You are on page 1of 26

J. Adhesion Sci. Technol., Vol. 21, No. 7, pp.

637– 661 (2007)


 VSP 2007.
Also available online - www.brill.nl/jast

Fundamentals of hot-melt pressure-sensitive adhesive


tapes: the effect of tackifier aromaticity

EMMETT P. O’BRIEN ∗ , LOUIS T. GERMINARIO, GARY R. ROBE,


TIM WILLIAMS, DOUG G. ATKINS, DEBBIE A. MORONEY and
MARK A. PETERS
Eastman Chemical Company, P.O. Box 1972, Kingsport, TN 37662, USA

Received in final form 16 February 2007

Abstract—We revisit the effect of varying the tackifier aromaticity on adhesive performance in
a styrenic block co-polymer (SBC)-based pressure-sensitive adhesive tape. Aromaticity strongly
affects the compatibility between the tackifying resin and base polymer and is defined as the relative
number of protons attached to an aromatic ring. Dynamic mechanical analysis, peel resistance,
tack, shear resistance and atomic force microscopy were utilized to detect changes in adhesive
performance due to resin aromaticity. The results indicate that increasing the aromaticity of the
tackifier enhanced the compatibility between the tackifier and styrene domains. As a consequence, the
high-temperature cohesive performance measured by the shear adhesion failure temperature decreased
due to plasticization of the end-blocks by the aromatic resins. However, other properties of interest
such as the glass transition of the rubbery matrix, tack, peel resistance and holding power were less
affected by the resin aromaticity. We note that this paper is of limited scope because we have focused
this research on styrene–isoprene–styrene, which is very compatible with a wide range of tackifiers.
More significant differences would be expected in formulations based on styrene–butadiene–styrene
or styrene–isoprene–butadiene–styrene polymers in which the compatibility between the resin and
midblock would be more sensitive to the resin aromaticity. In addition to our results, we review the
literature and some fundamentals of formulating SBC-based adhesives and PSA testing.

Keywords: Tackifier aromaticity; ring-and-ball-softening point; butadiene; nanostructure; micro-


domains; morphology; end-block associating resin; mid-block-associating resin; release paper.

1. INTRODUCTION

Pressure-sensitive adhesives (PSAs) are a class of soft polymeric materials that


adhere to surfaces under light contact pressure and short contact times without

∗ Towhom correspondence should be addressed. Tel.: (1-423) 229-2338; Fax: (1-423) 229-3217;
e-mail: eobrien@eastman.com
638 E. P. O’Brien et al.

chemical reaction or solvent evaporation [1]. In addition, these materials display


significant debonding resistance after adhering to the surface. The characteristic
properties of PSAs are due to their softness (low modulus at the application
temperature) and viscoelastic properties [2, 3]. Under small applied loads, the
material behaves like a liquid and readily deforms and wets the surface permitting
the attractive intermolecular forces between the adhesive and substrate surfaces to
form a bond. The debonding resistance and capability to support significant load
are attributed to viscous dissipation (cavitation and fibrillation) and to an increase
in adhesive modulus as the strain rate or debond velocity increases [4].
Styrenic block co-polymers (SBCs) are an important class of high-performance
polymers used in the PSA tape industry which are typically composed of two
styrene end blocks connected by a diene midblock [5]. SBCs have useful properties
due to their two-phase structure consisting of hard glassy styrene phase and a
soft rubbery diene phase and can form either spherical, cylindrical or lamellar
microstructures [5]. The glassy domains are connected to one another via the
rubbery mid-block. As a consequence, the styrene domains form physical cross-
links between the rubbery mid-block domains. An important advantage of these
materials is that above the glass-transition temperature (Tg ) of styrene, the SBC
loses its cohesive strength and can be melt processed but below this temperature,
the styrenic domains greatly increase the cohesion of the system.
The shear modulus, G , of a PSA at ambient temperature ranges typically between
103 and 106 Pa, and the Tg of a PSA is typically below 10◦ C [6, 7] or 30 to 70◦ C
lower than the application temperature [8]. As a result, the properties of neat SBCs
are not suitable for most PSA applications. SBCs inherently have a Tg that is too
low and a G that is too high at room temperature. Viscoelastic processes cannot
be activated if the Tg is too far from the application temperature. Furthermore, the
adhesive cannot deform and wet the surface if it is too stiff. As a consequence, SBCs
must be formulated with tackifying resins and plasticizers to achieve the desired
properties. Most tackifying resins have a low molecular weight (Mw in the range
of 300 to 2000) and a Tg above ambient temperature, typically near 40 to 50◦ C. In
SBCs, these tackifiers increase the Tg of the mid-block (yet keep the value below
ambient temperature) and enable the adhesive to behave viscoelastically at room
temperature. They also reduce the rubbery modulus of the tackifier-SBC blend by
plasticizing the rubbery matrix and by increasing the fraction of the rubbery phase
[9 –11]. Plasticizers such as oils are also used to fine-tune the adhesive properties
by reducing the Tg , stiffness and/or viscosity [12]. Brown and co-workers [4]
studied styrene–isoprene–styrene (SIS)-based PSAs and elucidated the mechanism
of how tackifiers improved adhesion. They found that the highest adhesion values
were observed when cohesive failure occurred, as opposed to interfacial failure
or cavitation initiating at the interface. This agrees with the general observation
that misleadingly high peel values are usually observed for adhesives that exhibit
cohesive failure rather than interfacial failure. They also noted that it was beneficial
for a PSA to maximize strain, without reducing the maximum stress. From a
The effect of tackifier aromaticity on hot-melt PSA tapes 639

