Sunteți pe pagina 1din 761

Pressure-Sensitive Adhesives

and Applications
Second Edition, Revised and Expanded

Istvan Benedek
Wuppertal, Germany

MARCEL

MARCELDEKKER,
INC. -
NEWYORK BASEL
D E K K E R
The first edition of this book was published as Pressure-Sensitive Adhesives
Technology, István Benedek and Luc J. Heymans (Marcel Dekker, Inc., 1996).
Although great care has been taken to provide accurate and current information,
neither the author(s) nor the publisher, nor anyone else associated with this publi-
cation, shall be liable for any loss, damage, or liability directly or indirectly caused or
alleged to be caused by this book. The material contained herein is not intended to
provide specific advice or recommendations for any specific situation.
Trademark notice: Product or corporate names may be trademarks or registered
trademarks and are used only for identification and explanation without intent to
infringe.
Library of Congress Cataloging-in-Publication Data
A catalog record for this book is available from the Library of Congress.
ISBN: 0-8247-5059-4

This book is printed on acid-free paper.


Headquarters
Marcel Dekker, Inc., 270 Madison Avenue, New York, NY 10016, U.S.A.
tel: 212-696-9000; fax: 212-685-4540
Distribution and Customer Service
Marcel Dekker, Inc., Cimarron Road, Monticello, New York 12701, U.S.A.
tel: 800-228-1160; fax: 845-796-1772
Eastern Hemisphere Distribution
Marcel Dekker AG, Hutgasse 4, Postfach 812, CH-4001 Basel, Switzerland
tel: 41-61-260-6300; fax: 41-61-260-6333
World Wide Web
http://www.dekker.com

The publisher offers discounts on this book when ordered in bulk quantities. For
more information, write to Special Sales/Professional Marketing at the headquarters
address above.

Copyright ß 2004 by Marcel Dekker, Inc. All Rights Reserved.


Neither this book nor any part may be reproduced or transmitted in any form or by
any means, electronic or mechanical, including photocopying, microfilming, and
recording, or by any information storage and retrieval system, without permission in
writing from the publisher.

Current printing (last digit):


10 9 8 7 6 5 4 3 2 1

PRINTED IN THE UNITED STATES OF AMERICA


Preface to the Second Edition

The growing interest in the advances described in this book have called for
this second edition of this book. In the past few years pressure-sensitive
products have reached a maturity that warrants a detailed and critical
examination of their science and technology. This is a vast domain, and I
have tried to cover some of its special aspects in separate works. The volume
Development and Manufacture of Pressure-Sensitive Products (Marcel
Dekker, 1998) describes the whole domain of self-adhesive products;
Pressure-Sensitive Formulation (VSP, Utrecht) gives a detailed discussion of
a special, practical segment of pressure sensitivity. However, Pressure-
Sensitive Adhesives Technology (Marcel Dekker, 1996), the first of these
books, constitutes the main step on the way to understanding adhesive-
based pressure-sensitive products.
In the past decade advances in contact physics and mechanics have
allowed us to correlate the macroscopic aspects of adhesive bonding
and debonding with the macromolecular basis of the viscoelastomers. The
most important elements of this progress are described in a separate section
of the revised book. Developments in the practical examination and quality
assurance of pressure-sensitive products required an enlarged discussion
and reformulation of Chapter 10, ‘‘Test Methods.’’ Environmental
considerations made necessary the discussion of recycling methods, and
biodegradability of raw materials, product components, and pressure-
sensitive products. Other scientific and industrial advances (i.e., new raw
materials and improved coating technology) are also included in the
second edition. Thus, after undergoing a major revision, the second edition
of this book remains a comprehensive and convenient up-to-date source of

iii
iv Preface to the Second Edition

information for users in industry and academia. So far as I am aware,


this is the first single-author book on general aspects of pressure-sensitive
adhesive technology, and it has been my pleasure to assist in its success.

István Benedek
Preface to the First Edition

Since their introduction half a century ago, pressure-sensitive adhesives have


been successfully applied in many fields. They are used in self-adhesive
labels, and tapes and protective films, as well as in dermal dosage systems
for pharmaceutical applications, the assembly of automotive parts, toys,
and electronic circuits and keyboards. They have experienced an astonishing
growth rate, and the installed manufacturing and converting capacity has
also sharply increased. A specific engineering technology for pressure-
sensitive adhesives, surprisingly a special science, appears to be lacking.
Very few books deal with intrinsic features of pressure-sensitive adhesives.
The application of pressure-sensitive adhesives requires a thorough
knowledge of basic rheological and viscoelastic phenomena. Adhesive and
polymer scientists, however, are not very often employed as industrial
managers or machine operators. Therefore the need arises to investigate and
summarize the most important features of pressure-sensitive adhesive
technology and to explain the phenomena scientifically. This book covers
all the fields of manufacturing, conversion, and application and end-uses of
pressure-sensitive adhesives.
The classical approach would be to compile a treatise based on the
work of various experts, theoreticians, chemists, and engineers, thereby
coming up with a book consisting of a series of papers with a common title
only. We have, however, chosen a different approach. Based on our
experience as engineers (in both scientific activity and industrial areas) and
using the available technical literature, we have addressed all aspects of
pressure-sensitive adhesives. We have included the scientific basis of
suitability for specific applications (i.e., chemical and physical, rheology),
the raw materials, the manufacture (formulation) of the adhesive and of the
labelstock (converting the adhesive). We have selected self-adhesive labels
as the most complex self-adhesive laminate; we mainly discuss labels, but,
v
vi Preface to the First Edition

whenever possible, a comparison with and extension to other applications is


included. In order to illustrate the different topics and issues discussed, we
have referred to a number of commercially available products. It should be
kept in mind that these products are only mentioned in order to clarify the
discussion and in no way does it constitute any judgment about inherent
performance characteristics or their suitability for specific applications or
end-uses.
It is not the aim of this book to establish or complete the science of
pressure-sensitive adhesives, nor does it constitute a series of recipes. Rather
it serves as a practical aid to converters and those involved in the design and
use of pressure-sensitive adhesives.

István Benedek
Luc J. Heymans
Contents

Preface to the Second Edition iii


Preface to the First Edition v

1. Introduction 1
References 3

2. Rheology of Pressure-Sensitive Adhesives 5


1 Rheology of Uncoated PSAs 5
1.1 Properties of PSAs 6
1.2 Influence of Viscoelastic Properties on the
Adhesive Properties of PSAs 11
1.3 Influence of Viscoelastic Properties on the
Converting Properties of PSAs and PSA Laminates 25
1.4 Influence of Viscoelastic Properties on
End-Use Properties of PSAs 26
1.5 Factors Influencing Viscoelastic Properties of PSAs 31
2 Rheology of PSA Solutions and Dispersions 40
2.1 Rheology of PSA Solutions 41
2.2 Rheology of PSA Dispersions 44
3 Rheology of the Pressure-Sensitive Laminate 52
3.1 Influence of the Liquid Components of
the Laminate 54
3.2 Influence of the Solid Components of
the Laminate 58
3.3 Influence of the Composite Structure 81
References 82

vii
viii Contents

3. Physical Basis for the Viscoelastic Behavior of


Pressure-Sensitive Adhesives 89
1 TheRole of the Tg in Characterizing PSAs 90
1.1 Values of Tg for Adhesives 91
1.2 Factors Influencing Tg 92
1.3 Adjustment of Tg 104
1.4 Correlation Between the Main Adhesive, End-Use,
and Converting Properties of PSAs and Tg 107
2 Role of the Modulus in Characterizing PSAs 110
2.1 Factors Influencing the Modulus 113
2.2 Adjustment of the Modulus 124
2.3 Modulus Values 126
3. Contact Physics 129
3.1 Contact Mechanics for Elastic Materials 129
3.2 Contact Mechanics for Viscoelastic Materials 130
3.3 Micromechanical Considerations 132
4 The Role of Other Physical Parameters in
Characterizing PSAs 137
References 138

4. Comparison of PSAs 147


1 Comparison of PSAs with Thermoplastics and Rubber 147
1.1 Cold Flow 148
1.2 Relaxation Phenomena 155
1.3 Mechanical Resistance 156
2 Comparison Between PSAs and Other Adhesives 158
References 159

5. Chemical Composition of PSAs 161


1 Raw Materials 162
1.1 Elastomers 163
1.2 Viscoelastomers 176
1.3 Viscous Components 187
1.4 Components for In-line Synthesis 189
1.5 Components for Special Pressure-Sensitive
Formulations 197
1.6 Other Components 198
1.7 Release Coatings 199
2 Factors Influencing the Chemical Composition 205
2.1 Synthesis 205
2.2 Formulation 210
Contents ix

2.3 Physical State of PSAs 211


2.4 End-Use 216
2.5 Coating Method 220
2.6 Solid State Components of the Laminate 221
2.7 Product Build-Up 222
2.8 Environment Related Considerations 223
References 224

6. Adhesive Performance Characteristics 235


1 Adhesion-Cohesion Balance 235
1.1 Tack 235
1.2 Peel Adhesion 261
1.3 Shear Resistance (Cohesion) 307
2 Influence of Adhesive Properties on Other
Characteristics of PSAs 320
2.1 Influence of Adhesive Properties
on the Converting Properties 320
2.2 Influence of Adhesive Properties on End-Use
Properties 320
2.3 Influence of Peel Adhesion 320
2.4 Influence of Shear 321
3 Comparison of PSAs on Different Chemical Bases 321
3.1 Rubber-Based Versus Acrylic-Based PSAs 321
3.2 Acrylics and Other Synthetic Polymer-Based
Elastomers 322
References 341

7. Converting Properties of PSAs 353


1 Convertability of the Adhesive 353
1.1 Convertability of Adhesive as a Function of
the Physical State 353
1.2 Convertability of Adhesive as a Function of
Adhesive Properties 361
1.3 Convertability of Adhesive as a Function of the
Solid State Components of the Laminate 361
1.4 Convertability of Adhesive as a Function of
Coating Technology 361
1.5 Convertability of Adhesive as a Function of
End-Use Properties 362
2 Converting Properties of the Laminate 362
2.1 Definition and Construction of the Pressure-Sensitive
Laminate 363
x Contents

2.2 Printability of the Laminate 391


References 419

8. Manufacture of Pressure-Sensitive Adhesives 425


1 Manufacture of PSA Raw Materials 425
1.1 Natural Raw Materials 426
1.2 Synthetic Raw Materials 427
2 Formulating PSAs 429
2.1 Adhesive Properties 430
2.2 Formulating Opportunities 432
2.3 Tackification 433
2.4 Rosin-Based Tackifiers 461
2.5 Hydrocarbon-Based Tackifiers 468
2.6 Special Tackifier Resins 476
2.7 Tackification with Plasticizers 477
2.8 Cohesion Regulation 480
2.9 Coating Properties 481
2.10 Converting Properties 484
2.11 End-Use Properties 489
2.12 Influence of Adhesive Technology 505
2.13 Technological Considerations 541
2.14 Comparison Between Solvent-Based,
Water-Based, and Hot-Melt PSAs 543
References 549

9. Manufacture of Pressure-Sensitive Labels 559


1 Coating Technology 560
2 Coating Machines 563
2.1 Adhesive Coating Machines 564
2.2 Coating Devices/Coating Systems 565
2.3 Choice of Coating Geometry 585
2.4 Other Coating Devices 591
3 Coating of Hot-Melt PSAs 594
3.1 Roll Coaters for Hot-Melt PSAs 594
3.2 Slot-Die Coating for Hot-Melt PSAs 595
4 Drying of the Coating 598
4.1 Adhesive Drying Tunnel 598
5 Environmental Considerations 602
6 Simultaneous Manufacture of PSAs and PSA Laminates 612
6.1 Radiation Curing of PSAs 612
6.2 Siliconizing Through Radiation 613
Contents xi

7 Manufacture of the Release Liner 616


7.1 Nature of the Release Liner 616
7.2 Coating Machines for Silicones 619
7.3 Technology for Solvent-Based Systems 619
7.4 Technology for Solventless Siliconizing 619
8 Rehumidification/Conditioning 621
References 621

10. Test Methods 629


1 Evaluation of the Liquid Adhesive 629
1.1 Hot-Melt PSAs 630
1.2 Solvent-Based PSAs 630
1.3 Water-Based PSAs 630
2 Evaluation of the Solid Adhesive 639
2.1 Test of Coating Weight 639
2.2 Other Properties 640
3 Laminate Properties 643
3.1 General Laminate Properties 644
3.2 Special Laminate Properties 687
References 700

Abbreviations and Acronyms 709


1 Compounds 709
2 Terms 711
Index 715
1
Introduction

Adhesives are nonmetallic materials [1] used to bond other materials, mainly
on their surfaces through adhesion and cohesion. Adhesion and cohesion
are phenomena which may be described thermodynamically, but actually
they cannot be measured precisely. It was shown [2] that the most important
bonding processes are bonding by adhesion and bonding with pressure-
sensitive adhesives (PSAs). For adhesives working through adhesion
phenomena the adhesive fluid is transformed after bonding (i.e., the build
up of the joint) into a solid. In the case of PSAs, the adhesive conserves
its fluid state after the bond building too. Thus its resistance to debonding
is moderate and the joint may be delaminated without destroying the
laminate components in most cases.
Pressure-sensitive adhesives have been in wide use since the late 19th
century, starting with medical tapes and dressings. The first U.S. patent
describing the use of a PSA—for a soft, adhering bandage—was issued
in 1846 [3]. Ninety years later Stanton Avery developed and introduced the
self-adhesive label [4]. Two major industries resulted from these innovations:
pressure-sensitive tapes and labels. Industrial tapes were introduced in
the 1920s and 1930s followed by self-adhesive labels in 1935. About ten
years after that, pressure-sensitive protective films were manufactured. The
history of PSAs was described by Villa [5]. First, solvent-based PSAs using
natural rubber were developed (19th century). In the 1940s hot-melt
adhesives were introduced. Pressure-sensitive adhesives are adhesives that
form films exhibiting permanent tack, and display an adhesion which
does not strongly depend on the substrate [6]. The term PSA has a very
precise technical definition and has been dealt with extensively in the
chemical literature [7,8]. However, as discussed in [9], the technical term
of PSA in different languages (e.g. pressure-sensitive adhesive, auto-
collants, Haftkleber, etc.) is not completely clear. The recent development
1
2 Chapter 1

of pressure-sensitive products without a coated pressure-sensitive adhesive


layer, makes the definition of this product group more difficult [9,10].
The build up and classification of pressure-sensitive products have been
discussed in detail in [9].
The function of PSAs is to ensure instantaneous adhesion upon
application of a light pressure. Most applications further require that
they can be easily removed from the surface to which they were
applied through a light pulling force. Thus PSAs are characterized by a
built-in capacity to achieve this instantaneous adhesion to a surface without
activation, such as a treatment with solvents or heat, and also by having
sufficient internal strength so that the adhesive material will not break
up before the bond between the adhesive material and the surface
ruptures. The bonding and the debonding of PSAs are energy-driven
phenomena. Pressure-sensitive adhesives must possess viscous properties in
order to flow and to be able to dissipate energy during the adhesive bonding
process. However, the adhesive must also be elastic (i.e., it must resist
the tendency to flow) and, in addition, store bond rupture energy in order
to provide good peel and tack performance. Pressure-sensitive adhesives
should possess typical viscoelastic properties that allow them to respond
properly to both a bonding and a debonding step. For satisfactory
performance in each of these steps the material must respond to a deforming
force in a prescribed manner.
Polymers employed as PSAs have to fulfill partially contradictory
requirements; they need to adhere to substrates, to display high shear
strength and peel adhesion, and not leave any residue on the substrate
upon debonding. In order to meet all these requirements, a compromise
is needed. When using PSAs there appears another difference from wet
adhesives, namely the adhesive does not change its physical state because
film forming is inherent to PSAs. Thus, PSAs used in self-adhesive laminates
are adhesives which, through their viscoelastic fluid state, can build up the
joint without the need to change this flow state during or after application.
On the other hand, their fluid state allows controlled debonding
giving a temporary character to the bond. Because of the fluid character
of the bonded adhesive, the amount of adhesive (i.e., the dimensions of
the adhesive layer) is limited; the joint works as a thin-layer laminate or
composite. Because of this special, thin-layer structure of the composite,
the solid state components of the laminate exert a strong influence on
the properties of the adhesive in the composite. Therefore, there exists a
difference between the measured properties of the pristine adhesive and of
the adhesive enclosed within the laminate.
Adhesives, in general, and PSAs, in particular, have to build up
a continuous, soft (fluid), and tacky (rubbery) layer. The latter will adhere
Introduction 3

to the substrate. On the other hand, the liquid adhesive layer of the PSAs
working in the bond has to offer a controlled bond resistance. This
special behavior requires materials exhibiting a viscoelastic character. The
properties which are essential in characterizing the nature of PSAs comprise:
tack, peel adhesion, and shear. The first measures the adhesive’s ability to
adhere quickly, the second its ability to resist removal through peeling, and
the third its ability to hold in position when shear forces are applied [11].
These properties will be discussed in more detail in Chap. 6, which
describes the adhesive properties of PSAs. In order to understand the
importance of these properties, it is absolutely necessary to answer the
following questions:
What does the viscoelastic character of a PSA comprise?
What is the material basis (main criteria) for the viscoelastic behavior
of a PSA?

