however,
quite a different trend from the CHT diagram for the same resins (Figure
14.18) and for the TTT diagrams reported in the literature for epoxies on
glass fiber (Figure 14.13). To start to understand the trend shown in Figure
14.14, it is first necessary to observe what happens to the modulus of the
wood substrate alone (without a resin being present) when examined
under the same conditions of a wood joint during bonding. No significant
degradation occurs up to a temperature of 180°C as shown by the relative
stability of the value of the elastic modulus as a function of time. Some
slight degradation starts to occur at 200°C, but after some initial degrada-
tion, the elastic modulus again settles to a steady value as a function of time
and at a value rather comparable to the steady value obtained at lower tem-
peratures. Evident degradation starts to be noticeable in the 220–240°C
range, and this becomes even more noticeable at higher temperatures. The
effect of substrate degradation on the TTT diagram in Figure 14.14 can
then only start to influence the trends in gel and vitrification curves at tem-
peratures higher than 200°C, and it is for this reason that the region of the
curves higher than 200°C are indicated by segmented lines in Figure 14.13.
At a temperature <200°C, the trends observed are only due to the resin.

Substrate degradation Degradation

Solvent remobilisation from substrate


τ g∞
Devitrification Resin to τg∞
No
n
geGe
llelle Gelled glass
d
ru d ru
Tcure (°C)

Pseudo gel bb bb
curve er er

Gelation curve Vitrification curve

Liquid
To time– ∞
τgo-O°C
Due to solvent
Non gelled glass

In time

Figure 14.14 Generalized TTT diagram of water-carried PF resin adhesives on an


interacting lignocellulosic substrate (wood).
Rheology and Viscoelasticity of Adhesives 337

In this range of temperature, the eventual turning to longer time and stable
temperature of the vitrification curve, characteristic of the TTT diagrams of
epoxy resins, becomes also evident for the TTT diagrams of the water-borne
PF and other formaldehyde resins on lignocellulosic substrates, indicating
that diffusion hindrance at a higher degree of conversion also becomes, for
these resins, the determinant parameter defining reaction rate. What how-
ever differs from previous diagrams is that the trend of all the curves, namely,
gel curve, initial pseudogel (entanglement) curve, and start and end of vit-
rification curve, is the same. In epoxy resins, TTT diagrams of the trend of
the gelation curve are completely different from what was reported here. The
result shown in Figure 14.14 is however rather logical because if diffusion
problems alter the trend of the vitrification curve, then the same diffusional
problem should also alter the gel and pseudogel curves. This is indeed what
the experimental results in Figure 14.14 indicate. It may well be that in water-
borne resins, the effect is more noticeable than in epoxy resins. This is the
reason why it is possible to observe it for PF, UF, PRF, and MUF resins. With
the data available and with the limitation imposed by the start of wood sub-
strate degradation at higher temperatures, it is not really possible to say if the
gel curve and the vitrification curve run asymptotically towards the same
value of temperature at time = ∞, although the indications are that this is
quite likely to be the case. What is also evident in the trend of the two curves
is the turn to the right, hence the inverse trend of their asymptotic tendency
towards Tg∞. This turn cannot be ascribed to substrate degradation because
for very reactive resins, such as PRFs, such a turn already occurs at a tem-
perature lower than 150°C, hence much lower than the temperature at which
substrate degradation becomes significant. This inverse trend can only be
attributed to movements of water coming from the substrate towards the
resin layer as the curve trend indicates an easing of the diffusional problem
already proven to occur at such a high degree of conversion [24, 25].
Two other aspects of the TTT diagrams in Figures 14.13, 14.14, and
14.15 must be discussed, these being the trend of the curves at tempera-
tures higher than 200°C and the trend of the devitrification (or resin deg-
radation) curve. The segmented line trend and experimental points of all
the curves at temperatures higher than 200°C are clearly only an effect
caused by the ever more severe degradation of the substrate: degradation
of the substrate infers a greater mobility of the polymer network consti-
tuting the substrate, hence the continuation of the curves as shown in
their segmented part. That this is the case is also supported by the vir-
tual negative times yielded by the TMA equipment when the temperature
becomes extreme, as well as by the trend of the resin’s higher degradation
curve, which tends to intersect the vitrification curve at about 200–220°C
338 Adhesives for Wood and Lignocellulosic Materials