viscoelastic standpoint this means softening the adhesive at high extensions and low
deformation rates (during debonding) while maintaining or increasing the stiffness
of the adhesive at small extensions and high deformation rates (during debond
initiation). Tackifiers accomplish this by increasing the Tg and raising the modulus
at high strain rates, while diluting the entanglements of the rubbery phase. For
SBCs, this argument assumes that the tackifier resin is compatible with the mid-
block.
Chu and co-workers conducted some of the first research examining how tackifiers
affected the viscoelastic properties of adhesives. They studied the influence of
tackifier chemistry [13], molecular weight [14] and concentration [15] on the
viscoelastic and rheological properties of natural-rubber-based, styrene–butadiene-
rubber-based and SIS-based adhesives. They also looked at how viscoelastic
behavior was affected by a mid-block-associating tackifier [16], as well as by an
end block-associating tackifier [17]. Class and Chu [13] found that the tackifier and
base polymer must have similar chemistry to be compatible. However, tackifiers
that were expected to be compatible based on the chemistry showed evidence of
incompatibility above a critical Mw [14]. Class and Chu also showed that the
tackifier became immiscible with the rubbery phase at a critical concentration.
Immiscibility was evident from the two glass transitions (of the rubbery matrix)
observed from rheological measurements and the microstructure determined from
microscopy. However, the work of Class and Chu did not examine how the resin
characteristics affected adhesive performance. Readers interested in a review of the
chemistry of tackifying resins are referred to the work of Donker [18] and Benedek
[8, 19].
To determine whether a tackifier is compatible with a base polymer, it is useful to
characterize these materials by their relative aromaticity and polarity. Aromaticity
can be analyzed via nuclear magnetic resonance (NMR) spectroscopy to determine
the relative number of protons attached to an aromatic ring or via cloud point
determinations using an appropriately chosen solvent system [20]. Relative polarity
and aromaticity can be determined using the diacetone alcohol–xylene (DACP)
cloud point or methylcyclohexane–aniline (MMAP) cloud point, respectively [20].
In a general sense, the less aromatic the tackifier is, the more aliphatic the tackifier.
The convention of describing a tackifying resin as aromatic or aliphatic is also
useful for describing the base polymer. For example, the three monomers most
commonly used in SBCs, in order of increasing polarity and aromaticity are
isoprene, butadiene and styrene. It is expected that an aromatic tackifier will
preferentially associate with the styrene end-block and a less aromatic (or more
aliphatic) tackifier will associate more with the rubbery mid-block (i.e., with
isoprene and butadiene). It was, therefore, the goal of this study to systematically
look at the effect of resin aromaticity on the rheological properties, morphology and
adhesive performance. Furthermore, the question of how resin aromaticity affects
adhesive tape performance is very relevant today given increasing raw material
costs and availability restrictions of some key raw materials. One formulation
640 E. P. O’Brien et al.

option would be to replace all or part of the isoprene in SBCs with butadiene.
A butadiene-containing formulation requires the tackifying resin to be more polar
and/or aromatic to achieve compatibility by more closely matching the resin’s
solubility parameter with that of butadiene.
In addition to the common tests of adhesive properties developed by The Pressure
Sensitive Tape Council and the ASTM, we have employed atomic force microscopy
(AFM), which can image the surface and sub-surface structures of a PSA. Scanning
probe microscopy has been used to study the surface morphology [21] and the
ensuing changes that occur during the aging of PSAs [22, 23]. Paiva and Foster [22]
studied poly(ethylene propylene) with a rosin ester tackifier and found that tackifier-
enriched domains increased in size and that the polymer-enriched matrix increased
in stiffness as the adhesive was aged. Doring et al. [24] presented work which
showed that AFM was an excellent tool for characterization of the adhesive phase
morphology and correlated the results to probe tack measurements. In this work, we
have utilized AFM to characterize the morphology and determine structure-property
relationships of some PSA formulations as a function of tackifier chemistry.

2. EXPERIMENTAL

Many experimental details gathered from our research experience are included in
this section to provide the reader the information that is essential for working with
and solving problems in the area of SBC-based PSAs. This section is intended
to be particularly useful to the reader inexperienced in PSAs to circumvent some
commonly encountered problems. Readers experienced with SBC-based adhesives
may choose to skip this section.

2.1. Ingredients and formulations


Pressure-sensitive adhesives were formulated using a general-purpose PSA-tape
formulation containing 100 phr (parts per hundred resin) of base polymer, 110 phr
tackifier and 15 phr oil. The base polymer was Kraton D1161K (Kraton Polymers
LLC), a linear tri-block co-polymer based on styrene and isoprene (SIS) with
a polystyrene content of 15 wt% and a di-block (styrene–isoprene) content of
19 wt%. The reader interested in how increasing the di-block content leads to
an increase in tack and a reduction in cohesive strength is referred to the work of
Nakajima and co-workers [10] and Gilbert and co-workers [25]. The tackifiers were
Piccotac® 1095, Piccotac® 9095, Piccotac® 8095, Piccotac® 7590-C and Piccotac®
6095-E (Eastman), in order of increasing aromaticity. The oil was Calsol 5550
(Calumet Lubricants). Calsol 5550 is a naphthenic process oil composed of
4% aromatic carbons, 46% naphthenic carbons and 50% parafinnic carbons, as
measured by ASTM D3238. All specimens were mixed with 2 phr Irganox 1010
anti-oxidant (Ciba Specialty Chemical). The tackifying resins were characterized
using test methods typical for the PSAT industry: (1) MMAP and DACP cloud point
The effect of tackifier aromaticity on hot-melt PSA tapes 641

measurements, (2) softening point measurement, (3) dynamic mechanical analysis,


(4) differential scanning calorimetry and (5) molecular weight distribution.

2.2. Blending and coating


Adhesive formulations were solvent blended. 50 g of adhesive was mixed with
approximately 100 g of toluene in a small jar and placed overnight on a roller. The
solution was then poured into a large tray lined with Teflon® sheet and the solvent
was evaporated in a vacuum oven at 60◦ C overnight. Second heat scans from the
rheology measurements revealed that no residual solvent was left in the adhesive.
It is well known that different mixing conditions can lead to significant differences
in the final adhesive properties. Solvent blending and solvent coating eliminates the
effects of both high temperature and shear imparted by the mechanical mixer which
can significantly affect adhesive rheology and performance. However, solvent
blending may lead to difference in morphology due to difference in the rate of
solvent removal as a function of adhesive thickness [26]. The morphology can also
be significantly altered by annealing or aging [27]. Readers interested in the effects
of processing (solvent vs. melt blending, solvent vs. melt coating) on the adhesive
properties of PSAs are referred to the work of O’Connor and Macosko [26] and Kim
et al. [28]. As a rule of thumb, increasing mixer shear and temperature lead to better
mixing of the adhesive components and a breakdown in molecular weight and a
subsequent increase in tack and a dramatic reduction in shear performance. In some
cases, these effects may be significant enough to overwhelm any effects caused by
changing the adhesive formulation; and for this reason, some researchers choose
to solvent-blend and solvent-coat samples rather than use a mechanical mixer or
extruder.
The adhesive was melt coated onto 50 µm polyester film (Mylar) using a knife
coater heated to 150◦ C. The coating weight was approximately 20 ± 2 g/m2 ,
which gives a thickness of roughly 17.5 ± 1 µm. Note that the thickness of the
soft adhesive layer can significantly affect adhesive performance. For example,
removable protective films are designed to have low tack and are, therefore, very
thin. Conversely, duct tape is extremely tacky because the adhesive layer is very
thick. For these experiments, a release paper was used for ease of handling. Note
also that the release paper after contacting the PSA surface can adversely affect the
adhesive performance and, in the worst case, can contaminate the surface enough to
eliminate any rolling ball tack. Readers interested in the role of carrier and release
paper on tack should see Benedek [8, 29]. Initial and aged adhesive performances
were measured. Initial (un-aged) tape specimens were conditioned at 25◦ C and 50%
relative humidity for 24 h prior to testing. Aged specimens were conditioned for an
additional 2 weeks at 40◦ C and ambient humidity. Aging of specimens is very
important given that the adhesive properties may change significantly with time.
This is because the formulation is composed of low molecular weight tackifiers and
oils, which can diffuse readily in the rubbery matrix. Adhesion measurements were
642 E. P. O’Brien et al.

made on steel cleaned with acetone as the substrate. The number of specimens for
each test was at least five.