REFERENCES

1. DIN 16921.
2. R. Köhler, Adhäsion, (3) 90 (1970).
3. J.A. Fries, ‘‘New Developments in PSA,’’ in TECH 72, Advances in Pressure-
Sensitive Tape Technology, Technical Seminar Proceedings, Itasca, IL, May,
1989.
4. Der Siebdruck, (3) 69 (1986).
5. G.J. Villa, Adhäsion, (10) 284 (1977).
6. Vinnapas, Eigenschaften und Anwendung, 7.1. Teil, Anwendung, Wacker
GmbH, München, 1976.
7. R. Houwink and G. Salomon, Adhesion and Adhesives, Vol. 2, Chapter 17,
Elsevier Co., New York, 1982.
8. D. Satas, Handbook of Pressure Sensitive Technology, Van Nostrand
Rheinhold Co., New York, 1982.
9. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999.
10. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000.
11. J.P. Keally and R.E. Zenk. (Minnesota Mining and Manuf. Co., USA), Canad.
Pat. 1224.678/10.07.92 (US Pat. 399350).
2
Rheology of Pressure-Sensitive
Adhesives

Pressure-sensitive adhesives are viscoelastic materials with flow properties


playing a key role in the bond forming; their elasticity plays a key role in
the storage of energy (i.e., the debonding process). The balance of these
properties governs their time-dependent repositionability and bonding
strength (i.e., their removability). Their flow properties are useful in coating
technology and at the same time detrimental to the converting technology of
labels.
Generally PSAs are used as thin layers, therefore their flow is limited
by the physico-mechanical interactions with the solid components of the
laminate (liner and face) materials. On the other hand the solid components
of the laminate are generally thin, soft, viscous, and/or elastic layers,
allowing a relatively broad and uniform distribution of the applied stresses.
Thus the properties of the bonded adhesive (i.e., its flow characteristics) may
differ from those of the pure (unbonded) adhesive. Therefore in this chapter
the rheology of pure and coated PSAs will be dealt with separately.

1 RHEOLOGY OF UNCOATED PSAs

It remains difficult to examine the properties of pure (i.e., uncoated or


unbonded) PSAs, and to obtain generally valid information. Pressure-
sensitive adhesives are seldom used as thick layers between motionless rigid
surfaces (i.e., as fluids). On the other hand, as known from industrial
experience, the nature of the face stock material or of the substrate used,
and their characteristics and dimensions may significantly influence the
properties of the PSA laminate. Practically, this disadvantage is eliminated
by the use of normalized or standard solid state components. However, a
theoretical approach may be used for the investigation of pristine PSAs.

5
6 Chapter 2

1.1 Properties of PSAs


The adhesive and end-use properties of PSAs require a viscoelastic, non-
Newtonian flow behavior which is based on the macromolecular nature of
the adhesive. In order to understand the needs and means of viscoelastic
behavior one needs to summarize the most important material properties
specifically related to PSAs. Generally, adhesives in a bond behave like a
fluid or a solid. Fluids are characterized by their viscosity which influences
their mobility, whereas solids are characterized by their modulus which
determines their deformability. In an ideal case, for Newtonian fluids (or for
solids obeying Hooke’s law) the applied force (load) will be balanced by
the material’s own mechanical characteristics, that is, the viscosity  or the
Young’s modulus E:

 ¼ = 0 ð2:1Þ
E ¼ = ð2:2Þ

where  and  are the applied stresses, and  and  0 are the strain and shear
rate, respectively.
As indicated earlier, PSAs originate from a film-forming, elastomeric
material, which combines a high degree of tack with an ability to quickly wet
the surface to which it is applied, to provide instant bonding at low-to-
moderate pressure as a result of its flow characteristics. On the other hand,
PSAs possess sufficient cohesion and elasticity, so that despite their
aggressive tackiness they can be handled with the fingers and removed
from smooth surfaces without leaving any residue. Moreover, in order to
achieve bond strength they have to store energy (i.e., they must be elastic).
Fundamentally PSAs require a delicate balance between their viscous and
elastic properties. One should note that PSAs have to satisfy these
contradictory requirements under different stress rate conditions, that is,
at low shear rates they must flow (bonding) and at high peeling rates they
have to respond elastically (debonding).
Consequently, according to their adhesive and end-use properties,
PSAs cannot be Newtonian systems: they do not obey Newton’s law (i.e.,
there is no linear dependence between their viscosity and the shear rate).
Their viscosity is not a material constant, but depends on the stress value or
shear rate:

ðÞ ¼ = 0 ð2:3Þ

That is:

 ¼ a   n ð2:4Þ
Rheology of Pressure-Sensitive Adhesives 7

where a is the apparent viscosity, and n denotes the flow index [1]. For
Newtonian systems the exponent n is one, which implies that the viscosity
does not depend on the shear rate. As pointed out, the viscosity of PSAs
does depend on the shear rate. This is possibly due to their macromolecular
character. Pressure-sensitive adhesives are polymers containing long-chain
entangled molecules with intra- and intermolecular mobility. At low strain
rates, the viscous components of the polymer dissipate energy, and as a
result resistance to debonding forces is low. At higher strain rates, the
molecules have less time to disentangle, and to slide past one another; in
this case viscous flow is reduced, but the elastic modulus or stiffness of the
polymer increases [2]. This behavior results in additional stored energy, and
the debonding resistance intensifies accordingly.
Practically, the dependence of the adhesive performance character-
istics on the stress rate may be observed by peeling off removable PSAs at
different peel rates: at higher rates paper tear may occur. The stress rate-
dependent stiffening is an increase in the elastic contribution to the rheology
of the polymer. When the elastic components are predominant more of the
bond rupture energy is stored, resulting in higher peel and tack properties.
The end-use properties of PSAs result from the nonlinear viscoelastic
behavior of the adhesive material, and the elastomeric polymer basis of
PSAs imparts them such a viscoelastic behavior. It is evident that the same
stiffening effect is apparent when the polymer temperature decreases. In this
case the polymer molecules are again restricted in their ability to flow, and
the modulus increases. Consequently the adhesive properties of PSAs are
also temperature dependent. Thus one always has to take into account
that the viscoelastic properties of PSAs are strain-rate and temperature
dependent. Zosel [3] demonstrated that the separation or debonding
energy of the adhesive joint is a function of the thermodynamical work of
adhesion and of a temperature and rate-dependent function (depending
on the viscoelastic properties). Accordingly, PSAs would absorb less or
more energy depending on the rate (frequency !) of the applied stress.
Practically, end-use situations with different stress rates may be simulated
experimentally by applying a strain to a thin sample of the material and
measuring the output stress. If the material is an ideal solid, its response is
completely in phase with the applied strain. A viscoelastic fluid, such as a
PSA, displays a mixture of solid-like and liquid-like responses. Therefore the
output stress curve is deconvoluted into an in-phase part (related to energy
storage) and an out-of-phase part (related to energy loss). The coefficients of
the in-phase and out-of-phase parts are called the energy storage modulus
and the energy loss modulus [4].
According to the theory of Lodge [5] the rheological state of a viscous
liquid subject to a sinusoidal deformation will be described by the following
8 Chapter 2

equation:

12 ¼ ð!Þ  0  ! cos ! t ð2:5Þ

The response of the elastic solid may be described as follows:

12 ¼ Gð!Þ  0  sin ! t ð2:6Þ

where G is the shear modulus. For the reversible and irreversible work of
deformation one can write:

12 ¼ Gð!Þ  0  sin ! t þ rev 0 rev  ! cos ! t ð2:7Þ


12 ¼ irr  0 irr  ! cos ! t ð2:8Þ

where  denotes the stress, ! the angular speed,  the viscosity, and  0 the
amplitude of the deformation.
The storage modulus G0 increases with the frequency:

G0 ¼ ðrev  ! þ GÞ cos  ð2:9Þ

whereas the viscosity decreases with the frequency;  is the loss angle.
Pressure-sensitive adhesives must display irreversible work of defor-
mation during bonding and reversible deformation work upon debonding.
The ratio of both kinds of deformation work (i.e., of stored and dissipated
energy) characterizes the behavior of PSAs. In general the energy state of the
viscoelastic polymer may be described as follows:

ðtÞ  ½G0 ð!Þ cos ! t þ G00 ð!Þ sin ! t ð2:10Þ

where G0 is the storage modulus and G00 the loss modulus, and

loss tan  ¼ loss modulus=storage modulus ¼ G00 =G0 ð2:11Þ

Tan  is a damping term and is a measure of the ratio of energy


dissipated as heat, to the maximum of energy stored in the material. One can
suppose that the term ‘‘loss tan ,’’ as an index of the amount of stored or
lost energy (i.e., of the contribution of the elastic and viscous part of the
polymer) may also characterize the adhesive properties. It was shown that
loop, peel, and quick stick show a good correlation with loss tan  [4]. It was
demonstrated that PSAs intended for similar application also exhibit similar
Rheology of Pressure-Sensitive Adhesives 9

rheological properties. The correlation between adhesive properties and the


dynamic shear storage modulus appears quite good [6]: hence the concept
of ‘‘window of performance’’ as a function of the storage modulus of the
adhesive was developed [7]. These moduli, the storage and the loss moduli,
can be displayed as a function of the temperature (Fig. 2.1).
The storage modulus starts high at low temperatures where all motion
within the polymer is frozen and the material behaves like a glass. At higher
temperatures it drops off and exhibits a plateau region which represents the
elastomeric response generally encountered at normal end-use temperatures;
the storage modulus then decreases further when softening begins. The tem-
perature region through which the polymer changes from a glassy (hard)
state into a liquid (rubber-like) state, this second order transition point (with
a continuous differential of the free enthalpy, but discontinuous, second
order differential of the Gibb’s free energy) is called the glass transition
temperature (Tg), and has a special significance in the characterization of
PSAs. Above the Tg the time-temperature superposition principle can be
applied. Differences in viscoelastic parameters around the glass transition
can be directly related to the side chain size and mobility of the polymer.

Figure 2.1 Dependence of the modulus on the temperature. 1) Storage modulus;


2) loss modulus; 3) tan .
10 Chapter 2

At Tg there occurs a change in the thermodynamic state which can be related


to a mechanical energy loss function such as the loss modulus. The loss tan 
peak does not occur at the glass transition but in the transition zone between
the glassy and rubbery regions. Ferry [8] pointed out that in this region of
transition there is no change in the thermodynamic state. The loss tan  peak
is the midpoint of this transition zone where the ratio of loss modulus and
storage modulus reaches a maximum. The energy loss maximum at this
point has considerable influence on the tack of the system. The storage
modulus definition can be simplified as a hardness parameter [9]. Optimum
PSA performance can be quantified using the storage modulus; ideally the
value of the storage modulus should vary between 20 and 80 kPa.
Chu [7] correlated PSA performance and dynamic mechanical
performance (DMA) properties, whereby PSA performance, especially for
tapes, was related to the storage modulus G0 at room temperature and the
loss tan  peak temperature of the system. Chu also showed that the PSA
application window for a high cohesive strength tape adhesive requires G0
values at room temperature between 50 kPa and 200 kPa with loss tan 
peak temperature limits between 10 and þ10 C. Optimum G0 values for
permanent PSA labels were determined to be around 20 kPa at room
temperature.
Pressure-sensitive adhesives were defined using viscoelastic application
windows relating the storage modulus G at room temperature to the loss tan
 peak of the adhesive. In water-based adhesives the viscoelastic relationship
is not as simple, that is, it was determined that for the most commonly used
polymers in water-based dispersions these relationships may not apply. In
hot-melt and solvent-based PSAs a close and predictable relationship exists
between the loss tan  peak, defined as the dynamic Tg, and the Tg as
measured by differential scanning calorimetry (DSC). For water-based
adhesives the relationship varies depending upon the polymer type used.
The loss tan  peak temperature of an acrylic can differ from the Tg (DSC)
by as much as 30 C. The phenomenon is much less pronounced for styrene-
butadiene rubber (SBR)-type polymer dispersions. This variation is also
valid for an adhesive dispersion containing a tackifier. The consequence of
this is that the loss tan  peak temperature cannot be used to predict and
define PSAs performance in a viscoelastic application window. According to
Bamborough [9] and applying the rule proposed by Chu, it may appear that
an SBR would require considerably more compatible resin than an acrylic
polymer; a soft acrylic (AC) PSA (like Acronal V 205) would require a
higher amount of compatible resin than a hard one (like Acronal 80 D).
Adhesive formulators know this not to be the case. Experienced adhesive
formulators know that for Acronal V 205 an optimum concentration of
tackifier would be 30 parts per 100 parts polymer whereas for Acronal 80 D
Rheology of Pressure-Sensitive Adhesives 11

one needs 80 parts (see Chap. 6). The loss tan  peak temperature for water-
based adhesive systems is not a reliable predictor of PSA performance
characteristics.
It is proposed that the loss modulus peak temperature of the adhesive
should be 50 C below the operating temperature of the adhesive.
Bamborough [9] proposed using the loss modulus peak temperature in the
glass transition region instead of the loss tan  peak temperature as a means
of predicting PSA performance for water-based adhesives. Despite the
above-illustrated discrepancies concerning DMA for adhesive characteriza-
tion, the use of dynamic mechanical spectroscopy to measure modulus
changes, and differential scanning calorimetry to measure shifts in the Tg of
the adhesive are now common methods for rheological studies of adhesives
[10]. In order to understand the practical benefits of such investigations the
rheology of PSAs needs to be studied.
As discussed above PSAs are special products allowing instantaneous
bonding due to their liquid-like flow and solid-like debonding resistance
due to their elasticity. Such viscoelastic behavior can be characterized
rheologically by the main parameters of a liquid (viscosity) and a solid
(modulus) taking into account their dependence on the time and
temperature (see storage and loss modulus ratio). It is evident that the
temperature domain allowing such viscoelastomer-like behavior has to be
taken into account also. Therefore as a supplemental rheological parameter
the value of the Tg should be used also. Extending the use of the William-
Landel-Ferry time-temperature superposition principle allowed an easier
rheological characterization of PSAs. It is known that the viscoelastic
behavior of amorphous polymers is a function of the time-temperature
dependence, the dependence of the ratio between recoverable energy during
a given deformation and energy losses on the experimental conditions,
characterized by the stress rate and temperature, that is of the validity of
the time-temperature superposition principle. This principle states that the
viscoelastic properties at different temperatures can be superposed by a shift
of the isotherm data along the logarithmic time-frequency scale. As
discussed in [11], by ‘‘replacing’’ the time with the temperature it was
possible to develop ‘‘full plastic’’ PSPs, i.e., polymer films having built-in
pressure sensitivity.

1.2 Influence of Viscoelastic Properties on the Adhesive


Properties of PSAs
The essential performance characteristics when characterizing the nature
of PSAs are tack, peel adhesion, and resistance to shear. The first property
represents the adhesive’s ability to adhere quickly (initial grab), the second
12 Chapter 2

measures its ability to resist removal by peeling, and the third characterizes
the adhesive’s ability to resist flow when shearing forces are exerted.
Generally speaking the first two characteristics are directly related to each
other, but inversely related to the third one [12].
Tack and peel tests imply a high loading rate, whereas shear will be
mostly measured in a static manner. The first phase in the use of PSAs
(bonding) generally occurs slowly, whereas the second step (debonding
during converting or end-use) imposes higher stress rates. The balance of
the adhesive properties, and of the adhesive/converting/end-use properties
reflects at the same time the need for the balance of the viscoelastic
characteristics. In this chapter the influence of the viscoelastic properties on
the adhesive properties will be briefly examined.