Degradation

Devitrification curve
No Ge
n lle gelled glass
ge
lle d
Tcure (°C)

d ru
ru bb
bb er
er Vitrification curve
Pseudo gel curve

Gelation curve

Liquid To time–∞
τgo-O°C
Due to solvent
Non gelled glass (frozen sovent)

In time

Figure 14.15 Generalized CHT diagram of water-carried PF resin adhesives on an


interacting lignocellulosic substrate (wood).

or higher, this being a clear indication that one is measuring the changes
in the reference system, the substrate itself, and that these are at this stage
so much more important than the small changes occurring in the resin to
be able to dominate the whole complex system, which is the bonded joint.
The CHT and TTT diagrams pertaining to water-borne formaldehyde-
based polycondensation resins on a lignocellulosic substrate should then
appear in their entirety as shown in Figures 14.14 (TTT) and 14.15 (CHT).

14.14 Experimental Results


Before describing some more recent results in DMA, it will be worth
describing the considerable series of experiments that were done in deter-
mining by TMA by studying the curing behavior and viscoelastic prop-
erties of PF, MUF, UF, and tannin adhesives and their correlation with
the board properties obtained. The method of following the progressive
curing, hardening, and degradation of the adhesives directly on a wood
joint was developed by the research group of Pizzi in the late 1990s and
early 2000s [31–35]. To obtain the TMA samples, the system consisted in
placing the adhesive in its viscous liquid state between two layers of beech
Rheology and Viscoelasticity of Adhesives 339

wood veneer of 0.6-mm thickness as shown in Figure 14.16, and a three-


point bending mode was applied. The adhesive thickness was 0.2 mm, but
the total tested sample dimensions were slightly different from those in
Figure 14.16 at 21 × 6 × 1.4 (mm). The specimens were then tested by
TMA in three points bending on a span of 18 mm exercising a force cycle
of 0.1/0.5 N on the specimens with each force cycle of 12 s (6 s/6 s).
PF, MUF, and UF resin adhesives were tested in this manner at a con-
stant temperature rate of increase of 10°C/min up to the hardening of
the adhesive glue mixes and the TMA results related to the IB strength of
boards made with the same resins [31]. Equally, commercial pine (Pinus
radiata) and mimosa (Acacia mearnsii) bark tannin extract solutions in
water were tested by TMA at different temperatures and at different per-
centage paraformaldehyde hardener addition both under isothermal con-
ditions and at a constant temperature rate of increase of 10°C/min up to
the hardening of the tannin resin glue mix, and the results were compared
with those obtained with the ABES equipment [36]. Equally in the case of
PF and UF adhesives, the effect of humidity on viscoelastic response and
cross-linked and entanglement networking of formaldehyde-based wood
adhesives were tested by the same TMA technique in bending [31]. The
principle of this TMA (and DMA) approach is that as the adhesive is still
a viscous liquid, only the lower beech veneer lamina resents the bending

1.0

141
0.8
Modulus/max. modulus

0.6
119
173
0.4 181
193

0.2

0.0
0 50 100 150 200
Temperature (°C)

Figure 14.16 TMA curve of the hardening of a PF resin in situ in a beech wood joint.
Increase of MOE of the joint as a function of temperature at 10°C/min constant heating
rate (O); first derivative (Δ) (from Ref. [37]).
340 Adhesives for Wood and Lignocellulosic Materials

stress imparted to the sandwich, but as the resin starts progressively to


thicken and to cure, the stress is equally progressively felt and transmit-
ted to both the glue line and to the upper beech lamina of the sandwich.
It is still thus possible not only to follow the progressive variation of the
adhesive to liquid to solid, but also to rapidly compare different adhesives,
or also to compare how the viscoelasticity of an adhesive behaves under
different curing conditions [34].
TMA on wood joints bonded with PF adhesives, as in Figure 14.16, has
shown that, frequently, the joint increase in modulus does not proceed in
a single step but in two steps, yielding an increase of the modulus first
derivative curve presenting two major peaks rather than the single peak
obtained for mathematically smoothed modulus increase curves [37]. This
behavior has been found to be due to the initial growth of the polycon-
densation polymer leading first to linear polymers of critical length for the
formation of entanglement networks. The reaching of this critical length is
greatly facilitated by the marked increase in concentration of the PF poly-
mer due to the loss of water on absorbent substrates such as wood coupled
to the linear increase of the average length of the polymer due to the ini-
tial phase of the polycondensation reaction. The combination of these two
effects lowers markedly the level of the critical length needed for entangle-
ment. Two modulus steps and two first derivative major peaks then occur,
with the first due to the formation of linear PF oligomer entanglement
networks, and the second one due to the formation of the final covalent
cross-linked networks. The faster the reaction of phenolic monomers with
formaldehyde, or the higher the reactivity of a PF resin, the earlier and at
lower temperature the entanglement network occurs, and the higher is its
modulus value in relation to the joint modulus obtained with the final,
covalently cross-linked resin.
Suitable sample geometry for the DMA measurement of liquid adhe-
sives can be the “wire mesh geometry” as shown in Figure 14.17 with the
following details:

• Support for liquid or paste materials.


• Note that the wire mesh is used on a 45° bias; otherwise, the
experiment would be dominated by the high stiffness wire.
When used in the 45° bias, the properties of the resin will
dominate the response.
• Controlled thickness, sensitive measure of Tg.
• Can extract thermoset G , G via composite analysis.
Rheology and Viscoelasticity of Adhesives 341

Stationary
clamp
Wire mesh sample
45° bias

Movable
clamp

Figure 14.17 Wire mesh geometry.

Lei and Frazier [38] used filter paper since this was proven to be suit-
able as the substrate for the preparation of DMA testing specimens to
predict the curing behavior of PF resin adhesives for its stability during
the curing temperature span. With this method, the curing behav-
ior of PF resin was monitored by DMA in tensile-torsion mode. With
the strain curves, the onset of curing temperature of PF resin could be
determined clearly. The curing degree of PF resin could be calculated
by the integral area in strain curves. The method to combine storage
modulus (G ), tan δ, and strain curves together could explain the curing
behavior of PF resin more comprehensively than the commonly used
method using only G and tan δ curves. The DMA test results of PF resin
with different viscosity and with accelerator implied the reliability of
this novelty method.
The viscoelastic properties of the blends of MF resin and poly(vinyl
acetate) were studied by DMTA by Kim and Kim [39]. The DMTA ther-
mogram of MF resin showed that the storage modulus (E) increased as
the temperature was further increased as a result of the cross-linking
induced by the curing reaction of the resin. E of MF resin increased both
as a function of increasing temperature and with increasing heating rate.
Kim and Kim [40] followed on the early work of other groups [9, 21, 30]
on the curing behavior and viscoelastic properties of tannin-based adhe-
sives. They studied the curing behavior and viscoelastic properties of two
types of tannin-based adhesives, wattle and pine, with three hardeners by
DMTA:

(a) Paraformaldehyde
(b) Hexamethylenetetramine
(c) TN [tris(hydroxylmethyl)nitromethane]
342 Adhesives for Wood and Lignocellulosic Materials

28 mm
5 mm 0.2 mm tannin-based adhesive
0.6 mm sliced beecd veneer
1.4 mm

Figure 14.18 Sample configuration for the DMTA (three-point bending mode) test.

Paraformaldehyde was shown to be more reactive with wattle tannin


than the other hardeners, while hexamethylenetetramine was the most
reactive with the pine tannin.
The storage modulus (E ), loss modulus (E ), and loss factor (tan δ) of
each adhesive system were obtained by DMTA. With increasing tempera-
ture, as the tannin-based adhesives hardened, the storage modulus (E )
increased in all adhesive systems.
Kim and Kim [40] imitated the same testing system first developed
much earlier for formaldehyde resins by another research group [9, 21,
26, 30–34, 36, 37, 41], as shown by Figure 14.18. As the adhesives were in
the viscous liquid state, to obtain the DMTA samples, the tannin-based
adhesives were placed between two layers of beech wood veneer of 0.6-mm
thickness as shown in Figure 14.18, and a three-point bending mode was
applied. The adhesive thickness was 0.2 mm, and the total tested sample
dimensions were 28 × 5 × 1.4 (mm). During the DMTA experiments, the
static force was kept at 20% of the dynamic force and the frequency was
maintained at 1 Hz with a strain of 0.05.