2.3. Rheology (dynamic mechanical analysis)


Dynamic mechanical analysis (DMA) is a method used to study the mechanical
characteristics of polymers (and other materials) and has been applied for rheologi-
cal studies (i.e., for the flow behavior of adhesives) [30 –32]. Rheological properties
of the adhesive (prior to coating and with no backing) were measured on a TA In-
struments Ares RDA3 Rheometer in a parallel-plate geometry and auto strain mode.
The diameter of the plates was 8 mm and the gap was 1.704 mm. The frequency was
10 Hz and the heating rate was 6◦ C/min. The maximum strain was set at 5.0%. Sec-
ond heat scans were also performed to determine if there was any residual solvent
and if there was any significant adhesive morphology change due to heat history.
Rheology data were very repeatable and the uncertainty in the glass transition was
approximately ±0.5◦ C.
The reader interested in understanding how rheology can be used to predict
adhesive performance is referred to the work by Chu [30], White and co-workers
[31], Mazzeo [32] and Benedek [8, 29]. A number of parameters were obtained
from the rheology experiments to predict structure–property relationships and
ultimately adhesive tape performance. These include the first tan δ maximum
(tan δmax ) temperature, which is a measure of the Tg of the rubbery matrix. In
addition, the tan δ maximum (tan δmax ) peak height indicates how much energy the
adhesive can dissipate [33]. The storage modulus G at room temperature (25◦ C)
was also noted, to quantify how compliant the adhesive was at the application
temperature. The tan δ minimum (tan δmin ), above the Ig of the rubbery phase,
was also measured to characterize the cohesive strength of the adhesive. Lastly,
the third cross-over temperature was measured. The third cross-over temperature
(tan δ = 1max ), illustrated in Fig. 5b for the formulation containing Piccotac® 1095,
is the temperature near the Tg of the glassy polystyrene domains at which the storage
and loss moduli are equal and, therefore, tan δ is equal to 1. The third crossover
temperature (tan δ = 1max ) is the temperature where the adhesive begins to flow and
loses its cohesive strength. Note there are three points in the tan δ spectra where
tan δ is equal to 1. The first and second cross-over points, where tan δ is equal
to 1, occur at temperatures much lower than the third, slightly below and above the
Tg of the rubbery matrix. However, the first and second cross-over points are less
important than the third and do not readily correlate to end use performance. In
this research, we determined if any of these rheological parameters correlated to
adhesive performance and the resin aromaticity.
DMA is an excellent tool for determining (1) the general rheological properties
of adhesives (such as Tg , storage modulus G at room temperature (25◦ C), or
tan δ = 1max ) and (2) if the adhesive components are compatible. However, it is
never enough to make conclusions based solely on DMA. Adhesive performance
The effect of tackifier aromaticity on hot-melt PSA tapes 643

testing (peel, tack and cohesive strength) is always recommended. Numerous


examples exist where an adhesive with “good” rheological properties exhibits poor
adhesive performance, often due to incompatibility of some components, surface
migration of an immiscible component or because the coated adhesive layer is
too thin. For this reason, DMA is often regarded as a screening tool to warrant
further adhesive testing. We note that a good rule of thumb for determining if
two adhesive formulations are different is if the difference between the Ig values
is 3◦ C or more. We also caution the reader that when comparing formulations
with different resins it is important to first characterize the resin (aromaticity,
softening point and molecular weight) to determine the source of any observed
differences.

2.4. Tack

While everyone agrees that PSAs are tacky materials, there is no general agreement
on a definition of what tack actually is. Tack is such an elusive property of a PSA
that there are at least three common ways of measuring it [4]. All of these tests
measure the ability of the adhesive to form a bond in a very short time with minimal
contact force. None of these: rolling ball tack (RBT), loop tack, or probe tack
produce results that correlate well with each other. The adhesives produced for this
study were all tested for RBT and loop tack.
Rolling ball tack experiments were performed in accordance with PSTC-6 [34].
In this experiment, an 11.1 mm diameter generic steel ball is rolled down a bench
top ramp on to a 25 mm × 250 mm strip of tape. The distance the ball travels along
the tape is recorded. The less the ball rolls down the tape, the tackier the adhesive.
The ramp was a Cheminstruments Rolling Ball Tack Tester. We caution the reader
that the cleanliness and quality of the ball can significantly affect the RBT value.
As mentioned previously, the use of release paper can also significantly affect the
rolling ball tack results due to transfer of silicone from the paper to the tape surface.
This agrees with work by Tse [35], who investigated the effect of oil on tack and
found that RBT was influenced more by surface than bulk properties. It is important
to remember that high tack levels correspond to low RBT distances (the ball sticks
to the tape and does not roll). If the sample has an RBT distance greater than 30 cm,
the adhesive tack was not enough, the ball was not clean or was not chosen correctly
or the coating weight was too low.
Loop tack tests [36 –38] were performed on a Cheminstruments Loop Tack Tester
in accordance with PSTC-16 [34]. The cross-head displacement rate was 5 mm/s.
A 25 mm × 125 mm loop of tape was used in the experiments. The free loop of
tape, unrestricted by the grips, was 75 mm long. The maximum force per unit width
of the specimen was recorded. The initial height, measured from the bottom of the
grips to the substrate surface, was 50 mm. The maximum displacement was 44 mm
and the dwell time at maximum displacement was 1 s.
644 E. P. O’Brien et al.

2.5. 180◦ peel test


The peel energy or peel force per unit width was measured in accordance with
PSTC 101: Peel Adhesion of Pressure Sensitive Tape Test Method A — Single-
Coated Tapes, Peel Adhesion at 180◦ Angle [34]. Rectangular strips of 25 mm ×
250 mm dimensions were tested using a Universal Testing Machine (UTM) at
5 mm/s cross-head displacement rate.