Influence of Viscoelastic Properties on the Tack of PSAs


According to Rivlin [13] the separation energy after a short contact time and
low pressure is a measure of the tack. A short contact time and low pressure
during application of PSAs imply a high wetting ability. For bonding to
occur there is an a priori need for wetting of the substrate. As confirmed by
Sheriff et al. [14] and by Counsell and Whitehouse [15] tack is a function of
wetting. Good wetting supposes sufficient fluidity of the adhesive, and
fluidity is characterized by viscosity.
Tack Dependence on the Viscosity. According to Zosel [16] tack is
measured in two steps, namely the contact step and the separation step.
During the first step, contact is made in the geometrical surface points,
which increase to a larger area through wetting out, viscous flow, and elastic
deformation. Wetting out implies high fluidity, as characterized by an
adequate viscosity of the adhesive. Wetting out (i.e., covering the surface
by the fluid adhesive) is followed by bonding due to the viscoelastic
deformation of PSAs. On the other hand, debonding assumes the defor-
mation of the laminate, the creation of two new surfaces, and deformation
of the new surfaces. Thus, it may be concluded that for bond-forming a high
deformation with a medium elasticity is required, whereas for debonding a
medium deformation with a high elasticity are required.
The tack may also be characterized as separation energy [17,18].
During debonding high tack means that the adhesive absorbs a high amount
of deformation energy, which dissipates on the break of the bond [19]:
a high ability to store energy implies elasticity, a high energy at break means
high cohesion. Thus, tack depends on elasticity and cohesion. Therefore,
for high tack, a low bonding viscosity, a high debonding viscosity, and high
elasticity are required. Factors influencing the viscosity and the elasticity of
the polymer will also influence the tack. The polymer’s own characteristics
Rheology of Pressure-Sensitive Adhesives 13

and the environmental conditions (experimental parameters) influence its


viscous and elastic behavior. Studying the performance of carboxylated
styrene-butadiene rubber (CSBR) latexes, Midgley shows that tack depends
on the Mooney viscosity (i.e., on the molecular weight, MW) [20].

Tack Dependence on the Modulus of Elasticity. Loss moduli correlate


with PSA debonding tests; McElrath [21] studied the debonding frequency
of PSA tests and their location on the loss modulus master curve. A good
agreement between adhesive properties and loss modulus was demonstrated.
Although absolute correlations have not been established, Class and Chu
[22] suggested that the location and the shape of the modulus curve in the
transition zone is important for PSA performance. In accordance with
Dahlquist’s criterion for a minimum value of compressive creep compliance
to achieve tack, Class and Chu said their data indicated a maximum
modulus value. Dahlquist [23, 24] related tack to modulus, showing that the
compression modulus should not be much higher than 105 Pa. Very high
modulus adhesives do not possess sufficient conformability to exhibit
pressure-sensitive tack. The optimum tack properties of PSAs are obtained
when the room temperature modulus falls within the range of  5  105 to
1  105 Pa, and the Tg lies between 10 to þ10 C [25].
Hamed and Hsieh [26] showed that for a given total bonded area, test
specimens containing noncontact regions of sufficiently small scale can
exhibit a higher peeling force than those in which the noncontact ones are
large. For an elastomer there is a critical flaw size below which adhesive
strength will remain unchanged. Rubbery materials with higher elasticity
have a smaller critical flaw size. Thus the modulus influences the tack.
Rubbery materials with a more elastic response exhibit tack properties that
are more sensitive to interfacial flaws, compared to those that respond more
by viscous flow. When failure occurs by viscous flow, the tack is relatively
insensitive to interfacial flaws, but when the strain rate is sufficiently high, so
that the elastomer responds elastically, stresses may be concentrated at the
edges of interfacial flaws causing a reduction of the strength.

Tack Dependence on Experimental Parameters. Viscosity and elastic


moduli of PSAs are not intrinsic material characteristics (i.e., they depend
on the experimental parameters used, such as the temperature and time, and
the strain rate). Thus, a similar dependence of the tack on time, temperature,
and the strain rate has to be taken into account. This dependence is
illustrated by the quite different values of the tack obtained using different
experimental techniques (rolling ball, quick stick, or loop tack) which are
characterized by different time and strain rates, and by the sensitivity of the
adhesive properties to environmental conditions (see Chap. 10). Different
14 Chapter 2

tack methods are adequate for different PSPs [27]. With the probe tack test
the load is regulated, as in the lamination of protective films. Polyken tack is
less strongly affected by the resin softening point (it is applied under load!)
during tackifying. Loop tack simulates the actual application conditions of
labels, which are blown onto a surface by using air pressure. Rolling ball
tack is more complex, it is friction related.
Tack Dependence on Temperature. Most general-purpose adhesives
are formulated to have tack at room temperature. If the adherent tempera-
ture is lower than room temperature, a higher degree of adhesive cold flow is
required to provide proper wet-out. In reality it is very difficult to apply PSA
labels at low temperature conditions. Deep freeze label (or tape) adhesives
must be specially formulated, with a low viscosity at low temperatures [28].
This behavior is due to the limited flow of the adhesive at lower tem-
peratures, a phenomenon governed by the strong temperature dependence
of the viscosity and of the elastic modulus. Deep freeze labels should
display the same viscoelastic properties at 40 C as at þ20 C [29]. It is
recommended that such adhesives have a lower storage modulus value than
common labels (104 Pa). As shown by Hamed and Hsieh [26] tack is a
function of test temperature and strain rate, and the experimental data may
be shifted from a master curve. However, the tack behavior as a function of
the temperature and strain rate is more complex than that of the cohesion.
Tack Dependence on Strain Rate. It was shown earlier that tack
depends on the bonding and debonding process, on debonding (separation)
work, and thus tack also depends on the strain rate. This phenomenon may
be observed by testing the tack using methods measuring the debonding
resistance (force), like loop tack or Polyken tack. Tack varies as the
separation speed of the Polyken test is changed [4]. The dependence of the
tack value on the increasing speed was confirmed by Hamed and Hsieh [26]
too. They observed a nonlinear, discontinuous increase of the tack with the
measurement speed. Tack values rise to a first maximum and then, after a
period of slight decrease, rise continuously with the measurement speed.
Data obtained from Polyken tack measurements show a correlation between
tack and loss tan  peak values [4]. Tack dependence on the debonding rate
was also confirmed by McElrath [21]. He demonstrated that loss modulus
values depend on the applied frequency, and different tack test methods
(loop, quick stick, and probe tack) exhibit maxima at different frequencies.

Influence of Viscoelastic Properties on the Peel of PSAs


Peel and peel strength are measured by separating an adhesive applied to a
substrate at some angle with respect to the substrate, usually at an angle
Rheology of Pressure-Sensitive Adhesives 15

of 90 or 180 [30]. Similar to tack, the measurement of the peel adhesion
involves a bonding step before the debonding or peeling step.
Tack measures the resistance to separation of the adhesive after a
short contact time, or by light pressure. Peel is measured after a relatively
long or very long contact time (at least 102 longer contact time than in the
tack test) after application of a light or medium pressure. The time available
for bond forming (wetting and penetration of the surface) during the first
contact step is longer for peel measurements than for tack. It follows that
flow properties of the adhesive during the bonding step are less critical than
for the tack measurement. On the other hand, the debonding resistance
will depend on the viscous nature/elasticity balance requiring its precise
adjustment in order to achieve peelability (removability or repositionability)
and on the strain rate, which influences the separation resistance in a more
pronounced manner than during the measurement of the tack.
Peel Dependence on Viscosity. Like tack, peel implies a bonding and
debonding step, with the time for the latter lasting longer. Supposedly the
influence of the viscosity on the bonding step during a peel measurement
is less important than for the tack. On the other hand, the debonding
resistance of the joint is increasingly proportional to the viscous flow of the
adhesive (i.e., a high peel needs a solid-like adhesive). In formulation
practice the regulation of the peel by (bonding) viscosity is used for special,
rubber-based, low peel products (e.g., tapes and protective films), where
mechano-chemical destruction of the base elastomer (mastication) is carried
out [31, 32] in order to manufacture a soft adhesive. The value of the peel is
also a criterion for the distinction between removable and permanent labels.
The peel dependence on the viscosity (modulus) and its theoretical basis will
be discussed in more detail in Section 1.4.
Peel Dependence on the Elastic Modulus. Special PSAs intended for
medical and surgical applications should display a low value of the modulus
(and a high value of the creep compliance) in order to allow removability
[33]: the higher the creep compliance, the greater the adhesive residue left on
the substrate. Creep compliance values greater than 2.3  10 Pa1 are not
preferred.
The peel force is related to the storage modulus G0 of the adhesive.
High tack, removable adhesives should have a low G0 , and the storage
modulus should not vary much with the peel rate. On the other hand, the
debonding takes place in a much higher frequency range than the bonding
process. In order to maximize the peel force, the highest possible G0 value in
the high frequency range is needed. Satas [6] showed, that solution polymers
with a higher G0 slope at higher frequencies than emulsion polymers, exhibit
higher peel adhesion, as is generally the case with acrylic solution polymers.
16 Chapter 2

Changing the chemical composition (e.g., sequence distribution) may lead to


a change (increase) in modulus and eventually to a decrease of the peel force
[34]. Regulating the diblock/triblock ratio in segregated copolymers may
reduce/increase the plateau modulus, i.e., soften or harden the polymer, and
influence the adhesive and converting properties [35].

Peel Dependence on Dwell Time and Strain Rate. As known from


plastic film processing and use, adhesion and self-adhesion depend on the
contact time. Self-adhesive (i.e., adhesiveless) protective films may build up
a 100% increase of their adhesion as a function of the end-use conditions
[36]. The build-up of the adhesion with the time is a general phenomenon
due to the macromolecular nature and viscoelasticity of such polymer films.
In a similar manner contact forming or bonding of the adhesive assumes its
viscoelastic deformation. Viscous, slow flow is time dependent and thus
bond forming will also be time dependent. According to [36] peel build-up as
a function of the time is a result of contact build-up in time. Full contact
build-up is referred as self-healing and in general the self-healing time (theal)
depends on the viscoelastic properties of the material, the aspect ratio, the
spacing of the asperities, and the surface properties. For certain application
conditions the time for self-healing (i.e., no external loading) is given by:

theal ¼ Að, mÞLCo o2 ½Co =C1 1=m =W ð2:12Þ

where A(, m) is a dimensionless constant depending on m, the creep


exponent, and a dimensionless parameter:
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffi
A ¼ ð2 2=Þ ð L=hÞ ð Co W Þ ð2:13Þ

where W/Co o2 is a characteristic length of the local bonding process and


[C1/Co]1/m is the characteristic relaxation time of the material. Eq. (2.12)
implies that the healing is very sensitive to the cohesive stress. Because of
this materials that have identical creep behavior and work of adhesion can
have very different healing times. The healing also depends on the height of
the asperities through the parameter . This parameter has the following
interpretation: if the viscoelastic polymer is replaced with an elastomer with
modulus 1/Co then the two surfaces will jump in contact with each other
with no externally applied load and the contact area will continue to
increase until the surfaces are in full contact if  was larger than some
critical value. This critical value is typically of order 1. The formulation of
this work implicitly assumed that  is below the critical value. For typical
values of h, L, and W, the elastic modulus which would cause the surfaces to
Rheology of Pressure-Sensitive Adhesives 17

jump in contact and a self-heal is of order 0.1 MPa. Therefore this work is
more relevant for viscoelastic adhesives or uncrosslinked rubbers of higher
modulus.
The time dependence of the bonding step is illustrated by the influence
of the dwell time on the peel values. As discussed in Chap. 10, application
pressure and temperature influence the peel and its build-up too. Therefore,
as proposed in [37] for special PSPs static and dynamic dwell time should
be tested.
Like the pressure used during application, the contact time (i.e., the
interval between bonding and debonding) takes into account the relatively
low mobility (flow rate) of PSAs. A higher pressure or a longer dwell time
should help adhesive flow and increase the peel resistance. It is known from
PSA characterization, that for the FTM-4 high speed release test the
samples are placed between two metal plates under a pressure of 6.87 kPa to
ensure good contact. In a similar manner for PSPs with very low tack and
instantaneous peel (e.g., protective films) bonding depends on the applied
pressure and temperature also. A special class of such products (warm
laminating films) is used at increased laminating temperature and pressure
[38]. Foam-like transfer tapes are also used under pressure to increase initial
bond strength [39]. As discussed in [40] time-dependent peel control is one of
the main requirements for PSA. Generally, until equilibrium is reached, the
peel resistance increases with increasing contact pressure and time [26,41]
(Table 2.1; Fig. 2.2).
As can be seen from Fig. 2.2, the peel force increases with the dwell
time of the adhesive, that is, the viscous and elastic deformation of the
adhesive need a period of time (depending on the viscosity and experimental
conditions). As will be discussed later, the time dependence of the bonding
imposes the use of normalized dwell times for peel measurement purposes.
Generally the peel/dwell time dependence is not linear [42], but partially
crosslinked silicone rubber exhibits peel values which increase linearly with

Table 2.1 Peel Adhesion as a Function of Increasing Dwell Times

Peel Adhesion (N/25 mm)

Dwell Time (week) Sample 1 Sample 2 Sample 3

0 0.32 0.12 0.09


1 0.44 0.24 0.21
2 0.47 0.30 0.30
3 0.55 0.35 0.35
4 0.60 0.40 0.47
18 Chapter 2

Figure 2.2 Dependence of the peel values on the dwell time. Peel from
polyethylene as a function of the coating weight for different dwell time values.
Dwell time of 1) 20 min; 2) 0 min.

the dwell time [43]. The adhesion build-up over time for the main types
of PSPs (e.g., permanent, repositionable, and removable adhesives) was
discussed in detail in [44]. It has been demonstrated that peel build-up
depends on the adherend surface quality [37]. The build-up of the peel force
in time has a special importance in the design of removable adhesives.
The dependence of the separation energy on the contact time was also
demonstrated by Zosel [3]. The ‘‘memory effect’’ in the adhesion of
rubber to rigid substrates is well known [45]. After readhering the adhesive,
the peel force is lower. After initial peeling of the adhesive, it regains its
original peeling resistance only after a period of time (or much faster with
the help of some solvent). The memory as a function of the dwell time
is associated with some rearrangements of the molecular structure of
the rubber at the interface. In a special case (e.g., diaper closure tapes) the
removable adhesive allows reliable closure and refastening. In this case the
time dependence of the flow/deformation of the adhesive and its influence
on the peel may be observed during the debonding process. Such tapes have
to exhibit a maximum peel force at a peel rate between 10 and 400 cm/min
Rheology of Pressure-Sensitive Adhesives 19

and a log peel rate between 1.0 and 2.6 cm/min [46]. Another influence of the
dwell time may be observed, when examining the influence of the storage
and aging on the peel adhesion value. In general there is a build-up of the
peel resistance with time.
It should be emphasized that stress transfer during debonding occurs
by means of the solid components of the laminate. Therefore strain rate
depends upon the carrier and substrate. The denaturation of the peel values
by carrier deformation (i.e., strain rate) is discussed in detail in [47].
Influence of the Peeling Rate. At low peel rates the viscous flow, and
the deformation of the adhesive layer are dominant for the peel resistance.
Therefore the peel resistance increases with increasing peeling rate. At high
peeling rates the elastic character of the adhesive dominates, thus in this
region the peel resistance no longer depends on the peeling rate. Of practical
importance is the very pronounced dependence of the peel force on the
peeling rate (Fig. 2.3); the peel force increases with the peel rate [48]. This
behavior requires the use of normalized peeling rates for the measurement of
the peel adhesion force.

Figure 2.3 Dependence of the peel force on the peeling rate. 1 through 8 are
different tackified water-based acrylic formulations; the peel adhesion from glass was
measured.
20 Chapter 2

The dependence of the peel on the peeling rate may also be observed
for very low peel forces (peeling from release liner) [49]. McElrath [21]
showed that (as theoretically supposed) the loss modulus depends on the
frequency, as does peel adhesion, and displays a maximum for a given
frequency value. Removable adhesives having a plateau region in the
storage modulus/frequency plot, exhibit a peel force independent of the peel
rate [6]. Kendall et al. [25] studied the dependence of the peel on the Tg; they
also stated that the peel maxima change with the testing rate. Higher test
rates or lower ambient temperatures produce maxima at lower resin levels,
lower resin softening points, or lower elastomer Tg. The peel of different
water-based PSAs displays large differences after storage at 100 C. As
shown in [25] adhesive break (in the adhesive layer) appears above a certain
peeling rate (2.5 mm/min), or below a certain temperature (25 C) only.
It may be possible to remove a label from a paper substrate if it is
peeled off very slowly, but the same label will certainly tear if it is peeled
off quickly. Since the peel force is related to the storage modulus of the
adhesive, high-tack removable adhesives should possess a low storage
modulus which does not vary much with frequency (rate of peel). At any
given temperature peel adhesion is observed to increase as the peel rate is
increased [2]. At low strain rates the peel forces are much lower. Under these
conditions, the viscous components of the polymer dissipate significant
amounts of energy and, as a result, resistance to peel forces is low. At higher
strain rates the molecules have less time to disentangle and to slide past
one another. This behavior results in more stored energy, and peel forces
intensify accordingly. One can conclude that at least theoretically, each
adhesive may be considered as a low peel adhesion adhesive (see Chap. 6)
provided a very low peel rate is applied.