References
1. Barnes, H.A., Hutton, J.F., Walters, K., An Introduction to Rheology, Elsevier,
Amsterdam, The Netherlands, 1989.
2. Chanda, M., Advanced Polymer Chemistry, Marcel Dekker, New York, 2000.
3. Anonymous, How does a rotational viscometer work? http://ttwud.org/
how-does-a-rotational-viscometer-work/ 2000.
4. Rheosys, A basic introduction to viscometers and viscosity, http://www.rh[9]
eosys.com/intro.html 2011.
5. Yang, X., Organic fillers in phenol–formaldehyde wood adhesives, PhD
Dissertation, Virginia Polytechnic Institute and State University, Blacksburg,
VA, 2014.
6. Gupta, R.K., Polymer and Composite Rheology, CRC Press, Boca Raton, FL,
2000.
7. Klimets, M., Torque and its calculation, http://earthphysicsteaching.home
stead.com/Torque_Lab.html 2002.
Rheology and Viscoelasticity of Adhesives 343

8. Halasz, L., Vorster, O.C., Pizzi, A., Guasi, K., A rheology study of the gelling
of phenol–formaldehyde polycondensates. J. Appl. Polym. Sci., 80, 898–902,
2001.
9. Pizzi, A., On the correlation of some theoretical and experimental param-
eters in polycondensation cross-linked networks. J. Appl. Polym. Sci., 63,
603–617, 1997.
10. Laborie, M.-P., Investigation of the wood/phenol–formaldehyde adhesive
interphase morphology, PhD Dissertation, Virginia Polytechnic Institute
and State University, Blacksburg, VA, 2002.
11. Steiner, P.R. and Warren, S.R., Rheology of wood–adhesive cure by torsional
braid analysis. Holzforschung, 35, 273–278, 1981.
12. Kim, M.G., Nieh, W.L.-S., Meacham, R.M., Study of phenol–formaldehyde
resol resins by dynamic mechanical analysis. Ind. Eng. Chem. Res., 30, 798–
803, 1991.
13. Palmese, G.R. and Gilham, J.K., Time–temperature-transformation (TTT)
cure diagrams relationship between Tg and the temperature and time of cure
for a polyamic acid/polyimide system. J. Appl. Polym. Sci., 34, 1925–1939,
1987.
14. Winter, H.H., Baumgaertel, M., Sockey, P.R., A parsimonious model for
viscoelastic liquids and solids, in: Techniques in Rheological Measurement,
A.A. Collyer (Ed.), pp. 123–160, Chapman and Hall, London, 1997.
15. Prime, B.R., Dynamic Mechanical Analysis of Thermosetting Materials,
2013. https://ruc.udc.es/dspace/bitstream/handle/2183/11489/CC-80%20
art%2013.pdf?sequence=1&isAllowed=y.
16. Ward, I.M., Mechanical Properties of Solid Polymers, 2nd edn., John Wiley &
Sons, New York, 1979.
17. Franck, A.J., Adhesives Rheology, Brochure Rheometrics, AFERA Congress,
Chester, UK, 1992.
18. Kendall, E.W., Benton, L.D., Trethewey, B.R., Jr., Thermal analysis of the
polymerization of phenol formaldehyde resins. Proc. TAPPI Plast. Laminates
Symp., 1, 117–132, 2000. https://imisrise.tappi.org/TAPPI/Products/DIL/
DIL0408.aspx.
19. Enns, J.B. and Gillham, J.K., Time–temperature–transformation (TTT) cure
diagram: Modeling the cure behavior of thermosets. J. Appl. Polym. Sci., 28,
2831–2844, 1983.
20. Wisanrakkit, G. and Gillham, J.K., Continuous heating transformation
(CHT) cure diagram of an aromatic amine/epoxy system at constant heating
rates. J. Appl. Polym. Sci., 41, 1985–1995, 1990.
21. Pizzi, A., On the correlation of some theoretical and experimental parame-
ters in polycondensation cross-linked networks, Part 2: Interfacial energy and
adhesion on cellulose substrates. J. Appl. Polym. Sci., 65, 1843–1847, 1997.
22. Pizzi, A., Extension of simple polycondensation gelation theories to simple
radical and mixed polycondensation/radical gelation. J. Appl. Polym. Sci., 71,
517–521, 1999.
344 Adhesives for Wood and Lignocellulosic Materials