2.6. Shear adhesion failure temperature


Shear adhesion failure temperature (SAFT) was measured using a Cheminstruments
30 Bank Shear Tester. These experiments were conducted to measure the high-
temperature cohesive or shear properties of the adhesive tape. A 25 × 25 mm area
of PSAT was adhered to stainless-steel coupons with a standard 2 kg roller [34].
A mass of 1 kg was suspended from the tape. Samples were placed in an oven
and equilibrated at 40◦ C. The oven temperature was then increased at 0.5◦ C/min.
The temperature at which the adhesive failed cohesively was recorded as the shear
adhesion failure temperature.

2.7. Holding power


Holding power measurements were made with a modification of PSTC-107 method,
180◦ Shear Adhesion of Pressure Sensitive Tapes [34] using a Cheminstruments
30 Bank Shear Tester. These experiments were conducted to measure the room
temperature cohesive or shear properties of the adhesive tape. Instead of the
25 mm × 25 mm contact area described in PSTC-107, a 12.5 mm × 12.5 mm area
of PSAT was utilized. The PSAT was adhered to stainless steel coupons with a
standard 2 kg roller [34] and a mass of 1 kg was suspended from the tape. The
time (h) at which the adhesive failed cohesively was recorded as the holding power.
After reaching a maximum time of 167 h (10 000 min), the adhesives were removed
from the test apparatus. Note that these adhesives hold much longer than 167 h if
the standard bonded area of 25 mm × 25 mm is used. As a consequence, reducing
the bonded area, and effectively increasing the shear load on the tape by a factor
of 4, is a useful method to reduce the time to failure and to increase the sensitivity
for detecting differences in PSAT performances.

2.8. Atomic force microscopy


In order to better understand the structure-property behaviors of these soft materials,
AFM was utilized in TappingMode™ (intermittent contact). Samples of tape
approximately 1 × 2 cm2 were cut and attached to glass slides using a double-
sided tape. Surface imaging was performed with a commercial atomic force
microscope, Dimension series D3000 AFM (Digital Instruments). The probes were
NanoDevices Metrology ProbeTM , Multi75 (LOT Oriel), with a spring constant of
5 N/m, resonance frequency of 75 kHz and tip radius <10 nm. AFM was also
The effect of tackifier aromaticity on hot-melt PSA tapes 645

used to provide a measure of surface viscoelastic response by phase imaging in the


intermittent contact mode. Phase image contrast was calibrated using TiO2 particles
dispersed on a polymer substrate. Under these conditions, TiO2 particles are bright
with darker surrounding. Therefore, the hard phase appears white and the soft phase
is dark.

2.9. Cloud point measurements


Cloud point measurements were made to determine the relative aromaticity, polarity
and compatibility of the tackifying resin [18, 39, 40]. In these tests, tackifying resin
is dissolved in either methylcyclohexane – aniline (MMAP) or diacetone alcohol –
xylene (DACP) in a 25 mm diameter test tube at 140◦ C. The solution is then
allowed to cool to ambient temperature and the temperature where the solution
becomes cloudy and the resin becomes insoluble is recorded as the cloud point
temperature. To determine the relative aromaticity, 5 g resin is dissolved in 15 ml
MMAP, a 1:2 mixture (by volume) of methylcyclohexane and aniline. To determine
the relative polarity, 5 g resin is mixed with 10 g DACP, a 1:1 (by weight) mixture of
xylene and diacetone alcohol (4-hydroxyl-4-methyl-2-pentanone). As the polarity
increases, DACP cloud point decreases. The relative aromatic and aliphatic contents
of tackifying resins are expressed usually by plotting the MMAP cloud point as a
function of DACP cloud point. MMAP and DACP cloud points are linearly related
for most resins; the lower the value of the MMAP cloud point, the lower the value
of the DACP cloud point [18]. The more aromatic and less aliphatic the resin is,
the lower are the cloud points. We note that the standard deviation in MMAP cloud
point is typically 2◦ C, but can be as much as 6◦ C. The standard deviation in DACP
cloud point is much more than the MMAP cloud point and can be as much as 12◦ C.
We also employed NMR to determine the aromaticity of the resin. NMR avoids
the differences introduced by the operator and the age and moisture content of the
DACP which are known to affect the cloud point measurement. However, NMR
cannot be used alone to predict resin–polymer compatibility because NMR does
not account for the effect of molecular weight on compatibility. Although these
results are not presented here, the relative aromaticity measured by NMR correlates
very well to both MMAP and DACP cloud points.

2.10. Thermal transition of the resin


Since tackifier resins are amorphous polymers they do not exhibit clear changes
of state such as melting point. The most readily detectible state change for these
materials is the Tg , a second-order transition typically defined as a discontinuity in
the rate of change in heat capacity of a material as the temperature changes. The
glass transition is normally measured by differential scanning calorimetry (DSC).
An indication of the resin Tg can be obtained by subtracting 50◦ C from the ring-
and-ball-softening point. The softening point of the resin is actually a viscosity
measurement and does not correspond to a change of physical state of the test
646 E. P. O’Brien et al.

material, unlike a Tg or melting point. It is widely accepted in the adhesives industry


due to ease of measurement and low equipment cost. Widely used instruments for
measuring softening points are the Herzog Ring and Ball Tester and the Mettler
Drop Point Tester. For the Herzog method, the ring and ball softening point (RBSP)
is defined as the temperature at which a disk of the 19.7 mm × 4.4 mm sample
imbedded in a ring is forced downward under the weight of a 9.5 mm diameter
steel ball. The specimen is heated in silicone oil at a rate of 5 ± 0.5◦ C/min. The
Mettler Drop Point Tester uses a sample ring 1.0 mm in diameter and 0.8 mm deep
and a disposable ball 0.8 mm in diameter. The test is run in air at a heating rate of
5 ± 0.5◦ C/min and the end point occurs when a liquid drop of the resin falls from
the ring. Typically, the Mettler RBSP is a few degrees lower than the Herzog RBSP.

2.11. Molecular weight distribution


Size-exclusion chromatography (SEC) was used to characterize the molecular
weight distribution of the resins. The Mn , Mw , Mz and polydispersity (PDI)
were determined versus polystyrene standards. The standard deviation was 1%.
Molecular weight is very important for compatibility; the higher the molecular
weight, the less compatible the resin is. Mz is regarded to be the most important
of the molecular weight parameters, as it is the highest molecular weight species
that are the least compatible. The lower molecular weight fractions reflected in Mn
are more readily mixed with the base polymer. Furthermore, if one compares the
molecular weights among different resins, one will discover that Mn and Mw are
very similar, but that Mz values can be significantly different.