Temperature Dependence of the Peel. As is known, a maximum peel


strength implies a certain modulus value and viscosity. Both modulus and
viscosity depend on the temperature. With the increase of the temperature
the viscosity of PSAs decreases. Therefore the increase of the temperature
improves the tack and instantaneous peel, and exerts a negative influence on
the cohesion. Such an influence is illustrated in the end-use of several PSPs.
It was shown for ethylene acrylic acid copolymers that peel adhesion
depends on the temperature [30]. The peel of an untackified PSA at 0 C is
about 210% lower than the peel value at 23 C. For tackified formulations
peel reduction at 0 C may attain 300% [50]. The importance of laminating
temperature of acrylic PSA coated protective films on plastic plates was
demonstrated in [51]. The dependence of the peel value and of the break
nature on the peel rate is a common phenomenon observed in the
delamination of PSPs during their application also. Such dependence
Rheology of Pressure-Sensitive Adhesives 21

causes the so-called inversion of the adhesive break when tapes are unwound
at too low a temperature (lower than 10 C) or with too high running speed.
It is well known in packaging applications that the winding resistance for
tapes depends on the temperature and the winding rate (i.e., the peel is
temperature dependent). As shown in [52] the unwinding resistance (Rw) is a
function of temperature (Tw) and unwinding speed (vw).

Rw ¼ f ðTw , vw Þ ð2:14Þ

Generally the unwinding force depends on a number of parameters


such as the adhesion force, the moduli of elasticity of the adhesive and
carrier film, the thicknesses of the adhesive and film, and the width of the
tape. The modulus of the elasticity of the adhesive depends on the time and
temperature, i.e., on the unwinding speed and temperature.

Influence of Viscoelastic Properties on the Shear of PSAs


Cohesive strength is measured as shear or shear strength, which is the
resistance of adhesive joints to shear stress, and is measured as a force per
unit area at failure. The shear force is applied in a plane parallel to the
adhesive joint.
Different authors have formulated a definition of the cohesive
strength. If deformation by shearing is considered like a flow, the flow
limit (FL) is given by the following correlation [53]:

FL ¼ ðE  We  b2 Þ=4h3 ð2:15Þ

where We denotes the maximum elastic deformation of the sample, b the


width of the adhesive surface in the stress direction, and h the thickness of
the adhesive layer. It can be seen from the above relation that the flow limit
(i.e., the cohesion) is a function of the elastic modulus of the adhesive.
Considering that the strength of a PSA joint depends on the viscosity
, the adhesive layer thickness h, and time t, the interdependence of these
factors may be formulated as follows [54]:

F ¼ =ðh2  tÞ ð2:16Þ

The influence of the viscoelastic properties on the adhesive character-


istics of PSAs depends on the nature of the stresses applied during tack, peel,
or cohesion measurements. Köhler [55] demonstrated that tensile or shear
stresses produce different levels of deformation in PSAs. When applying a
22 Chapter 2

tensile force on a circular, laminated PSA layer, with a viscosity of 103 Pa,
according to the relation:

3    R4
P¼ ð2:17Þ
4 t h2

where P is the applied force, R is the sample radius, t is the time, and h is the
thickness of the adhesive layer, one needs a force of 100 kg/cm2 during 1 sec
in order to achieve a certain deformation; but applying a shear stress,
according to the Newtonian relation:

 ¼  0 ð2:18Þ

where  0 ¼ 1 sec1, only a 2.5 kg/cm2 force is needed. Thus it may be


concluded that shear-stressed pressure-sensitive joints need lower force
levels in order to undergo a deformation (i.e., pressure-sensitive joints are
weaker than classical ones).
As illustrated by the relations (2.15)–(2.18) the shear resistance
depends on the adhesive’s viscosity and elasticity. If the adhesive is
considered a Newtonian fluid, the shear resistance SH is given as a function
of the viscosity , the adhesive thickness h, and the tensile rate :

SH ¼ ð  Þ=h ð2:19Þ

If the solid obeys Hooke’s law the shear resistance depends on the shear
modulus G and the tensile rate :

SH ¼ ð  GÞ=h ð2:20Þ

Shear Dependence on Viscosity. Toyama and Ito [56] used the data of
creep testing (shear resistance) to calculate viscosity. Increasing the
molecular weight of CSBRs increases the Mooney viscosity [20]. At the
same time an increase in the molecular weight will improve the cohesive
strength (shear) of carboxylated SBR. For styrene block copolymers the role
of viscosity in shear measurement is observed at higher temperatures where
the molecular association does not work. At room temperature shear
resistance is due to the elasticity of the polymer, determined by segregation
and its parameters (e.g., sequence length and distribution), at higher
temperatures it is given by viscous flow influenced by copolymer
composition and global molecular weight.
Rheology of Pressure-Sensitive Adhesives 23

Shear Dependence on the Modulus. According to Woo [57] the shift


factors used to plot shear measurements at different temperatures on a
master curve were almost identical to the shift factors used for modulus
curves. Thus the time to fail in shear can be predicted from the viscoelastic
function. Correlations between the viscoelastic and PSAs performance
properties can be made by looking at those temperature regions which
correlate (via the time-temperature superposition principle) to the time
scales of the adhesive performance parameters. The magnitude of G0 in the
upper temperature range is indicative of internal strength or cohesion.
Initial peel, which is largely dependent upon the wet-out characteristics of
the polymer, is governed by G0 and tan  in this same temperature range.
Similarly, loop tack and quick stick properties are wet-out dependent, but
are associated with faster relaxation times and thus correlate with room
temperature viscoelastic properties [58]. The decrease of the modulus at high
temperature reduces the shear resistance. Milled rubber possesses a lower
elastic modulus and shorter modulus plateau than rubber from dried latex
[25]. For instance the G0 value at 20 C for milled smoked sheet is about
105 Pa. The thermo-mechanical degradation of the rubber through milling
does not change the tan  peak temperature (Tg). It does reduce the modulus
at high temperatures. This modulus reduction relates to the lower shear
performance of solvent-based systems.
The possibility to characterize application field-related properties of
PSAs and to correlate them with the rheology of the adhesive is illustrated
by formulating adhesives for medical tapes [59]. The adhesive used for
medical tapes is characterized by a dynamic shear modulus of about
1–2  104 Pa, a dynamic loss modulus of 0.6–0.9  104 Pa, and a modulus
ratio or tan  of about 0.4–0.6 as determined at an oscillation frequency
sweep of 1.0 rad/sec at 25% strain and 36 C. Adhesives with moduli higher
than the acceptable range display poor adhesive strength, while adhesives
with moduli below the acceptable range exhibit poor cohesive strength and
transfer large amounts of adhesive to the skin upon removal [59].

Shear Dependence on Time and Strain Rate. The shear resistance of


the adhesive depends on its internal cohesion. The cohesion is a function of
the inherent viscosity or modulus and thus it depends on the parameters
influencing the viscosity or modulus. Both viscosity and modulus are not
intrinsic material characteristics, they depend on the temperature and time
(i.e., on the nature and time history of the applied forces). Therefore, the
shear resistance will depend on the temperature and the time: in labeling
practice this behavior is illustrated by the low temperature resistance of the
adhesive joints and by the differences between statical and dynamical shear.
As is known, the viscosity depends on the loss modulus (viscous flow is the
24 Chapter 2

phenomenon that ‘‘loses’’ the energy). It also depends on the shear rate
(frequency):

0 ¼ G00 =! ð2:21Þ

At low frequencies 0 and the steady-state viscosity are the same. A static
shear test is based on low frequency deformation; therefore creep may be
regulated by viscosity.
The cohesive strength (shear) increases with increasing test rate or with
decreasing temperature [26]. When non-Newtonian liquids are subjected to
variable shear rates, the plot of the shear stress/shear rates no longer shows
a linear relationship [60]. In a diagram with double logarithmic scaling this
curve becomes a straight line. The mathematical equivalent of these two
curves is given by the following equations:

 ¼ K  D00 ð2:22Þ
ln  ¼ ln K þ n ln D ð2:23Þ

where D denotes the shear rate, K is a viscosity related coefficient, and n is


an exponent of the ‘‘power law equation’’ defining the shear rate dependence
on the viscosity. This exponent is known to vary for polymers between
0.3–1.0. Practically, it was demonstrated that an increase of the test rate
produces an increase of the cohesion and of the peel up to a critical value
[26]. As shown for cohesive strength measurements as a function of
temperature and shear rate [26], the principle of strain rate-temperature
equivalence can be applied.
The strain rate influences the shear resistance by bonding too. Bonding
is a diffusion and time dependent process. This is illustrated by the pressure
dependence of the lap shear adhesion for adhesives; the higher the pressure
the greater the bond strength [61].

Shear Dependence on Temperature. As discussed earlier, the cohesive


strength increases with increasing test rate or with decreasing temperature.
Shear data obtained as a function of the test temperature and shearing rate
can be shifted horizontally to form a single master curve, illustrating the
principle of time-temperature equivalence. Hamed and Hsieh [26] found
a good agreement between experimental and calculated values, using the
universal Williams-Landel-Ferry relationship for an amorphous rubber.
The finding that data can be shifted to form a master curve is evidence
of the importance of chain segmental mobility in controlling the shear
strength. If the chain segmental mobility (i.e., the ability to relax) is high
Rheology of Pressure-Sensitive Adhesives 25

(high temperature or low test rate) then the fracture stress is small. On the
other hand, a large stress is required to rapidly tear apart a PSA sample,
since the chains have little time to rearrange their microstructure in order to
accommodate the applied stress. This statement appears valid for peel
adhesion too. One should recall that the removability of a PSA label
strongly depends on the peeling rate.

1.3 Influence of Viscoelastic Properties on the Converting


Properties of PSAs and PSA Laminates
The converting properties of PSPs were discussed in detail in [62]. PSPs are
manufactured generally as web-like products. Most of them are laminates.
They are applied as web-like laminates or finite products. Before application
they have to be finished. In this case finishing means the transformation
of the continuous web-like product that has the optimal geometry for
manufacture into a product that has the optimal characteristics for use.
Convertability is the sum of the convertability of the adhesive and that of
the laminate. The converting properties of the adhesive and of the laminate
depend on the rheology of the adhesive. It has to be pointed out that, except
for hot-melt PSAs, the rheology of the uncoated adhesive (with an inherent
fluidity required for processing purposes) is different from that of the
converted material.

Converting Properties of the Adhesive


The liquid adhesive must be coated onto a release liner or face material.
Good coatability implies adequate machinability or processing properties
on the coater (metering roll, drying tunnel). During manufacturing,
transport, and coating, the adhesive fluid is subjected to shear forces,
going through a more or less pronounced change of the viscosity. The
coated shear-thinned adhesive has to wet the web, and the wet-out depends
on its viscosity. Except for hot-melt PSAs, the coated liquid adhesive layer
has to allow the elimination of the carrier liquid (drying) in order to form
a solid adhesive layer. Evaporation of the carrier liquid depends on its
diffusion through the adhesive layer (i.e., on the viscosity of the adhesive).
It may be concluded that the viscosity of the adhesive, i.e., the time (shear
rate)/temperature dependence of the viscosity, influences the coatability and
convertability of PSAs. It was mentioned earlier that, except for hot-melt
PSAs, the other PSAs are dispersed or diluted systems, their rheology being
different from that of the converted material. The coating-related rheology
will be covered in Section 3.
26 Chapter 2

Convertability of Laminate
Convertability of the PSA laminate means its ability to be processed into
finished products (labels, decals, etc.) by operations which influence its
dimensions (e.g., slitting, cutting or die cutting, embossing, folding,
perforating etc.) or its surface quality (e.g., printing, laquering, etc.). The
flow properties of the adhesive influence the migration or penetration,
oozing, and cold flow, thereby limiting the convertability of the laminate.
Thus the viscosity of the adhesive and its time/temperature dependence (i.e.,
a nonlinear character) determine the converting properties of the PSA
laminate. It should be mentioned that the converting properties of the
laminate really depend on the interaction of the PSA-laminate components.
These properties will be discussed in more detail in Section 3.2.

1.4 Influence of Viscoelastic Properties on End-Use


Properties of PSAs
The most important end-use properties of PSAs are the propensity to
labeling (dispensing) and bonding behavior. For some special PSPs the
debonding characteristics have to be taken into account too. The
application technology of other PSPs (e.g., tapes and protective films) is
less sophisticated therefore this chapter describes the label application only.
The end-use properties of various PSPs have been discussed in detail in [62].
Labeling is influenced by the adhesive properties (peel and tack) and by the
dispensing properties of the label. Removability or peeling off is influenced
by the adhesive properties. Thus the parameters influencing the adhesive
properties will also characterize the end-use properties.

Label Application Technology


Label application technology refers to labeling. Labels are either applied by
hand or with mechanical processes. Generally the label dimensions are
decisive for the choice of the application technology. Large labels will only be
applied by hand. On the other hand, according to the application technology,
reel and sheet laminates are manufactured. For PSA reels and sheets, a quite
different adhesive/cohesive balance is required (i.e., quite different tack/peel/
shear values or flow properties). High speed labeling guns need high tack
PSAs (so-called touch blow labels use no mechanical contact in label
application), whereas for high speed cutting, high modulus and high cohesion
PSAs are required. Dispensing and labeling were discussed in detail in [63].
In the label industry a basic difference exists between roll and
sheet supplied laminates (face material/adhesive/release liner) [40].
Pressure-sensitive adhesives for sheet applications must resist flying knife
Rheology of Pressure-Sensitive Adhesives 27

and guillotine cutting. A poor selection of the adhesive results in part in


oozing of the adhesive (gum deposits) on the cut surfaces and smearing of
the edges of the label stock with subsequent poor feed to the printing press.
The requirements are generally less critical for roll applications. Converting
in the paper label industry involves processes such as die cutting of roll
stock, and guillotining of sheet stock. Formulating approaches that improve
the high frequency modulus of the adhesive will enhance converting [2] since
die cutting and guillotining are such high frequency processes. The less
viscous and the more rigid the response of the polymer during the processes,
the cleaner the process tends to be. If viscous flow within the polymer is
significant during the converting operation, poor die cutting or poor
guillotining (knife fouling) can result.

Removability of PSAs
Conventional PSAs can be classified as either nonpermanent (2.7–9.0 N/
25 mm for 180 peel adhesion) or permanent (above 9 N/25 mm for 180 peel
adhesion). Repositionable PSPs are a special class of removable pressure-
sensitive products (labels and tapes) that stick to various surfaces but
remove cleanly and can be reapplied. The final adhesion builds up over a few
hours. Adhesives included in the nonpermanent category are used in the
manufacture of removable tapes and labels, protective laminates, and other
less durable products [64].
For nonpermanent, so-called removable adhesives, the flow properties
and cohesion of the adhesive as well as the anchorage of the adhesive to the
face stock are critical. In an ideal case, if the bond to the substrate is
nonpermanent, then a clean release from that substrate is encountered and
the adhesive remains on the face material. Another requirement for good
removable adhesives is the low peel level with a permanent character (i.e., no
build-up of the peel in time is acceptable). A clean release from the substrate
and no build-up of the peel with time are the minimal requirements for
removable PSAs. On the basis of the adhesive characteristics it is possible to
formulate the rheological criteria necessary for removable and permanent
adhesives.

Criteria for Removability. For certain applications PSPs are required


that display a low peel force and give a clean, deposit-free separation from
the substrate. The criteria for removability were discussed in detail in [44].
Generally a special balance between tack, peel adhesion, cohesion, and
anchorage is required in order to ensure removability. Removability requires
a breakable bond at the adhesive/substrate interface. Bond breaking is an
energetic phenomenon. For removability the whole debonding energy should
28 Chapter 2

be absorbed by the adhesive itself in such a manner that no failure occurs in


the adhesive mass. In this case the energy is used only for the viscous flow and
elastic motion of the macromolecules. The major danger is that if the viscous
flow is too pronounced, bonds will fail in the adhesive mass. (It should be
mentioned that for special removable protective films the regulation
absorbtion of the debonding energy is achieved using a deformable carrier
material [65].) Some low tack, crosslinkable dispersions are good removable
PSAs. Their good elasticity yields low cold flow and low build-up of the peel
during storage [66]. On the other hand, a certain degree of cold flow is needed
in order to obtain removability. Adding 2–20% plasticizer during the
polymerization of acrylics leads to a removable adhesive but the plasticizer
may migrate out after storage [67]. Furthermore cohesive strength is
necessary to avoid stringing. It must be stated that the adhesion depends
on both the rheological features and chemical affinity. It was shown [68] for
certain PSAs that the apparent separation energy was approximately the
same when pulled away from either a glass or a Teflon surface. The adhesion
of these adhesives must be attributed largely to their rheological features
rather than to selective wettability.