23. Pizzi, A., On the correlation equations of liquid and solid 13C NMR, TMA,
Tg and network strength in polycondensation resins. J. Appl. Polym. Sci., 71,
1703–1709, 1999.
24. Pizzi, A., Lu, X., Garcia, R., Lignocellulosic substrates influence on TTT and
CHT curing diagrams of polycondensation resins. J. Appl. Polym. Sci., 71,
915–925, 1999.
25. Pizzi, A., Zhao, C., Kamoun, C., Heinrich, H., TTT and CHT curing dia-
grams of water-borne polycondensation resins on lignocellulosic substrates.
J. Appl. Polym. Sci., 80, 2128–2139, 2001.
26. Pizzi, A., Garcia, R., Deglise, X., Thermomechanical analysis of entanglement
networks—Correlation of some calculated and experimental parameters.
J. Appl. Polym. Sci., 67, 1673–1678, 1998.
27. Properzi, M., Pizzi, A., Uzielli, L., Influence of preparation parameters on
the variation of CHT curing diagrams of MUF polycondensation resins.
J. Appl. Polym. Sci., 81, 2821–2825, 2001.
28. Garnier, S. and Pizzi, A., CHT and TTT curing diagrams of polyflavonoid
tannin resins. J. Appl. Polym. Sci., 81, 3220–3230, 2001.
29. Pascault, J.P. and Williams, R.J.J., Glass transition temperature versus
conversion relationships for thermosetting polymers. J. Polym. Sci.,
Part B, Polym. Phys., 28, 85–95, 1990. https://doi.org/10.1002/polb.1990.
090280107.
30. Garnier, S., Huang, Z., Pizzi, A., Commercial tannin adhesives-bonded
particleboard IB forecasting by TMA bending. Holz Roh Werkst., 59, 46,
2001.
31. Kamoun, C., Pizzi, A., Garcia, R., The effect of humidity on cross-linked
and entanglement networking of formaldehyde-based wood adhesives.
Holz Roh Werkst., 56, 235–243, 1998.
32. Laigle, Y., Kamoun, C., Pizzi, A., Particleboard, I.B., forecast by TMA
bending in UF adhesives curing. Holz Roh Werkst., 56, 154, 1998.
33. Zhao, C., Garnier, S., Pizzi, A., Particleboard dry and wet IB forecasting
by gel time and dry TMA bending in PF wood adhesives. Holz Roh Werkst.,
56, 402, 1998.
34. Zhao, C., Pizzi, A., Garnier, S., Fast advancement and hardening acceler-
ation of low condensation alkaline PF resins by esters and copolymerized
urea. J. Appl. Polym. Sci., 74, 359–378, 1999.
35. Kamoun, C. and Pizzi, A., Particleboard IB forecast by TMA bending in
MUF adhesives curing. Holz Roh Werkst., 58, 288–289, 2000.
36. Lecourt, M., Humphrey, P., Pizzi, A., Comparison of TMA and ABES as
forecasting systems of wood bonding effectiveness. Holz Roh Werkst., 61,
75–76, 2003.
37. Garcia, R. and Pizzi, A., Cross-linked and entanglement networks in ther-
momechanical analysis of polycondensation resins. J. Appl. Polym. Sci., 70,
1111–1116, 1998.
Rheology and Viscoelasticity of Adhesives 345

38. Lei, H. and Frazier, C.E., A dynamic mechanical analysis method for pre-
dicting the curing behaviour of phenol–formaldehyde resin adhesive.
J. Adhes. Sci. Technol., 29, 981–990, 2015.
39. Kim, S. and Kim, H.-J., Thermal stability and viscoelastic properties of
MF/PVAc hybrid resins on the adhesion for engineered flooring in under
heating system; ONDOL. Thermochimica Acta, 444, 134–140, 2006.
40. Kim, S. and Kim, H.-J., Curing behaviour and viscoelastic properties of
pine and wattle tannin-based adhesives studied by dynamic mechanical
thermal analysis and FT-IR-ATR spectroscopy. J. Adhes. Sci. Technol., 17,
1369–1383, 2003.
41. Zhao, C., Pizzi, A., Kuhn, A., Garnier, S., Fast advancement and hardening
acceleration of low condensation alkaline PF resins by esters and copoly-
merized urea. Part 2: Esters during resin reaction and effect of guanidine
salts. J. Appl. Polym. Sci., 77, 249–259, 2000.

You might also like