3. RESULTS AND DISCUSSION

3.1. Tackifying resin properties


Table 1 shows the relative aromatic and aliphatic contents of the tackifying resins
characterized using MMAP and DACP cloud point measurements and NMR. In
addition to the relative aromaticity, the RBSP (Herzog) and molecular weight
information are also shown in Table 1. The tackifying resins listed in order of least
aromaticity to highest aromaticity are: (1) Piccotac® 1095, (2) Piccotac® 9095,
(3) Piccotac® 8095, (4) Piccotac® 7590-C and (5) Piccotac® 6095-E. Therefore,
we would expect 6095-E, the most aromatic and most polar of the resins, to have
the greatest affinity for the aromatic styrene end blocks of the base polymer and
the least affinity for the isoprene midblock. As shown in Fig. 1, the MMAP
cloud point is linearly related to the % aromaticity measured by NMR. These
tackifying resins were chosen to focus our investigation on the effect of resin
aromaticity on adhesive properties and minimize the influences of molecular weight
and softening point. Ideally, in such a study, the resins would have different levels of
aromaticity, but identical softening points and molecular weights. However, due to
The effect of tackifier aromaticity on hot-melt PSA tapes 647

Table 1.
Properties of the tackifying resins used in this research

Resin Aromaticity MMAP DACP RBSP Mz × 10−3 Mw × 10−3 Mn × 10−3 PDI*


(%, NMR) cloud cloud (◦ C)
point point
(◦ C) (◦ C)
Piccotac® 1095 0.5 94 47 99 3.3 1.7 0.82 2.1
Piccotac® 9095 2.0 88 47 95 4.0 1.9 0.93 2.0
Piccotac® 8095 5.2 76 38 95 5.7 2.2 0.94 2.3
Piccotac® 7590-C 7.5 65 27 93 3.2 1.6 0.82 1.9
Piccotac® 6095-E 14.8 35 −10 96 5.0 1.6 0.67 2.4
* Measured by SEC.

Figure 1. The MMAP cloud point as a function of % aromatic content of the tackifying resin
determined by NMR. The aromaticity and MMAP cloud point are directly correlated.

the polymerization process, it is difficult to synthesize resins with the same softening
point, same molecular weight but different aromaticity. Therefore, we have chosen
resins that vary in aromaticity but have similar softening points (between 93 and
99◦ C) and different molecular weight distributions. An additional consequence
is that we could not determine how the molecular weight of the resin influenced
adhesive performance. The similarities in softening point and Tg are highlighted
by examining Figs 2 and 3. Figure 2 shows the DSC results for the different
neat tackifying resins. Figure 3 shows the thermal transition measurements of
neat tackifier resins measured by Herzog RBSP, Mettler RBSP and DSC (5 and
10◦ C/min). These results show the close similarity in thermal properties among the
resins as well how the value of the glass transition temperature of the tackifying
resins varies depending on the analytical method.
648 E. P. O’Brien et al.

Figure 2. The heat flow (exotherm-up) measured by DSC for the different resins. The legend is
as follows: (3) Piccotac® 1095; (Q) Piccotac® 9095; (P) Piccotac® 8095; (2) Piccotac® 7590-C;
(×) Piccotac® 6095-E.

Figure 3. Thermal transition measurements of neat tackifier resins. The legend is as follows:
(Q) Mettler RBSP; (") Herzog RBSP; (P) DSC 5◦ C/min; (!) DSC 10◦ C/min. These results show
the close similarity in thermal properties among the resins as well how the value of the Tg of the
tackifying resins varies depending on the analytical method.
The effect of tackifier aromaticity on hot-melt PSA tapes 649

3.2. The effect of resin aromatic content


3.2.1. DMA. Figure 4 shows storage modulus G and tan δ as a function of
temperature for the five different formulations with varying resin aromaticity. There
appear to be two molecular relaxations below ambient temperature, in addition to
the high temperature styrene phase, which suggests that there are two phases present
in the rubbery matrix. The primary relaxation, near 0◦ C, is the relaxation of the
tackifier-rich isoprene phase. Near −40◦ C is a small shoulder in the tan δ that is
likely the isoprene-rich phase. This feature has also been previously observed from
rheology by Tse [11] and Class and Chu [14, 15]. The presence of two phases
in the rubbery matrix is also indicated by the AFM images. Readers interested in
more information on the existence of multiphase structures in tackified PSAs and
the miscibility of resins are referred to Benedek [19].
Figure 5 shows that increasing the aromaticity leads to a decrease in the tempera-
ture where the modulus dramatically falls off (above 70◦ C, as shown in Fig. 5a), as
well as a decrease in the third cross-over point, tan δ = 1max point (Fig. 5b and 5c).
Again, the tan δ = 1max temperature is defined as the highest temperature in the
rheology spectrum where the value of tan δ is equal to unity (G = G ). This is the
temperature at which the adhesive loses significant cohesive strength and viscous
flow begins to dominate the rheology of the system as the styrene domains soften
and begin to flow. An increase in the tan δ = 1max temperature usually indicates
an increase in the high temperature shear properties (SAFT). Based on Figs 4, 5b

Figure 4. Storage modulus G and tan δ as a function of temperature for each adhesive formula-
tion. The legend is as follows: (3) Piccotac® 1095; (2) Piccotac® 9095; (P) Piccotac® 8095;
(1) Piccotac® 7590-C; (Q) Piccotac® 6095-E.
650 E. P. O’Brien et al.

(a)

(b)
Figure 5. The storage modulus, G , as a function of temperature for each adhesive formulation. The
tan δ as function of temperature for each adhesive formulation. A line is drawn to indicate the third
cross-over temperature (tan δ = 1max ), where tan δ is equal to 1. The third cross-over temperature,
tan δ = 1max is an indication of the high-temperature cohesive strength of the adhesive. The third
cross-over temperature, tan δ = 1max , as a function of the % aromatic content of the resin. Increasing
the aromaticity of the resin lowers the temperature where the adhesive begins to flow.
The effect of tackifier aromaticity on hot-melt PSA tapes 651

(c)
Figure 5. Continued.