How to Achieve Removability? A special adhesion/cohesion and/or


plasticity/elasticity balance is required in order to achieve removability.
Flow properties allow rearrangement, relaxation, and transformation of the
debonding energy into viscous flow, therefore softening of the adhesive is
a prerequisite for removability. It is well known from the field of caseine-
based adhesives [69] that polyethylene glycols added to caseine and dextrine
glues control the humidity absorption, and thus the tear of gummed papers
at low environmental humidity, and avoid blocking at high humidity,
thereby acting as plasticizers (at a loading rate of 1–2%).
Flow properties are correlated to creep which, in turn, is a function of
molecular weight and plasticizer loading. Higher molecular weights and low
plasticizer levels reduce the tendency to creep. For removable PSAs the
molecular weight should be limited. Permanent adhesives need no fluidity
during debonding, as they must display debonding resistance when subject
to high stress rates and large forces. They need to display a modulus and
viscosity increase as a function of the stress rate. On the other hand,
removable adhesives need easy debonding and viscous flow. The applied
stress must be able to cause movement within the bulk of the material.
Theoretically the applied stress for debonding can be minimized by using
stress-resistant polymers, fillers, and primers; stress-resistant polymers are
those which develop a controlled crosslink density or are soft internally.
Primers are somewhat flexible and generally promote good anchorage to the
face stock. Thus, they can absorb expansion or impact stresses without
Rheology of Pressure-Sensitive Adhesives 29

adhesive failure. Generally, as discussed in [44], removability can be achieved


by regulating the chemical composition of the adhesive, and the structure and
geometry of the adhesive, and the carrier properties and by means of the
manufacturing technology. The main methods used to change the chemical
composition (and macromolecular characteristics) of the adhesive compo-
nents, e.g., tackifying, softening, and hardening (crosslinking, filling, etc.)
influence the removability too. Formulation windows with boxes shifted to
the left and up (i.e., for lower frequencies) are harder [70].

Influence of the Viscosity/Modulus on Removability. As shown


earlier, hardening of the adhesive increases its modulus (i.e., decreases its
ability to creep and its ability to wet the surface), and thus decreases the tack
(i.e., peel adhesion after very short dwell times) as well as the peel resistance.
It can therefore be assumed that hard adhesives display low peel adhesion
(i.e., are removable). According to [6] there are two basic types of removable
adhesives, namely PSAs which wet the surface poorly and PSAs which
adhere easily but are easily removed. The first type of PSAs is highly elastic
and has low tack. Its removability depends on the adhesive not establishing
a good contact with the surface. Sometimes the poor contact is built-in
mechanically. These PSAs might have inclusions that prevent the adhesive
from achieving full contact with the surface, as the adhesive contains
physical particles that do not deform completely and therefore limit the
contact area between adhesive and substrate. These PSAs possess a high
modulus. Thus, hardening of the polymer (modulus increase) is a possibility
to obtain removable adhesives.
As discussed in [71] the build-up of ordered multiphase structures
can modify the chain mobility. Such structures include fillers, crosslinks,
crystallites, and molecular associations. Fillers modify the Tg also. Really
such hardening of the adhesive using a composite structure works like other
methods for the control of the contact surface (e.g., pattern-like contact
points, pattern-like nonadhesive points, decrease of the coating weight etc.).
It should be mentioned that in filled hard adhesive formulations the
absorbtion of the debonding energy is improved too [72]. As illustrated in
[73] the degree of filling (e.g., filling, filling and foaming) strongly influences
the adhesive characteristics and adhesion build-up. The rheology of such
reinforced systems and its influence on the adhesive and end-use properties
were discussed in detail in [74]. The most important fillers used in PSAs and
their functions are listed in [40]. According to [75] current removable
adhesive formulations include high molecular weight polymers that reduce
flow on the surface and prevent adhesion build-up.
The second type of removable PSAs displays high tack and is
characterized by a low modulus. In order to soften the adhesive such
30 Chapter 2

formulations use plasticizers or other low molecular weight viscous com-


ponents (paraffinic oil, polyisobutylene etc.). The presence of 8% plasticizer
may improve the elongation by 3000% [66]. The softening effect is
illustrated by the lowering of the minimum film-forming temperature
(MFT) (e.g., 6% plasticizer decreases the MFT from 23 C to 0 C), or the
decrease of the tensile yield strength (e.g., 6% plasticizer decreases the
tensile yield strength from 13.0 to 1.0 N/mm2). The elongation of PVAc
PSAs (normally 6–10% at break) was significantly increased by adding
plasticizers to up to 100–1000% [76]. Differences in plasticity may be due to
the molecular weight of the plasticizer or different interactions between
chain segments and plasticizers. Long-chain plasticizers exert a strong
plasticizing effect. Generally plasticizers decrease the modulus value and
influence the frequency dependence of the loss tan  [4]. For nonplasticized
adhesives loss tan  is low at low frequencies. For plasticized adhesives loss
tan  shows a maximum at lower frequencies, and peel and quick stick also
exhibit maxima at lower frequencies.
Plasticizers may be incorporated in peelable PSAs to soften the
adhesive and thus improve peelability. However, some plasticizers can exert
a tackifying effect on adhesive polymers and this may limit the amount used.
Typical plasticizers include phthalate esters and polyalkylene ether
derivatives or phenols, the principal requirement being compatibility with
the main adhesive polymer and the tackifier so as to avoid or minimize
migration of the plasticizer. It is possible to add quite large quantities of
plasticizer, even up to 50% by weight on the solid adhesive; however,
usually 10–20% by weight are added [6]. Other plasticizers which can be
employed, include the well-known extruder oils (aromatic, paraffinic, or
naphthenic) as well as a wide variety of liquid polymers. Satas [6] showed
that the addition of a plasticizing oil which is compatible with the mid-block
of a styrene-isoprene-styrene (SIS) block copolymer results in an adhesive
with lower peel force. Such plasticizers are used for HMPSAs.
In some cases softening and hardening of the adhesive are used
together. Mechano-chemical destruction of the macromolecules of natural
rubber (mastication) followed by postcrosslinking leads to PSAs used for
removable protective films. Such crosslinked natural rubber formulations
are softer than certain common noncrosslinked acrylic recipes [77]. Internal
crosslinking by special comonomers and the simultaneous use of plasticizers
have also been suggested [78].

Influence of Viscoelasticity on Applied Labels


After some time labels show a higher peel resistance than freshly applied
ones (i.e., the peel resistance increases with the dwell time). This build-up of
Rheology of Pressure-Sensitive Adhesives 31

the peel resistance is due to the flow and contact build-up of the adhesive.
Therefore, one should pay attention when evaluating the peel force, that is,
peel should always be measured after a well-defined (normalized) dwell time.

1.5 Factors Influencing Viscoelastic Properties of PSAs


The material characteristics, chemical composition/structure, and environ-
mental and experimental conditions influence the viscoelastic properties
of PSAs. The influence of the material characteristics on the viscoelastic
properties of the PSAs will be discussed first. Numerous empirical
correlations between the molecular structure, molecular weight, molecular
weight distribution (MWD), branching, and rheology are generally valid for
chemically quite different polymers [79].

Influence of Material Characteristics on Viscoelastic


Properties of PSAs
The molecular character of the base polymer, its chemical composition, and
structure influence its viscoelastic properties. The influence of the molecular
weight on the viscoelastic properties of the PSAs will be examined first.
Viscosity and the elastic modulus are the most important parameters
characterizing the viscoelastic behavior of PSAs. Both parameters depend
on the molecular weight of the base polymer.
Dependence of the Viscosity of PSAs on the Molecular Weight. The
viscosity of macromolecular compounds is a function of the molecular
weight. Their zero viscosity 0 depends on the weight average molecular
weight according to a correlation of the form [80,81]:

0 ¼ f ðMW Þ ð2:24Þ

where  ¼ 3.4.
Zosel [16] demonstrated the strong influence of the molecular weight
of acrylic PSAs on their viscosity and on their tack. The increase of the
molecular weight in the range of 103–107 produces an increase of the
viscosity from 102 to 1010 Pas. Zosel also showed the existence of a
maximum in the tack/molecular weight graph for polyisobutylene. The
pronounced increase of the viscosity with the molecular weight limits the
usable molecular weight range for hot-melt PSAs, limiting (unfortunately)
their cohesion as well. (According to [82] for such adhesives 80 Pas is the
preferred maximum viscosity value for high speed coating.) It is well known
to skilled formulators that block copolymers have lower viscosities than
natural rubbers. (Their viscosity depends on sequence distribution and
32 Chapter 2

branching too). The solution viscosity of the high polymers also increases
with their molecular weight. Thus a viscosity-imposed balance between
processing and end-use properties, that is, a solid content/molecular weight
balance for solvent-based PSAs is always a limiting factor, keeping in mind
that the solid content of common solvent-based PSAs does not exceed
30–40%. On the other hand the additives in an adhesive formulation, their
nature, and their concentration also influence the viscosity of the blend.
Practical examples are given by plasticizers and tackifiers. Micromolecular
tackifiers (plasticizers) may be considered as diluting agents. Thus the
viscosity of the polymer ‘‘solution’’ obeys an exponential law, depending on
the polymer concentration C, namely
Z
¼ ðCÞ ð2:25Þ

At first rubber resin formulations were used for PSAs. To achieve


viscoelastic behavior, the elastic component (rubber) had to be mixed
(formulated) with a viscous component. Either macromolecular or
micromolecular compounds can be used as viscous components. The best
known are resins and plasticizers; both are suggested as tackifiers.
Generally, low molecular weight resins and relatively high molecular
weight solvents are used as tackifiers. Common HMPSA formulations
incorporate a high level of plasticizer, usually naphthenic oil or liquid resin.
The level of the tackifier resin influences the resulting viscosity. As an
example, Zosel [16] demonstrated that the ‘‘zero viscosity’’ of tackified PSAs
decreases up to 80% resin loading level, resulting in better flow properties.
The diluting effect of the tackifier resin depends on its own viscosity too.
Generally, soft (liquid) resins impart aggressive grab and quick stick, while
hard resins help retain good cohesive strength and creep resistance. Thus
resins with a softening point below about 50 C impart tack, but poor
cohesive strength, while those above 70 C lead to poor tack properties [83].
It was shown earlier that the peel adhesion depends on the viscosity.
Peel is a function of molecular contacts which depend on the diffusion. The
rate of the adhesive diffusion into the face stock is inversely proportional to
(MW)2. Thus the number of contact points decreases with (MW)2, and thus
tack and peel decrease with the molecular weight [26]. On the other hand the
chain elasticity increases with the molecular weight as well as the so-called
‘‘critical flaw size,’’ resulting in a lower peel at a higher molecular weight.
Dependence of the Elastic Modulus on the Molecular Weight. Raw
materials for PSAs are characterized by two special features: the Tg and the
modulus of elasticity. Both of these properties and their dependence on the
Rheology of Pressure-Sensitive Adhesives 33

base polymer will be examined in Chap. 3. The dependence of the elastic


modulus on the molecular weight will be studied in this chapter. It should be
mentioned that toughness properties of polymers generally increase with
increasing molecular weight up to a point, then level off. An extension
towards higher temperatures of the plateau region of the dynamic modulus
master curve can be achieved by the increase of the molecular weight of
CSBR [20].

Influence of Chemical Composition/Structure on Viscoelastic


Properties of PSAs
As discussed earlier and detailed in the next chapter, the rheological
properties of PSAs are characterized by viscosity and modulus, and the
Tg of the base polymers. All these parameters depend on the chemical
composition/structure of the PSAs. The dependence of the viscosity on the
chemical composition/structure is discussed next.
Viscosity of the Base Elastomer. As known from polymer chemistry
the inherent viscosity of polymers depends on the intermolecular mobility of
the polymer molecules. This mobility is a function of the chain length
(molecular weight) and structure (e.g., branching) and reciprocal interaction
of the polymer molecules (polar forces). It must be noted that both the
viscosity and its dependence on the temperature are a function of the chemical
composition/structure of the polymer. The flow activation energy (e.g.,
Arrhenius) depends on branching [80]. The chemical composition of the base
elastomer also influences the viscosity of the PSAs. Acrylic hot-melt PSAs
with non-Newtonian behavior differ from compounded hot-melt PSAs and
from conventional solvent-based rubber/resin PSAs, and allow for the
development of high shear adhesives with a low processing viscosity [84].
Styrenic block copolymers have a relatively low solution or melt viscosity,
since in solution or at temperatures above the Tg of polystyrene, the
systematic polystyrene domain within the block copolymer no longer exists.
Generally the viscosity of the adhesive is a function of the raw material class,
but viscosity changes are possible in the same raw material class also. For
instance the saturated styrene block copolymers (e.g., Kraton G) possess
higher solution viscosity than the common unsaturated ones [85].
Influence of Tackifier Resin. Pressure-sensitive formulations are
based on an elastomer or viscoelastomer and a viscous component. The
most important viscous components are tackifier resins. Exhibiting a surface
free energy close to that of the elastomer, the resin (filler) is perfectly wetted
by the organic matrix, and as a consequence, good dispersions (i.e., lower
viscosities) are obtained [86]. Addition of resins to rubber-based adhesives
34 Chapter 2

decreases the viscosity [87]. The addition of low melting point resins to a
hot-melt formulation with block copolymers (Kraton GX 1657) decreases
the melt viscosity [88]. As discussed in Chap. 5, the mutual compatibility of
the formulation components strongly influences their rheological properties.

Influence of Environmental and Experimental Conditions on


Viscoelastic Properties of PSAs
The coating of PSAs, the converting of the laminate, and the end-use of the
pressure-sensitive labels occur under very different temperature and stress
rate conditions. Because of their macromolecular chemical basis, the
properties of the PSAs are time/temperature dependent. Thus, for adequate
manufacturing and use their time/temperature sensitivity must be studied.
For a better understanding of this dependence the temperature and time
dependence of the viscoelastic properties will be examined separately. First,
the dependence of the viscosity and of the modulus as a function of the
temperature will be reviewed.
The physical and mechanical characteristics of a polymer depend on
its chemical composition/structure. Chemical and macromolecular char-
acteristics determine the internal mobility of the polymer chains. On the
other hand, the ambient temperature influences the rate of chemical and
physical interactions (i.e., the chain mobility) and thus the viscoelasticity of
the polymer. Liquid characteristics (viscosity) or solid behavior (modulus)
of the macromolecular compound are both functions of the temperature.
Adhesion properties require a special balance of the viscous liquid/elastic
solid state. Therefore the temperature also influences the adhesive
properties.

Dependence of Viscosity on Temperature. In general the viscosity is a


function of temperature. For low molecular weight, nonassociated liquids
(Newtonian systems) the viscosity depends on the temperature according to
the equation of Arrhenius:

 ¼ A exp ðEo =RT Þ ð2:26Þ

where Eo represents the activation energy, R is the universal gas constant,


and T the absolute temperature. For highly viscous liquids [89]:

ln  ¼ A exp ½B=ðT þ CÞ þ D ð2:27Þ


Rheology of Pressure-Sensitive Adhesives 35

where A, B, and D are experimental parameters. For non-Newtonian


systems the Williams-Landel-Ferry equation is valid, that is:

Eo
 ¼ A exp ½B=ðT  To Þ þ ð2:28Þ
RT

where A and B are experimental parameters, and To is the temperature


at which the free volume is zero. This equation remains valid for
T < Tg þ 100 K. Generally the temperature-dependent shift factor aT for
the viscosity is given by [90]:

aT ¼ o ðTÞ=o ðTo Þ ð2:29Þ

where o is the viscosity at zero shear rate. For amorphous polymers


(e.g., PSAs):

ln aT ¼ ½C1 ðT  To Þ=½C2 þ ðT  To Þ ð2:30Þ

where C1 and C2 are material-dependent parameters. For partially crystal-


line polymers (e.g., resins):

ln aT ¼ Eo =Rð1=T  1=To Þ ð2:31Þ

For the temperature dependence of the dynamic viscosity, two models


generally are considered. The model of Hirai and Eyring [91] indicates the
existence of an energy barrier between holes in a liquid and this barrier is the
rate determining step of viscous flow; the model applies transition state
theory to explain the temperature dependence of viscosity, namely:

 ¼ 0 expðE=RTÞ ð2:32Þ

The constant 0 includes both the molecular volume of the liquid and
the activation entropy of flow and is generally assumed constant. Another
model [92] supposes the formation of holes (i.e., the free volume as the rate
determining step), introducing the expression:

 ¼ A0 exp ðB  Vo =Vf Þ ð2:33Þ

A0 and B are two constants, and Vo and Vf are the occupied and the free
specific volume of the liquid, respectively. The temperature dependence
of the viscosity is explained by that of the volumes; by rearranging the
36 Chapter 2

well-known Williams-Landel-Ferry equation [93]:


 
 C1 ðT  Tg Þ
log  log T ¼  ð2:34Þ
o C2 þ ðT  Tg Þ

where the constants C1 and C2 depend on the fractional free volume.