Figure 6. The Tg of the tackifier-rich isoprene matrix as a function of the % aromatic content of
the resin. No correlation exists between the % aromaticity of the resin and the glass transition of the
tackifier-rich isoprene phase.

and 5c, a decrease in the tan δ = 1max with increasing resin aromaticity indicates
that the tackifying resin is preferentially associating with the styrene end blocks. In
fact, we see that the relative difference in tan δ = 1max between the least and most
aromatic resins is approximately 30◦ C. This indicates that the Tg of the styrenic
phase is shifting downward as the domains are being plasticized by resins which
have a lower Tg than styrene.
Figure 6 shows that the Tg of the tackifier-rich rubbery phase does not correlate to
the aromatic content of the resin. Although the aromatic resins clearly associate
with the polystyrene domains, the total amount of resin in the formulation that
652 E. P. O’Brien et al.

migrates from the rubbery midblock to the polystyrene domains remains small.
Consequently, there is no heither a significant change in the resin content in the
rubbery matrix nor a decrease in the Tg of the tackifier-rich rubbery phase as the
resin aromaticity changes. This is not unexpected, given that the styrene content of
the SIS base polymer is 15% and it is only 6.5% of the total adhesive formulation;
thus, as a consequence, the relative amount of resin that plasticizes the end-blocks is
low. Furthermore, the Tg of the rubbery matrix was found not to correlate with either
the softening point or Tg of the resin, as would be predicted by the Fox equation
[41, 42]. This was not surprising because the softening points and Tg values of
these resins were so similar (see Figs 2 and 3).

3.2.2. Adhesive performance. The adhesive performance data, initial and aged,
as a function of aromatic content of the tackifying resin are shown in Fig. 7.

(a)

(b)
Figure 7. Adhesive performance data as a function of the % aromatic content of the resin. Initial data
are shown with (P) and the aged data are shown for (2). Loop tack, peel force, RBT, SAFT, holding
power. Only the SAFT values correlate to the aromaticity of the resin.
The effect of tackifier aromaticity on hot-melt PSA tapes 653

(c)

(d)

(e)
Figure 7. Continued.
654 E. P. O’Brien et al.

The trends change slightly after aging which shows that these materials reach
equilibration during aging. This is a common observation in the PSA industry,
and is not surprising, given that these are rubbery materials which are above their
Tg even at ambient temperature. These results show that increasing the aromaticity
of the resin results in a strong decrease in the high temperature performance of
these adhesives, as measured by SAFT (Fig. 7d). Furthermore, the decrease in
SAFT correlates well to the tan δ = 1max measured by rheology (Fig. 5c). Peel
force, loop tack and RBT do not appear to be affected by the resin aromaticity.
Interestingly, the holding power of the adhesive formulated with the most aromatic
resin (Piccotac® 6095-E) is greater than for the other less aromatic resins. We
believe that increasing the aromaticity of the resin somewhere between 7.5 and
14.8% (measured by NMR) causes the resin to change from a mid-block-associating
resin to an end-block-associating resin. End block-associating resins increase the
room temperature shear strength by increasing the size of the styrene domains, as
well as by increasing the relative volume fraction of the glassy phase. These glassy
spherical domains, of course, act as physical cross-links and increase the holding
power. Another formulation technique to increase holding power is to use a resin
that is immiscible with the rubbery matrix. In this case, the resin forms incompatible
glassy domains in the rubbery matrix that behave similar to a reinforcing agent or
filler.
There is much discussion in the literature about the storage modulus, G , and
loss modulus, G which describe the viscoelastic behavior of PSAs. These two
parameters are particularly useful for describing if an adhesive meets the Dahlquist
criterion (G  105 Pa) [43] and if the compound lies within the viscoelastic
window for PSAs [6, 7, 44]. However, in these experiments no correlation
between resin aromaticity and either G or G was consistently observed in our
experiments; nor was there any correlation between the adhesive performance and
either G or G . Readers interested in the effect of tackifying resin on the modulus
of PSAs are referred to the classic work by Ferry [45], Brown and co-workers [4],
Cantor [42], Nakajima and co-workers [10] and Benedek [19].
It is also of interest to determine if the aromaticity of the resin affected the
melt and coating behavior. Figure 8 shows the melt viscosity measured at 177◦ C
with a Brookfield viscometer as a function of the aromatic content. Interestingly,
the viscosity as a function of the % aromatic content follows a similar trend to
rolling ball tack values and holding power. The resins with a middle range of
aromaticity have the highest tack and lowest viscosity. Furthermore, the higher
viscosity samples have higher holding power. This is also not surprising given that
adhesive performance is a balance between cohesion and tack. That is, holding
power is often directly correlated to viscosity and inversely related to tack.

3.3. AFM
AFM was used to obtain phase images (soft vs. hard) of the surface and the
subsurface of the adhesive tapes. Surface measurements were made by “light
The effect of tackifier aromaticity on hot-melt PSA tapes 655

Figure 8. The melt viscosity at 177◦ C as a function of % aromatic content of the resin. No correlation
between the viscosity and resin aromaticity was observed.

tapping”. Subsurface images were obtained by tapping at higher forces, which


reduces the contribution of the soft surface and increases the contribution from the
hard domains below the surface. As an example, surface and subsurface images
of the tape formulated with the least aromatic resin (Piccotac® 1095) are shown
in Fig. 9. The surface is absent of any polystyrene and is mostly soft tackifier
(dark phase). It is not surprising to find the surface rich in soft material, given
that the soft phase would possess some molecular mobility and the surface would
thermodynamically favor the presence of low molecular weight resin and oil.
The subsurface of the PSA shows a spherical morphology characterized by glassy
styrene domains in a rubbery matrix of isoprene and tackifier. We observe that
the AFM images consist of white styrene domains imbedded in an inhomogeneous
matrix (Figs 9 and 10) suggesting that that rubbery matrix is not composed of
a single phase. Rheological evidence also supports the existence of a two-phase
rubber matrix from the bimodal tan δ curve (Fig. 4). The rubbery matrix is likely
composed of a tackifier-rich phase and an isoprene-rich phase. This was surprising,
because if there were indeed two phases we would expect the tackifier to be
incompatible with the isoprene midblock. If the resin was incompatible, we would
expect to observe an isoprene-rich phase and a second phase composed mostly of
immiscible resin. Surprisingly, we observed the opposite, a resin-rich phase of
isoprene and resin and a second phase composed mostly of isoprene (an isoprene-
rich phase). The question of why the tackifier would partition in such a way can
be reconciled if we realize that the tackifier and oil are the major components and
isoprene is the minor component in this adhesive formulation. Tse [11] and Class
and Chu [14, 15] also confirmed the presence of a shoulder at −40◦ C using rheology
and transmission electron microscopy (TEM) in natural rubber-based and SIS-based
adhesives. However, the reasons for the existence of the third phase were not clearly
explained.
656 E. P. O’Brien et al.

Figure 9. AFM phase images of the surface and subsurface of the lowest aromaticity resin
(Piccotac® 1095). Each image is a 2 × 2 µm2 area of the sample. In the AFM image, the relative
stiffness of the sample area is indicated by color. The harder the surface, the lighter the image will
appear; or the softer the surface, the darker the image will appear. The surface (left image) is mostly
dark, suggesting that the surface is composed of soft rubbery adhesive. In the sub-surface (right image)
the morphology of the adhesive consists of glassy styrene domains imbedded in a rubbery matrix.

Bulk images of all the samples are shown in Fig. 10. We observe that the
formulations made with Piccotac® 8095 and 7590-C show the greatest contrast
between rubbery and hard domains. Specimens formulated with the least aromatic
resins, Piccotac® 1095 and 9095, exhibit less definition and more of the softer
isoprene-rich phase. Perhaps increasing the aromaticity, relative to Piccotac® 1095
and 9095, or reducing the melt viscosity (Fig. 8) leads to a better dispersion
of the tackifier in the rubbery phase and produces more contrast and definition
of the microstructure. The more defined microstructure may also explain why
Piccotac® 8095 and 7590-C produce the tackiest adhesives. Piccotac® 6095-E,
the most aromatic resin, shows the coarsest structure, and is the only one of the
five specimens with a significantly different styrenic phase appearance. This is
potentially one of the most important observations from this series of experiments
because the difference in morphology of the adhesive made with Piccotac® 6095-E
may explain some of the characteristics seen through rheological analysis and
adhesive testing. The swelling of the styrenic domains could account for the higher
room temperature holding power of this adhesive. Since more of the Piccotac®
6095-E tackifier is apparently partitioning into the styrenic domains, this adhesive
behaves as if it has a higher crosslink density since the effective weight fraction of
the glassy domains is higher for this sample than any of the others tested. This is an
important observation also because apparently at some point between 7.5 and 14.8%
The effect of tackifier aromaticity on hot-melt PSA tapes 657

Figure 10. AFM phase images of the bulk of all adhesive blends. Each image is a 2 µm × 2 µm area
of the sample. Note that the most aromatic resin (Piccotac® 6095-E) produces the coarsest structure
and the intermediate level aromatic resins (Piccotac® 8095 and 7590-C) produce the finest structure.

aromaticity the tackifier begins to partition significantly into the styrenic phase of
the SBC.
The diameter of the polystyrene domains was measured via particle size analysis
on the subsurface images using the AFM software. Table 2 summarizes the
average and standard deviation of the measured particle diameter size of the glassy
polystyrene domains for the different adhesive formulations. The size of the glassy
domains for the adhesive formulated with Piccotac® 6095-E was much greater than
for the formulations with low aromaticity resins (0.5 to 7.5% aromatic), which was
supported by a two sample t-test. Particle size data, therefore, reinforce the finding
that among the resins in this study, the most aromatic resin (Piccotac® 6095-E)
associates most with the end block. This also suggests that between an aromaticity
level of 7.5% and 14.8%, the resin becomes an end block-associating tackifier.
Since this study was confined to using only commercially available tackifiers, it is
beyond the scope of this work to more precisely define the cross-over point between
midblock and end-block-associating resins. In addition, potential effects of other
variables such as the resin molecular weight, Tg , or processing may also impact the
extent of resin partitioning to the styrenic phase.
Phase angle measurements on the bulk phases were also performed for all
formulations. The phase angle measures the average elasticity or “hardness” of the
scanned 2 × 2 µm2 area. A greater phase angle means a harder surface. Figure 11
658 E. P. O’Brien et al.

Table 2.
Particle diameter size of the glassy polystyrene domains for the
different adhesive formulations measured by AFM

Resin Aromaticity (%) Glassy domain size (nm)


Piccotac® 1095 0.5 14.3 ± 4.5
Piccotac® 9095 2.0 15.7 ± 5.7
Piccotac® 8095 5.2 14.0 ± 3.8
Piccotac® 7590-C 7.5 14.9 ± 4.5
Piccotac® 6095-E 14.8 22.1 ± 10.6

Figure 11. The AFM phase angle as a function of aromatic content. Increasing the aromaticity of the
resin resulted in a harder adhesive, i.e., higher phase angle.

shows that the phase angle increases as the resin aromaticity increases. This is not
unexpected, given that the rheology measurements show that level of aromaticity
correlates to the extent of end block-association. As the resin aromaticity increases,
one would expect the domain size to gradually increase leading to an increase in the
relative volume fraction of the glassy spherical domains. We believe that the degree
of end block-association is reflected in the phase angle measurements.

4. SUMMARY
The effect of varying the tackifier aromaticity in an SBC-based adhesive was
tested using standard PSTC and ASTM tests, dynamical mechanical analysis
(DMA) and atomic force microscopy (AFM). Using adhesive performance tests,
the most significant effect was in the high temperature performance reflected in
the shear adhesion failure temperature (SAFT). No major differences due to resin
aromaticity were observed in rolling ball tack, peel resistance, loop tack or holding
power. DMA, a tool used to predict adhesive performance based on viscoelastic
properties, showed only small differences in the Tg of the rubbery matrix among the
formulations, which was not directly attributable to resin aromaticity. However, the
The effect of tackifier aromaticity on hot-melt PSA tapes 659

reduction in SAFT was reflected in the third crossover temperature, tan δ = 1max .
This was clearly due to plasticization of the polystyrene domains by the tackifying
resin. AFM was found to be a good tool to understand the effect of resin aromaticity.
AFM and DMA showed that the rubbery matrix consisted of two phases: one
phase composed mostly of tackifier and isoprene and a second phase composed
mostly of isoprene. AFM phase images also indicate that the most aromatic resin
(Piccotac® 6095-E) associates more strongly with the endblock than any of the
other resins with lower aromaticity. This suggests that at some point increasing the
aromaticity above 7.5% the resin begins to partition strongly into the styrenic phase.
The partitioning into the end-block was reflected in the increased values of holding
power. Future work will test the effect of resin aromaticity on SBC formulations
based on styrene–butadiene–styrene (SBS) or styrene–isoprene–butadiene–styrene
(SIBS) polymers in which the compatibility between the resin and midblock would
be more sensitive to the resin aromaticity. In this case, we will likely observe
more significant differences in adhesion, rheology and morphology due to tackifier
aromaticity.
We note that there must be a caveat attached when discussing any PSA experi-
ments as it is commonly observed that the adhesive properties can be significantly
altered by varying the weight fraction of each ingredient, the blending and coat-
ing procedures [26, 28], aging conditions (as we observed), adhesive thickness, and
backing film type and thickness. No attempt was made in this work to optimize
the adhesive formulation as would be the common practice in the industry. All of
the adhesive systems studied contained the same proportions of rubber, tackifier,
oil, and antioxidant. No doubt each of these formulations could be manipulated to
maximize one or more adhesive properties and that for each tackifier studied that
optimum formula will probably be different. Such optimization was beyond the
scope of this study. The formulations were chosen to provide a basis for compari-
son such that the only real performance requirement was that each formulation used
should have measurable peel adhesion, tack, holding power and SAFT.