Influence of Time on the Viscoelastic Properties of PSAs. The inter-


molecular mobility of the macromolecules is a time-dependent, relatively
slow process. This means that molecular rearrangements require a certain
time. Practically, this time may be longer, equal to, or (rarely) shorter than
the duration of the external stress. Therefore, internal (molecular) motion
does not always occur synchronously to the external one. Macroscopic flow
behavior and elastic recovery are characterized by viscosity and modulus.
Both of these properties are the result of internal microscopic interactions
(i.e., of the time needed for these interactions). Therefore, the viscoelasticity
of PSAs depends on the time and on the rate of the applied external stress.
Williams, Landel, and Ferry [93] defined a shift factor allowing the
correspondence between time and temperature influences. The shift factor
depends on the Tg and on the reference temperature [94]. At relatively low
strain rates, such as those acting during initial bonding or wet-out of
the surface, the polymer molecules have time to slide past one another
and undergo viscous flow [2]. This effect, as would be expected, is more
pronounced at higher temperatures, where molecular mobility is greater.
Storage and end-use of PSAs are both low strain processes. Alternatively,
strain rates are significantly higher during peel measurements. Here the
molecules are not able to flow past one another sufficiently fast, and the
polymer stiffens. This stiffening constitutes an increase in the elastic
contribution to the rheology of the polymer. When the elastic components
predominate, more of the bond rupture energy is stored resulting in a higher
peel and tack. High speed converting and end-use operations (e.g., die
cutting, removal of labels) are high strain processes [2].
The time and temperature dependence of the viscoelastic behavior
of polymers may be interpreted by the corresponding dependence of the
relaxation spectra and by conformational changes within the macromole-
cular chains as shown by Schneider and Cantow [95]. The corresponding
contributions to the viscoelastic functions are expressed by the frequency
shift factor aT which allows the superposition of viscoelastic isotherms into
composite curves at To, the reference temperature [95]:

T ¼ ½2
o T To =½2
o To T ð2:35Þ
Rheology of Pressure-Sensitive Adhesives 37

From Eq. (2.33) it follows that the temperature dependence is


connected mainly to the translational friction coefficient per monomeric
unit
o although the mean square end-to-end distance per monomeric unit a2
may vary significantly with temperature for most polymers. The shift factor
is related to the dynamic viscosity 0 according to:

T ¼ ð0 To o Þ=ð0o T Þ ð2:36Þ

where is the density of the polymer. Intermolecular polymeric motion


influences the polymer flow, and this motion is a function of time and
temperature. From the mechanics of fluids or solids, it is known that
mechanical properties decrease as temperature increases (i.e., the molecular
motion is enhanced). In a similar manner there is an opportunity for
molecular motion if the rate of the applied force is lowered. Thus high
temperature and low forces act similarly on the molecular motion and on
phenomena related to it (e.g., viscosity).
Dependence of the Viscosity on the Shear Rate. Pressure-sensitive
adhesives are viscoelastic materials, that is, their mechanical properties are
temperature and time dependent. The material characterization requires the
measurement of a material function such as viscosity or modulus as a
function of temperature and time. As can be seen from Fig. 2.4, the viscosity
is a function of the shear rate.
At low shear rates, there is a short interval where the viscosity does not
depend on the shear rate (this is the so-called zero-viscosity). At higher shear
rates, there is a linear or nonlinear decrease of the viscosity as a function
of the shear rate (Newtonian or non-Newtonian systems). The decrease of
the viscosity with the increase of the shear rate is given by the elastic
component of the adhesive [96]. The existence of a shear rate-dependent
viscosity domain implies that static methods for testing the cohesive strength
of PSAs cannot yield real information about their behavior at high stress
rates (peel, cutting). This statement can be confirmed through the definition
of the adhesion failure energy (separation energy W), upon debonding
through peel or loop tack [97]:
Z
W¼ F  v dt ð2:37Þ

where v is the debonding (separation) rate and F the applied force. The
exponent value for a Newtonian liquid that has a shear rate-independent
viscosity is n ¼ 1 [see Eq. (2.4)]. Changes in the polymer structure may be
associated with changes in the viscosity-shear rate dependence. Detailed
38 Chapter 2

Figure 2.4 Dependence of the viscosity on the shear rate. Decrease of the
Brookfield viscosity for different (1 through 3) water-based acrylic PSAs as a
function of the stirring rate.

studies of elastomers showed mastication increased n from 0.25 to 0.5 [98].


Mastication in a mixer yields less change of the polymer elasticity than
mastication on a roll mill. Consequently rubber-based PSA formulations
manufactured without a priori calendering or roll mixing exhibit a higher
viscosity and a very broad range of viscosity values (see Fig. 2.5) as a
function of the solid content.
The viscosity of PSAs influences the contact build-up of the liquid
adhesive. For easy bonding low viscosity values are required. In order to
obtain a high debonding resistance (high peel and shear) high viscosity
values are needed. Thus one can wonder what low and high viscosity values
mean? What should be chosen, the low or the high viscosity version of the
adhesive? Generally, with common formulations the elastic component of
the polymer predominates. That is, high peel and high shear as exhibited by
high viscosity polymers are used, and bonding (flow) of the adhesive is
obtained by the use of longer dwell times. For those highly viscous polymers
the value of the viscosity is given indirectly as cohesion (shear). On the other
hand, removable PSAs are more plastic and possess a lower viscosity,
Rheology of Pressure-Sensitive Adhesives 39

Figure 2.5 Range of solution viscosity values. Tackified, unmilled rubber/resin


PSA: solution Brookfield viscosity as a function of the solid content.

causing stringing in many cases. Viscosities higher than 105 Pas are required
in order to avoid this phenomenon. Generally, viscosities of PSAs vary
between 104 and 105 [55].

Dependence of the Modulus on the Stress Rate. Like the viscosity the
modulus of the PSAs also depends on the stress rate. As an example, Zosel
[16] measured the modulus of tackified PSAs as a function of the resin
concentration, at a rate above 102 sec, and demonstrated the existence of a
maximum at a resin concentration of 50%. On the other hand Dahlquist,
working at faster rates (t < 102 sec) showed the existence of a maximum at
60% resin concentration [23].

Influence of the Elasticity/Plasticity Balance on the Properties of


PSAs. If the elastic components are predominant in the rheology of the
polymer, more of the bond rupture energy is stored, resulting in higher peel
and tack properties [2]. Some of the energy supplied by the application of
stress to a material is stored energy, in contrast to that which is dissipated in
overcoming viscous drag. The two recognized means of storing mechanical
energy are inertia (kinetic) and elasticity (potential) of which the latter
commands the most attention as the determining factor which allows us
to distinguish permanent adhesives from removable ones. If the time scale
of the measurement is short enough, any liquid will respond elastically to
prevent relaxation of the applied stress (e.g., high speed peel).
40 Chapter 2

2 RHEOLOGY OF PSA SOLUTIONS AND


DISPERSIONS

In order to be coated onto the face stock material the adhesive must be in
the liquid state. This state may be achieved by melting (hot-melt PSAs),
dissolving (solvent-based adhesives), or dispersing (adhesive dispersions) the
PSAs. Dissolving and dispersing the adhesive produce diluted adhesive
systems, where flow properties are due to the carrier liquid. The rheology of
diluted PSAs also depends on the carrier liquid. Generally, there are aqueous
and solvent-based adhesive solutions, and aqueous or solvent-based
adhesive dispersions. The most important are the organic solvent solutions
of adhesives (solvent-based adhesives) and the water-based dispersions of
adhesives (water-based adhesives). In fact most solvent-based adhesives like
rubber resin adhesives, contain some dispersed material, and most of the
dispersion-based aqueous adhesives contain some dissolved substances
(e.g., thickeners, protective colloids, tackifiers, etc.).
The rheology of PSAs in the liquid state (molten or diluted) is very
important for the coating process. As stated in [99] the main processing
parameter by coating is the viscosity of the adhesive. In this case one needs
to be concerned with coating rheology. Wetting out, coating rheology,
coating speed, and versatility are the most important flow parameters
influencing the convertability (coatability) of the adhesive. All of these
parameters depend on the physical state of the adhesive, either as a diluted
liquid system, yielding an adhesive layer by evaporation of the carrier liquid
(water or solvent), or as a 100% solid system, giving a solid adhesive layer
by a physical (hot-melt) or chemical [ultraviolet (UV)- or electron beam
(EB)-cured] process. Therefore the coatability and the coating rheology of
diluted and undiluted systems will be discussed separately.
The molten state of hot-melt PSAs differs from the coated, solid state
by the value of their viscosity only. Therefore, it may be assumed that the
rheological properties of hot-melt PSAs during converting do not differ
from those of solid PSAs. Hot-melt flow is not an isothermal process; it
is more sensitive to thermal transfer and temperature than the room
temperature cold flow of the solid, coated adhesive layer. However, it is
designed to possess an adequate fluidity at a given temperature (i.e., an
optimized viscosity). Because of the relatively high viscosity of the fluid hot-
melt PSA and the short period of the liquid state, phenomena encountered
by diluted PSA systems (e.g., wetting out, migration, mechanical stability,
and volatility of low molecular weight components) appear less important.
Diffusion problems for hot-melt PSAs are due not to the carrier liquid (and
do not occur during the coating) but mainly to the low molecular weight
polymers. Hence there is no need to discuss the coating rheology of hot-melt
Rheology of Pressure-Sensitive Adhesives 41

PSAs separately. Unfortunately, the nature and concentration of the carrier


liquid in a diluted PSA play a decisive role in their rheology. Thus it appears
necessary to discuss the rheological features of those systems. Because of
the basic differences between solution adhesives and polymer dispersions,
considerable differences in rheology exist and it is necessary to discuss the
coating rheology of both these diluted PSA systems separately.

2.1 Rheology of PSA Solutions


The rheological behavior of polymer solutions is a result of solute/solvent
concentration; it depends on the properties of the dissolved PSAs and of the
solvent system. The influence of the dissolved polymers and their chemical
composition and macromolecular characteristics will determine their
interaction with the solvent. On the other hand, like the rheology of the
solid polymer, the flow characteristics of the polymer solution also depend
on time and temperature. Coating rheology includes the flow properties of
the wet and dried adhesive, leading to a wettable, stable, running time-
independent liquid layer, with a smooth surface after drying under the
applied shear forces. Coating rheology of solvent-based PSAs is the
rheology of a true solution, thus mainly viscosity driven and more simple
than that of water-based systems. Such viscosity depends on the nature and
concentration of the formulation components [100]. Typical viscosity values
of acrylic solvent-based PSAs are situated in the range of 1700–
25,000 mPas. The first step of the coating process is the wet-out of the
solid surface by the flowing adhesive. Generally, wetting out is a function of
surface tension, viscosity, and density of the adhesive layer. For coating
of solvent-based formulations a surface tension higher than 38 mN/m is
required. For coating of aqueous formulations the surface tension must be
higher than 46 mN/m [101]. It should be kept in mind that solvent-based
adhesives are usually highly viscous systems (more than 10,000–
15,000 mPas) with low surface tension solvents as liquid carrier, and
being (at least partially) directly coated on high surface energy webs. Thus,
the wet-out of solvent-based PSAs is less difficult and less critical than that
of water-based ones. Surface tension depends on polymer and solvent
nature, viscosity, and density. Synthetic raw materials for solvent-based
adhesives are homogeneous, whereas natural raw materials are not. In
general rubber/resin adhesives display a broader variation of the solution
properties than acrylic-based ones. On the other hand solution polymeriza-
tion allows better chemical control, thus some macromolecular properties
of solvent-based polymers are superior to those of water-based ones (e.g.,
MWD, sequence distribution, stereoregularity, gel content, etc.). The use of
special solvents or solvent mixtures constitutes an additional possibility for
42 Chapter 2

surface tension control of solvent-based adhesives. Generally solvents used


for solvent-based PSAs have a low surface tension (25–30 mN/m), much
lower than that of water (72 mN/m) or even of aqueous dispersions (usually
35–42 mN/m). The surface tension of solvent-based PSAs causes less
concern than that of aqueous dispersions. Viscosity control of solvent-
based PSAs can also be realized by the use of fillers.

Special Features of the Rheology of PSA Solutions


Polymer solutions are multicomponent systems, where the rheological
properties of the components and their mutual interactions may play an
important role. The rheology of the basic polymer and of the solvent-system
interface is a function of the nature and concentration of the components. In
some cases the procedure for dissolving the basic polymers, respectively for
eliminating the carrier solvent, also influences the final adhesive properties.
These aspects will be examined in more detail in Chap. 9. In this section the
most important features of the rheology of polymer solutions are
summarized. The flow properties of PSA solutions are important for their
coating behavior.
The coating rheology of the polymer solutions may be different from
that of the polymer without the solvent carrier, due to the multicomponent
(system) character, and to the quite different manufacturing and end-use
conditions (stress rate and temperature) for PSA solutions.
Polymer-Dependent Characteristics. Polymer solutions are visco-
elastic, with a non-Newtonian character, their behavior being determined by
the nature and concentration of the components. Theoretically the most
important rheological characteristics of PSA solutions will be described by a
correlation depending on the macromolecular and chemical characteristics
of the polymer (solute) and the chemical nature and concentration of the
solvents used. The time/temperature dependence of the viscosity is also
pronounced. Solvent-based adhesives are mostly solutions of rubber/resin
blends or acrylic copolymers. Their viscosity depends on the chemical
composition and molecular weight of the base polymers. Mechano-chemical
degradation of the base polymer, e.g., mastication of natural rubber,
reduces the molecular weight and its dispersion. As discussed in [102] the
ratio of the viscosities of adhesive compositions based on unmilled or milled
rubber is about 5–10. For rubber/resin-based PSAs viscosity control is
limited to the change of the MW/gel content ratio. Because of the high
dispersion of these values, the use of an additional 10–15% of solvent in
these formulations is common practice. Resins tend to reduce viscosity and
increase solids content [103]. The molecular weight of solvent-based
adhesives is lower than that of water-based dispersions. However, these
Rheology of Pressure-Sensitive Adhesives 43

relatively low molecular weights are high enough to yield high viscosities. As
shown earlier viscosity directly influences the wetting out. The wet-out is a
function of surface tension, viscosity, and density. Surface tension depends
on the solvents used, and on the density of both polymer and solvent.

Solvent-Dependent Characteristics. Generally polymer solubility


depends on solute/solvent interactions, thus the viscosity of PSA solutions
depends on the nature and concentration of the solvents used. The viscosity
of solvent-based PSAs depends on the molecular weight, chemical nature,
and solids content of the formulation. With suitable molecular weight values
and a solids content of 20–25%, solvent-based PSAs display viscosities of
more than 10,000 mPas. High solids solvent-based acrylic PSAs have a
solids content higher than 60% and a Brookfield viscosity of 15,000 mPas
(at 25 C) at 65% solids [104]. Gravure cylinders can be used for high solids
(50–80%) highly viscous solutions having a viscosity of more than 50 Pas
and for low viscosities of 12–15 mPas too [105]. It should be kept in mind
that solvent-based PSAs differ from the water-based PSAs not only by their
higher viscosity but also through the easier viscosity control. Problems may
appear for crosslinked systems (for instance for crosslinkable, solvent-based
acrylates) where diluting may affect the stability of the adhesive [106]. As
discussed in detail in [99] for such systems viscosity is pot life related. As is
known from practice, it is possible to formulate solvent-based PSAs
according to their viscosity (for doctoring ease) or to their solids content
(for high drying speeds).
Rubber/resin PSAs mainly use hydrocarbon-based solvents, thus the
choice of the solvent as an aid cannot play an important role in the control of
their viscosity. Currently binary or ternary solvent mixtures allow better
viscosity control. On the other hand, the rheology of the solvent-based PSAs is
sensitive towards drying speed (i.e., the rate of the diffusion of the solvents).