Acknowledgements
The authors thank Judy Dean, Peter Chang and Eddie Forbes for performing the
thermal transition analysis of the resins. We also thank George Caflisch and
Mark Chase for the size-exclusion chromatography analysis. We are also grateful
to Arved Harding for help with the data analysis and both Jeremy Lizotte and
A. J. Pasquale for reviewing the manuscript. Lastly, the authors acknowledge Jos
Ooms, Roelof Luth, Chretien Donker, Peter Dunckley and the reviewers for valuable
discussions on pressure-sensitive adhesives.

REFERENCES
1. S. H. Moon, A. Chiche, A. M. Forster, W. H. Zhang and C. M. Stafford, Rev. Sci. Instrum. 76,
Art. No. 062210 (2005).
660 E. P. O’Brien et al.

2. D. Satas (Ed.), Handbook of Pressure Sensitive Adhesive Technology. Van Nostrand, New York,
NY (1989).
3. C. Creton, in: Processing of Polymers, H. E. H. Meijer (Ed.), pp. 707–741. VCH, Weinheim
(1997).
4. K. Brown, J. C. Hooker and C. Creton, Macromol. Mater. Eng. 287, 163–179 (2002).
5. P. Bahadur, Curr. Sci. 80, 1002–1007 (2001).
6. E. P. Chang, J. Adhesion 34, 189–200 (1991).
7. E. P. Chang, J. Adhesion 60, 233–248 (1997).
8. I. Benedek, Pressure-Sensitive Adhesives and Applications. Marcel Dekker, New York, NY
(2004).
9. M. F. Tse and L. Jacob, J. Adhesion 56, 79 (1996).
10. N. Nakajima, R. Babrowicz and E. R. Harell, J. Appl. Polym. Sci. 44, 1437–1456 (1992).
11. M. F. Tse, J. Adhesion Sci. Technol. 3, 551–570 (1989).
12. C. Galan, C. A. Sierra, J. M. Gomez Fatou and J. A. Delgado, J. Appl. Polym. Sci. 62, 1263–1275
(1996).
13. J. B. Class and S. G. Chu, J. Appl. Polym. Sci. 30, 805–814 (1985).
14. J. B. Class and S. G. Chu, J. Appl. Polym. Sci. 30, 815–824 (1985).
15. J. B. Class and S. G. Chu, J. Appl. Polym. Sci. 30, 825–842 (1985).
16. J. Kim, C. D. Han and S. G. Chu, J. Polym. Sci. Part B – Polym. Phys. 26, 677–701 (1988).
17. C. D. Han, J. Kim, D. M. Baek and S. G. Chu, J. Polym. Sci. Part B: Polym. Phys. 28, 315–341
(1990).
18. C. Donker, Proceedings of the Pressure Sensitive Tape Council, pp. 149–164 (2001).
19. I. Benedek (Ed.), Pressure-Sensitive Design, Theoretical Aspects (Vol. 1). VSP, Leiden (2006).
20. J. Simmons, Adhesives Age 28, (Nov. 1996).
21. J. Mallegol, O. Dupont and J. L. Keddie, J. Adhesion Sci. Technol. 17, 243–256 (2003).
22. A. Paiva and M. D. Foster, J. Adhesion 75, 145 (2001).
23. A. Paiva, N. Sheller, M. D. Foster, A. J. Crosby and K. R. Shull, Macromolecules 34, 2269–2276
(2001).
24. A. Doring, J. Stahr and S. Zollner, Proceedings of the Pressure Sensitive Tape Council, pp.
213–222 (2000).
25. F. X. Gilbert, G. Marin, C. Derail, A. Allal and J. Lachat, J. Adhesion 79, 825–852 (2003).
26. A. E. O’Connor and C. W. Macosko, J. Appl. Polym. Sci. 86, 335–3367 (2002).
27. S. Akiyama, Y. Kobori, A. Sugisaki, T. Koyama and I. Akiba, Polymer 41, 4021–4027 (2000).
28. D. J. Kim, H. J. Kim and G. H. Yoon, J. Adhesion Sci. Technol. 18, 1783–1797 (2004).
29. I. Benedek (Ed.), Developments in Pressure-Sensitive Products. CRC Press, Boca Raton, FL
(2005).
30. S. G. Chu, in: Handbook of Pressure Sensitive Adhesive Technology, D. Satas (Ed.), p. 158. Van
Nostrand, New York, NY (1989).
31. C. C. White, M. R. VanLandingham, P. L. Drzal, N. K. Chang and S. H. Chang, J. Polym. Sci.
Part B: Polym. Phys. 43, 1812–1824 (2005).
32. F. Mazzeo, Proceedings of the Pressure Sensitive Tape Council, pp. 139–148 (2002).
33. N. G. McCrum, B. E. Read and G. Williams, Anelastic and Dielectric Effects in Polymeric Solids.
Wiley, New York, NY (1967).
34. Test Methods for Pressure Sensitive Adhesive Tapes, Pressure Sensitive Tape Council, North-
brook, IL (2000).
35. M. F. Tse, J. Adhesion 70, 95–118 (1999).
36. R. H. Plaut, N. L. Williams and D. A. Dillard, J. Adhesion 76, 37–53 (2001).
37. Y. Woo, R. H. Plaut, D. A. Dillard and S. L. Coulthard, J. Adhesion 80, 203–221 (2004).
38. I. Rivals, U. Personnaz, C. Creton, F. Simai, P. Roose and S. van Es, Measurem. Sci. Technol.
16, 2020–2029 (2005).
39. M. F. Tse, J. Adhesion 66, 61–88 (1998).
The effect of tackifier aromaticity on hot-melt PSA tapes 661

40. J. A. Schlademan, Proceedings of the Pressure Sensitive Tape Council (2003).


41. T. G. Fox, Bull. Am. Phys. Soc. 1, 123 (1956).
42. A. S. Cantor, J. Appl. Polym. Sci. 77, 826–832 (2000).
43. C. A. Dahlquist, in: Adhesion Fundamentals and Practice, p. 2. McLaren, London (1966).
44. S. G. Chu, in: Adhesive Bonding, L. H. Lee (Ed.). Plenum, New York, NY (1991).
45. J. D. Ferry, Viscoelastic Properties of Polymers. Wiley, New York, NY (1980).

You might also like