Technology-Dependent Characteristics. The rheology of the polymer


solutions may depend on the manufacturing and coating technology.
Solvent-based PSAs are made by dissolving the basic polymer. Because
of the sensitivity of the macromolecular compounds towards mechano-
chemical destruction, the dissolving technology will influence the spread of
the viscosity values. On the other hand the same PSA solution may give
coatings of different quality, according to the geometry (shear) of the
metering device [99]. As stated in [99] the main parameters of the wet film
quality for solvent-based adhesives are the surface tension, solids content,
viscosity, solvent characteristics, and temperature. The quality of the dry
coating depends on the evaporation rate of the solvent, which is a function
of the coating thickness, solids content, solvent characteristics, and drying
44 Chapter 2

conditions. Another feature is given by the sensitivity of the coated adhesive


layer towards drying speed (rate of diffusion of the solvents used). Mainly
for rubber/resin adhesives based on hydrocarbon solvents, superficially
stripped, textured surfaces may be obtained if the solvent ‘‘boils’’ during
drying and the fluidity of the dry layer is lost.
Adherend-Dependent Characteristics. Liquid components of solvent-
based PSAs may interact (physically or chemically) with the contact surface
of the adherent, resulting in better adhesion.

Time/Temperature Dependence of the Viscosity of PSA Solutions


A shear-induced viscosity decrease of polymer solutions is governed by
molecular orientation and disentangling. For Newtonian liquids at low shear
rates it is possible to restore the polymer network in the solution through
relaxation. High shear implies the decrease of the network points. Therefore,
the shear modulus G is given, as a function of the network points vc,

G ¼ vc RT ð2:38Þ

The dependence of the viscosity on the shear is a general phenomenon


also valid for polymer (adhesive) dispersions. It should be mentioned,
however, that in a manner different from adhesive dispersions, the
mechanical stability of the system for polymer solutions is viscosity (or
solids content) independent.

2.2 Rheology of PSA Dispersions


The coatability of liquid adhesives is a function of wetting out, coating
versatility, coating rheology, and coating speed. On the other hand the
coating rheology influences the wet-out and the coating versatility, and
depends on the coating speed. The wetting out of water-based systems
constitutes a difficult problem because of the high surface tension of the
carrier liquid (water). Solvent/solute interactions in ‘‘real’’ solutions give rise
to the increase in viscosity. Water-based dispersions generally possess a
lower viscosity while ready-to-use PSA dispersions possess water-like low
viscosities. Their low density and the transfer coating technique used
also contribute to the coating difficulties. In a different manner from
solvent-based PSAs, improvement of the wet-out by the use of surface active
agents causes irreversible changes in the adhesive properties. Thus an
adequate coating rheology appears important.
Water-based dispersions have a limited shelf life and thus the coating
rheology also depends on the age of the dispersion. This is a function of the
Rheology of Pressure-Sensitive Adhesives 45

dispersed polymer particles (macromolecular and dispersion characteristics)


and of the carrier liquid as well. Only small changes in the composition
of the carrier liquid are allowed. It may be assumed that the two main
parameters influencing the coating rheology of the polymer solutions
(polymer and solvent nature) have a reduced importance for water-based
PSA dispersions. More important seem to be the dispersion nature and the
inhomogeneous character of the liquid, dispersed adhesive.
The flow of a dispersion depends on the concentration and
characteristics of the dispersed solid particles (particle size and particle
size distribution). The correlation between solids content and viscosity
includes the particle size and particle size distribution. On the other hand,
the inhomogeneous, anisotropic character of the adhesive layer persists after
the coating (wetting out and drying) of the liquid layer. The final water
content of the dried, coated adhesive (residual humidity) depends on the
humidity of the environment (i.e., it reaches an equilibrium value). The dried
dispersion-based PSAs will never have the rheology of the bulk adhesive (the
adhesive from the polymer particles) but remain in an intermediate state
between the rheology of the dispersed system and that of the base polymers.
Furthermore the wetting additives left in the adhesive layer also
contribute to the composite character of the PSA layer. Therefore it can be
assumed that the rheology of the water-based PSA layer is that of a com-
posite material (i.e., polymer particles surrounded by a layer of additives).
From experimental practice it is known that the rheology of water-based
PSAs is shear sensitive (i.e., dynamic wetting needs a certain level of coating
speed and a certain level of shear on the metering/coating device). The most
important rheological parameter of a water-based PSA dispersion is its
viscosity. The viscosity influences the coatability, the film-forming char-
acteristics, and the properties of the adhesive layer thus formed.

Viscosity of PSA Dispersions


In order to be transfer coated onto the face stock (or release liner) the
adhesive dispersion must wet-out the solid surface. Furthermore, the coated
adhesive has to permit the withdrawal of the carrier liquid (water), and the
coalescence of the solid, dispersed particles (film forming). The wet-out of
the adhesive and the film building process are strongly influenced by the
viscosity of water-based PSAs. On the other hand, the coating technology
(coating machines and coating speed) is viscosity dependent too. Coating
devices are very sensitive towards viscosity, thus the coating versatility (i.e.,
the ability of an adhesive to be used on different coating machines) depends
on the viscosity. The coating speed contributes to the wet-out process.
Excessive speed may produce foam or adhesive build-up on the machine.
46 Chapter 2

Both phenomena are a function of the viscosity of the adhesive. The


adhesive’s flow on the metering device (i.e., the viscosity) also influences the
coating weight of the PSAs. Because of the lower level of interactions
between the solute and solvent, generally the viscosity of the dispersions is
lower than that of true solutions.

Wet-Out of Water-Based Adhesives


As stated in [99], of the technological properties of dispersed adhesive systems,
wetting ability is the most critical performance characteristic for their
coating. As will be shown later, the wet-out is a function of the quality of
the face stock and/or release liner. Water-based PSAs are coated mainly by
transfer, that is, coated onto the release liner and subsequently transferred
onto the face stock. The surface quality of the release liner is a function of
the siliconizing technology (condensation/addition; solventless/solvent-
based/water-based silicones). The fast curing rate at low temperatures (ca.
5 sec at 110 C) of the addition-based systems allows mild coating conditions
for the paper (better paper/moisture equilibrium) [107]. A decade ago,
solventless systems were developed, permitting better release properties and
economical advantages. Today, they represent about 40% of the world
usage in self-adhesive applications [55]. These systems are based on very low
molecular weight polymers, which can only be applied to very high grade
papers in terms of pore sealing and smoothness of the surface, in addition
to which special (better) wetting properties of the water-based adhesives
are required. Generally, suitable water-based PSAs have to display good
wetting on the worst liners (i.e., on siliconized liners) and especially on
solventless silicone-coated release liners.
Theoretical Basis of Wetting Behavior. When in its liquid state, the
adhesive must be able to uniformly wet the web. The amount of work
expended under reversible or equilibrium conditions to disrupt the interface,
i.e., to coat it, is a function of the surface energy of the latex. Equilibrium
spreading occurs when the surface energy of the substrate ySV is higher than
the surface tension of the latex yLV and the internal surface tension between
latex and substrate surface ySL, as summarized by the following formula and
illustrated in Fig. 2.6:

SV ¼ LV cos þ SL ð2:39Þ

where is the contact angle. Thus for a given substrate, wetting is facilitated
if the surface tension of the latex or/and the internal surface tension is
minimized. Therefore the surface tension of the adhesive has to be decreased.
As typical surface tension values for wetting out, 42  105 mN/m, to
Rheology of Pressure-Sensitive Adhesives 47

Figure 2.6 The contact angle of a high surface tension liquid on a nonpolar
surface.

46  105 mN/m are required [108]. Wetting out on silicone coated release
is more difficult. Silicones have a surface tension of about 20 mN/m. The
surface tension of water-based dispersions depends on their formulation. For
instance the surface tension of SBR lattices is lower than that of acrylates.
Industrial coating requires not only wetting but also the stability of the
coated liquid layer. Unfortunately, this liquid layer formed by wetting the
surface is not stable, it suffers dewetting. Two types of dewetting phenomena
are possible, namely retraction of the latex coating from the edge and surface
cracking.

Static/Dynamic Wet-Out. In industrial practice the wetting failure of


greatest concern is the dewetting of the adhesive from the primary web,
between the gravure coating station and the drying tunnel. In the process of
latex coating at high speed the dynamic aspect of the wetting/dewetting is
also critical.
The real wetting/dewetting behavior of the latex cannot be approxi-
mated in the laboratory because of shearing of the dispersion on the coating
(metering) roll, which improves wetting, and the competition between
dewetting and drying imposed coalescence. Therefore dewetting on the
48 Chapter 2

coating machine is diminished by shearing of the dispersion layer and


coalescence of the adhesive. Industrial wetting is always better than on the
laboratory scale.
The surface tension of the latex and its contact angle influence the
wetting/dewetting process [109,110]. Dewetting problems can be overcome
by reducing both the surface tension and the contact angle of the latex. In
order to evaluate the coatability both parameters (surface tension and
coating angle) have to be measured. Generally the contact angle of the latex
is plotted against the surface tension of the latex for different surfactant
systems. The best wet-out is obtained for the lowest contact angle and
surface tension. Other factors affecting wetting/dewetting are the viscosity
and the density of the liquid adhesive.
Surface Tension (Static/Dynamic Values). The influence of the lower
surface tension values is especially noticeable, when trying to coat onto low
surface energy substrates. If the surface energy of the web is too high, and
the formula is not thickened, defects show up in the film (e.g., cratering or
pinholes). Reducing the surface tension of the latex helps to get good
wetting. On the other hand, the coated liquid layer tends to retract. In order
to avoid film retraction surfactants have to migrate rapidly to the interface
between the adhesive and the release liner so as to avoid dewetting. Thus for
PSA applications where the coating process contantly creates new interfaces,
a low static surface tension and a low dynamic surface tension are needed.
Generally the surface tension is a function of the following parameters,
namely the age of the latex [111], the polymerization yield [112], the solids
content of the latex [113], the temperature [113], the viscosity [113], the
particle size and particle size distribution [114]. The influence of the viscosity
on the surface tension can be determined by means of surface amplitude
decay through the Orchard [115] equation for leveling:
Z
lnðAo =A1 Þ ¼ ð164 =3 4 Þ 0  h31 dt= ð2:40Þ

where Ax denotes the surface amplitude at any given time,  denotes


effective surface tension, h is the mean film thickness, l is the wave length,
and  is the dynamic viscosity. From the above relationship it may be
concluded that the surface tension  s is correlated to the viscosity:

s ¼ f ðÞ ð2:41Þ

Contact Angle. A low contact angle enhances the wetting out. For
the evaluation of the coatability the contact angle needs to be measured.
Unfortunately measuring of the contact angle faces some difficulties as the
Rheology of Pressure-Sensitive Adhesives 49

contact angle depends on the surface tension and on the smoothness of the
surface; furthermore the contact angle is time dependent (i.e., there is a
wetting and a dewetting angle) and the contact angle also depends on the
dimension of the adhesive drop.

Influence of Viscosity on Wetting


Dewetting may be reduced by increasing the standing ability of the liquid
adhesive layer. The mobility of the liquid layer is inversely proportional to
its viscosity and density. This is confirmed by thickening and incorporating
fillers into the water-based formulations. Only slight density differences may
be obtained, hence wetting out will depend mainly on the viscosity. Highly
viscous aqueous dispersions are characterized by good wetting properties,
and generally the increase of the viscosity is a method to improve the
wet-out. Thus, one can write:

Wetting out ¼ f ðviscosityÞ ð2:42Þ

For aqueous solutions of surface active agents Kurzendörfer [116]


formulated a special dependence for the amount of rest liquid G (i.e., liquid
which does not dewet) and dynamic viscosity :

G ¼ ð8=3Þr    h3=2  ð0:5 0:5 =g0:5 Þ  ðtw  td Þ0:5 ð2:43Þ

where r is the average cylinder radius (for the test equipment used for
vertical immersion-wetting out tests), h is the pull-out length of the cylinder,
 is the specific weight of the liquid, tw is the pull-out time of the cylinder, td
is the dewetting time of the liquid, and g is the gravity constant. Thus,

G ¼ C1  0:5  d 0:5 ðtw  td Þ0:5 ð2:44Þ

This correlation confirms the dependence of the wet-out on the


viscosity and density of the latex. As will be shown in Section 1.4 of
Chap. 3, aqueous dispersions are characterized by a minimum film-forming
temperature. This temperature depends on the viscosity of the bulk polymer
in the interior of the emulsion particles.
Influence of Viscosity on Coating Versatility. As discussed in [100] the
wetting out and viscosity are the main parameters for the coatability of
water-based PSAs. Wetting out is strongly influenced by surface tension,
viscosity depends on the diluting ability of the system. Coating versatility
refers to the ability of water-based PSAs to be coated by different coating
techniques. Generally, coating technology is a function of the chemical
50 Chapter 2

composition, physical state of the adhesive, product class, and product


build-up. As discussed in [117] coating technology and a particular coating
device can require a certain viscosity and solids content and influence the
formulation. As listed [118] the viscosity range of water-based PSAs covers
150–9000 mPas. The application equipment for water-based PSAs differs
from that used for solvent-based formulations. Generally it works with low
viscosity. In Europe two methods dominate the coating technology, namely
the metering bar system and the reverse gravure. The metering bar system
(Meyer bar) is used mainly for tapes and protective films. Reverse gravure
is recommended for labels. Slot die coating was developed for coating of
water-based dispersions also because of the very broad range (0.1–10 Pas)
of viscosities which can be used [119]. For coating by reverse gravure the
high solids content (64%) has to be reduced. For specific viscosity ranges
different coating devices are used. Reverse roll coating by rotogravure needs
low viscosities of 150–400 mPas (17–30 sec Ford Cup 4; 17–23 sec with
DIN Cup 4 mm [120]); thickening makes this interval broader. Reverse roll
coaters are suggested up to a viscosity of 40 Pas [121]. As stated in [99]
water-based adhesive systems may require a different gravure set up of the
coaters. The manufacturing versatility of selected coating methods is
discussed in detail in [38, 100]. The base viscosity of the dispersion is a
function of the polymerization recipe. For instance, water-based acrylics
are synthesized with viscosities of around 100 to 500 mPas. Thus they are
adequate for gravure and reverse gravure coating [122]. Adjusting the
viscosity of an aqueous dispersion requires knowledge of the thickening/
diluting response of the dispersion. Ideally the adhesive should allow large
viscosity changes without important variation in solids content and
mechanical stability. Another problem concerns the shear sensitivity of
the viscosity. Because of the shear sensitivity of the adhesive on the machine
the apparent viscosity in the machine differs from the viscosity measured
under laboratory conditions. For water-based dispersions a high thixotropy
ratio is required. On the other hand, dispersions with different viscosities
may be coated directly or by transfer. The coating technology (transfer or
direct) influences the anchorage and migration of the adhesive and therefore
the end-use properties. Therefore an indirect influence of the viscosity on
these properties must be taken into account.

Influence of Viscosity on Coating Weight. As discussed in [38]


different coating methods and coating weights are used for the main PSP
(e.g., label, tape, protective films etc.). For labels coating weight values of 10–
30 g/m2, for tapes 5–40 g/m2, and for protective films 0.5–15 g/m2 are applied.
It is evident that the difficulty in achieving a standard tolerance value of the
coating weight depends on its absolute value. To keep the coating weight
Rheology of Pressure-Sensitive Adhesives 51

tolerance in the usual range within a maximum of 5% the viscosity of the
emulsion must be kept constant. For a PSA dispersion on a reverse gravure
coater it is ideal to have a high solid content with a low viscosity. A low
viscosity implies a high coating weight. The minimum viscosity depends
on the surface tension of the adhesive too. In practice the coating weight
depends on several parameters such as the viscosity, running speed, and
solids content. Viscosity is a function of the solids content; on the other hand
it influences the coating speed. The interdependence of these parameters,
and their influence on the coating weight is illustrated in Fig. 2.7.
As can be seen from Fig. 2.7 the interdependence of the coating
parameters (running speed, solids content, and viscosity) is not linear. The
coating weight increases with the solids content, with the running speed, and
with decrease of the viscosity up to a maximum level, then decreases.
During the coating process disturbing phenomena may occur such as
adhesive build-up on the machine, coagulum build-up on the machine, and
in the coated layer, turbulences in the coated adhesive layer and changes
in the viscosity and foam formation. These phenomena depend on the
rheology of the diluted system. Because of their practical importance, they
will be discussed in Chap. 9.

Figure 2.7 The interdependence of the coating weight, the solid content, the
viscosity, and the coating speed of PSA.
52 Chapter 2

Factors Influencing the Viscosity of Aqueous Dispersions. The


following factors influence the viscosity of aqueous dispersions [116],
namely: temperature; solids content; pH; dispersing agents, particle form,
size, and size distribution; viscosity of the dispersing media; and shear forces.
In general, the viscosity decreases with increasing temperature. The
increase of the solids content implies an increase of the viscosity. The
dependence of the viscosity on the pH is more complex, but generally it can
be assumed that viscosity increases with increasing pH. The dependence
of the viscosity on the dispersing agent is a function of the nature and
concentration of the dispersing agent. A broad particle size distribution
imparts low viscosity due to the mobility of the particles. Generally water-
based dispersions contain a lot of different additives, which influence the
viscosity of the dispersing media. Shear forces produce in most cases a
pronounced decrease of the viscosity (shear thinning). Too high a shear
force may destroy the system producing coagulum. Too low an emulsion
viscosity can cause wetting problems, foam problems, or adhesive bleed-
through when the emulsion is directly coated onto the face stock. The
emulsion viscosity can also be controlled by either controlling the emulsion
particle size, or by subsequent thickening/diluting.

3 RHEOLOGY OF THE PRESSURE-SENSITIVE


LAMINATE

Pressure-sensitive adhesives are mostly used in a coated, bonded state


enclosed within a laminate. Generally the adhesive acts as an intermediate
layer between face stock and release liner, and finally, between face stock
and substrate. Because of the relatively small thickness of the adhesive layer
(Fig. 2.8) its flow properties are strongly influenced by the properties of the
delimiting solid surfaces. On the other hand, at least one component of the
laminate (the face stock), but mostly both (face stock and release liner or
substrate) are flexible, soft materials allowing no uniform distribution of the
applied forces, and undergoing a pronounced deformation during convert-
ing or end-use. Thus one can suppose that the rheology of the bonded PSAs
will be strongly influenced by the components, structure, and manufacture
of the laminate.
Pressure-sensitive adhesives act like permanently liquid elastomeric
layers enclosed between two thin, flexible, soft but solid surfaces. The flow
of the adhesive film is strongly influenced by the adhesive/surface
interactions, and by the stress/temperature loading transmitted by and
through these delimiting surfaces. The adhesive/face stock, adhesive/release
liner, and adhesive/substrate interactions depend on the adhesive nature and
Rheology of Pressure-Sensitive Adhesives 53

Figure 2.8 Geometry of the PSA paper laminate.

on the characteristics of the solid components. One can conclude that the
rheology of the adhesive in the laminate depends on its thickness, and the
dimensions on the continuous/discontinuous nature of the laminate com-
ponents, and on the laminate build-up. It is relatively easy to determine the
rheological properties of the adhesive or of the materials used as face stock
or release liner, but there is no means for direct measurement of the
rheology of the pressure-sensitive laminate. Practically, one can investigate
the adhesive, converting, and end-use properties of the coated (bonded)
PSAs, and thus indirectly evaluate the rheology of the bonded PSAs. Both
the uncoated adhesive and the solid components influence the rheology of
the coated adhesive.
In Section 2 the most important features of the rheology of the
adhesive are summarized. The influence of the solid laminate components
on the rheology of the bonded adhesive will be described indirectly through
the most important adhesive properties (i.e., tack, peel adhesion, and shear).
Rheology depends both on the nature of a fluid and on its flow conditions;
these are a function of the relative volumes of the fluid and, in the special
case of coated adhesive, on its coating weight. Therefore before examining
the influence of the solid state components on the adhesive properties, the
influence of the coating weight needs to be considered first.
54 Chapter 2

3.1 Influence of the Liquid Components of the Laminate


Earlier the rheology of uncoated PSAs was examined. The conclusions of
this chapter may be taken as valid for the liquid component (adhesive) of
the laminate. However, these considerations must be completed by the
examination of the influence of the adhesive coating weight on the rheology
of pressure-sensitive laminates.
The importance of the flow of the adhesive in achieving pressure-
sensitive properties has been outlined. Flow properties depend on the flow
conditions, that is, fluid volume, rate, and interfaces. Pressure-sensitive
adhesives are thin-layer coatings, where surface roughness and fluid volume
may have a pronounced influence on the motion of the adhesive. Surface
influences may be at least partially avoided by increasing the distance of the
fluid in motion from the solid surface (i.e., by increasing the coating weight
of the adhesive). Changes of the coating weight will influence the adhesive
and converting properties of the PSAs.

Influence of Coating Weight on Adhesive Properties


The main adhesive properties of PSAs may be described as tack, peel, and
shear, and each one of these properties depends in a different manner on the
coating weight. As is known from labeling practice, a certain coating weight
is necessary in order to obtain a measurable tack. Different tack mea-
surement methods show a different sensitivity towards the coating weight.
On the other hand the nature of the solid components of the label also
influences the relation tack/coating weight. A detailed investigation of RB
tack dependence on the adhesive and carrier nature and thickness was
carried out in [27]. As demonstrated RB tack values increase with the
increase of the coating weight and the decrease of the film carrier thickness.
Such dependence is a function of the adhesive nature too.
For elastic materials the energy flow to the crack tip per unit area of
crack extension is given by the following correlation [123]:

G ¼ K12 =2E ð2:45Þ

where the parameter K1 is called the mode one stress intensity factor. For
viscoelastic materials, in contrast to elastic materials, K1 depends on the
history of contact and on the details of the crack opening and closing
process [124]. According to Schapery [125] in this case the rate of energy
flow to the crack tip is given by:

1
GðtÞ ¼ CðtÞK12 ðtÞ ð2:46Þ
2
Rheology of Pressure-Sensitive Adhesives 55

This equation is similar to the elastic criterion [Eq. (2.45)] except that the
time dependent C(t) replaces the constant E, thereby making G a dynamic
quantity.
The history of contact and the details of crack opening have to be
taken into account by measuring the peel too. Dwell time (or theal) depends
on the application conditions. The coating weight also influences peel
adhesion. Generally the peel increases with increasing coating weight. This
is a nonlinear dependence, and its relation depends on the substrate to be
adhered to [126,127]. It was recognized in the first years of the development
of PSAs that controlling the coating weight allows modification of the
permanent/removable character of the adhesive. A decrease of the coating
weight from 20–25 g/m2 to 15–20 g/m2 can yield a removable label [58]. The
use of less than 7 g/m2 of dry adhesive for labels removable from paper was
developed [66]. Coating weight reduction is the main method to ensure
removability of the slightly crosslinked adhesives used for protective films
[44]. For acrylic PSAs used for common protective films, coating weight
values of 1.7–12.0 g/m2 are used. As discussed in [44] for an ideal adhesive
the adhesion should not depend on the coating weight. On the other hand
the peel value attains a measurable value only above over a certain coating
weight. Contact build-up needs adhesive flow. For given application
conditions, the surface roughness, carrier deformability, and adhesive cold
flow are the main parameters of contact build-up. Therefore the dependence
of the peel force on the coating weight can be described by a simplified plot,
with the linear part of the dependence shifted over the critical coating weight
value:

P ¼ Cw  ð2:47Þ

where  ¼ f (adhesive, carrier, substrate, time, temperature) and is a


conformation factor, which depends on the parameters influencing the
contact build-up between adhesive and substrate surface and the so-called
free flow region of the PSA. That means that for too low coating weight or
too rough surface no adhesion builds up. It is evident that rough, nonpolar
surfaces and harder adhesives on rigid nonconformable carrier materials
lead to higher values of ; in such cases the critical coating weight is higher.
Like the tack dependence on the coating weight, the dependence of the peel
is strongly influenced by the nature and geometry (surface and bulk
characteristics of the carrier material, see Section 3.2 also). As stated in [44]
the resulting peel is a complex function of carrier deformation ("c ):

P ¼ f ð"c ="c Þ ð2:48Þ


56 Chapter 2

where  and  are exponents that account for the influence of the deformation
work of the peel force. For thick mechanically resistant or elastical materials
the value of  is low. As seen from the relation (2.48) both  and depend
on the adhesive, carrier, substrate, time, and temperature. For a given
application and the same PSP (same adhesive, coating weight, and carrier)
on a given substrate  and can be considered as parameters that depend
only on the time and temperature. Therefore the peel value depends on the
application conditions only, i.e., laminating pressure, time, and temperature.
Thus for the application of classic PSPs (tapes and labels) the intrinsic
adhesive properties of the product (regulated by the adhesive nature and
coating weight) are more important for peel build-up. For protective films
the application conditions are determinant. It should be noted that contact
build-up with the adherent depends on the adhesive geometry too. Thus, as
demonstrated in [45] special crosslinked adhesives coated onto protective
films with different coating devices (e.g., gravure cylinder line 30, line 40, and
Meyer bar) and with different coating weight, give the same peel value.
Shear resistance tends to be inversely proportional to the coating
weight [56]. It is to be mentioned that because of the roughness of the solid
components a minimum coating weight may be required in order to achieve
a good shear strength. For special applications (e.g., packaging, splicing,
or mounting tapes) the high speed application and conformability of the
product require a thick adhesive layer (50–90 mm) [127]. In order to achieve
high shear values such compositions have to be filled or crosslinked. It
should be mentioned that the coating weight influences the shear test also.
A suitable test for the measurement of shear properties of an adhesive is the
thick adherend shear test (TAST) since it is conducted using a tensile testing
machine [128]. However, finite element analysis shows that the stress
distribution in the thick adherend shear test is not pure shear. The values of
maximum shear strain gained from the thick adherend shear test are not
comparable to those obtained from tests in pure shear. The role of the
coating weight on the energy dissipation will be discussed in Chap. 6.

Influence of Coating Weight on Converting Properties


The coating weight strongly influences the convertability of the laminate. As
discussed in detail in [63] printing related product characteristics (shrinkage,
lay flat, curling, stiffness, and wrinkle build-up), confectioning properties
(cutting, die cutting, perforating, embossing, folding, and winding), and
dispensing/labeling properties are coating weight dependent. This depen-
dence will be examined in Chap. 7. Here, the influence of the coating
weight on the cold flow of the adhesive will be discussed. The cold flow of
the adhesive in laminates is hindered by the marginal interaction of the
Rheology of Pressure-Sensitive Adhesives 57

adhesive with the ‘‘walls’’ of the laminate. It depends on the wetting and
anchorage of the adhesive. Because of the local character of this interaction
the middle layer displays free flow. For an ideal adhesive, cold flow should
be maximum during bonding and minimum during debonding. That means
that high loss modulus (peak) at the practical frequency of bonding helps
obtain adequate bonding. On the other hand at the highest debonding
frequency the elastic behavior of the adhesive (storage modulus) is required.
Cutting, slitting, and die cutting are high frequency operations.
The surface properties (roughness, porosity, and surface tension)
influence the wetting out, but there exists an indirect influence of the
adhesive on converting properties through the coating weight as well. It is
evident that the thickness of the free flowing layer depends on the coating
weight and on the thickness of the anchored layer. The anchored layer
depends on the coating technology and on the nature of the solid laminate
components; the free flowing layer depends on the coating weight (Cw) and
coating technology:

Cold flow ¼ f ðfree flowÞ ¼ f ðCw , Coating technologyÞ ð2:49Þ

Therefore, in a first approximation, the cuttability depends on the coating


weight, as follows:

Cuttability ¼ f ðCoating technology=Cw Þ ð2:50Þ

Figure 2.9 illustrates the dependence of the cuttability on the coating


weight (i.e., the cuttability decreases with increasing coating weight).
The most important factors influencing the coating weight are the face
stock material, the substrate to be adhered to, the permanent/removable
character of the adhesive, and the coating conditions. As discussed in [27],
the parameter in Eq. (2.47) describing the dependence of the peel on the
coating weight is a function of the adhesive, carrier, substrate, and time/
temperature. Generally it may be stated that the optimum coating weight
value depends on the chemical composition and end-use of the PSA. The
adhesive performance also influences the coating weight to be used. The
coating weight of selected PSPs is listed in [129]. Solvent-based acrylics are
coated for different product classes with the following coating weight values:
5–10 g/m2 for protective films; 25–40 g/m2 for film-based tapes; 50–100 g/m2
for foam tapes; and 20–30 g/m2 for labels and tags. For labels peel is
regulated by formulation, not by coating weight. Labels can be considered
a special case of PSPs where coating weights are relatively high (above the
critical domain). On the other hand for the same product class various
58 Chapter 2

Figure 2.9 Dependence of the cuttability on the coating weight. Cuttability of soft
PVC film labels: 1) PSA with low PE-peel, medium tack, high cohesion; 2) PSA with
high PE-peel, low tack, low cohesion; 3) PSA with low peel, very low tack, high
cohesion; 4) PSA with high peel, high tack, low cohesion.

adhesive formulations (with different base elastomer and physical status) use
different coating weights. Protective films coated with solvent-based acrylics
or rubber resin formulations display a lower coating weight than dispersion
coated ones [63]. All these parameters and their influence on the coating
weight will be examined in Chap. 6.

3.2 Influence of the Solid Components of the Laminate


Some PSAs are self-sustaining (e.g., sealant tapes, mounting tapes, self-
adhesive protective films etc.). Such formulations do not need a solid state
carrier material. Their adhesive properties depend on the rheology of the
nonreinforced PSA. Most PSPs possess a carrier material (e.g., face stock
for labels). Both, the bulk and surface properties of the solid state carrier
material (for labels: face stock) influence the adhesive, converting, and
end-use properties of the pressure-sensitive products. The solid laminate
components affect the distribution of the applied stresses on the adhesive
(Fig. 2.10) and the flow of the liquid adhesive in the laminate.
Rheology of Pressure-Sensitive Adhesives 59

Figure 2.10 The influence of the solid state laminating components (face stock
and release liner) on the active and passive forces in the adhesive’s flow: FA; active
forces, compression, and impact stress on the laminate: Fp1; passive forces, resistance
against PSA flow, given by the adhesive/face stock interaction: Fp2; resistance against
PSA flow, given by adhesive/release liner interaction.

The interaction between the PSAs and the delimiting surfaces acts like a
brake or a resistance to flow.
The stress distribution depends on the elasticity, hardness, and flexi-
bility of the solid component (i.e., on its bulk properties) and dimensions.
Control of the flow is a function of the quality of the delimiting surfaces, that
is of the surface properties of the solid components. As a solid component
influencing the adhesive properties the adherent (i.e., the substrate) should
be considered as well. In order to elucidate the influence of each one of the
solid components of the PSA label (applied or not) a separate discussion of
these components has to be carried out. First the role of the face stock in
the design of the label will be examined in order to clarify its influence
on the rheology of the adhesive. The influence of the face stock material
on the rheological characteristics of the adhesive (modulus and viscosity)
will be discussed while characterizing the face stock material by its own
rheological properties (modulus). The best approach to describe these
phenomena is their indirect examination through the adhesive and
converting characteristics.
The properties of self-adhesive labelstock are primarily concerned
with the printability of the face stock, the conversion characteristics of the
60 Chapter 2

Figure 2.11 Face stock influence on the adhesive and converting properties. The
influence of the stiffness, elasticity, roughness, porosity, and surface tension on the
face stock directly and through coating weight on the adhesive and converting
properties of the label material.

laminate, and end-use adhesive performance. All these characteristics depend


on the face stock material used. As shown in Fig. 2.11, the face stock material
influences the properties of the PSA label through its bulk and surface
properties.
The bulk properties of the face stock material influence the adhesive
and converting properties directly and indirectly. The indirect influence is
exerted by the coating weight. The role of the porosity in the migration
is well known. If the adhesive wets the adherent (face stock) surface, its
performance is largely influenced by the bulk properties of the face stock
material.
The enormous importance of the surface quality of the solid com-
ponents of the PSA laminate on the properties of the PSA is illustrated by the
recent rheological studies concerning high and low surface energy substrates.
In the case of low surface energy substrates the rheological approach of
the properties is often not valid. The plasticity/elasticity balance of the face
stock material, its stiffness, and its capability to absorb energy, influence
the adhesive and converting properties of the pressure-sensitive laminate.

Influence of the Bulk Characteristics of Carrier Material on the


Adhesive Properties of the Pressure-Sensitive Laminate
The dependence of the most important adhesive characteristics (peel,
tack, and shear) on the bulk properties of the face stock material will be
Rheology of Pressure-Sensitive Adhesives 61

investigated. It should be pointed out that some of the tests of the adhesive
properties of PSAs involve deformation of the face stock material (e.g., tack,
peel); others (e.g., shear) are characterized by the deformation of the pure
adhesive layer only. Thus the bulk properties of the solid components of
the pressure-sensitive laminate influence more the tack and the peel of the
adhesive. Peel adhesion depends on the face stock used as well as on testing
conditions [44]. This strong dependence is due to the simultaneous
deformation of the face stock material and adhesive during peeling. Also,
some tack measurements induce face stock straining. In fact most of the tack
measurements may be considered as peel, shear, or tensile strength
measurements, and therefore they will be discussed in Chap. 6.
Face Stock Material and Peel Adhesion. From an energy point of
view debonding is associated with energy dissipation. Energy dissipation
during debonding is dependent on such parameters as interfacial adhesion,
polymer rheology, rate of peel, temperature, angle of peel, web stiffness, and
coating weight. Interfacial adhesion and web stiffness are characteristics of
the face stock material. They also influence the angle of the peel. On the
other hand, the peel rate, especially in the starting phase or onset of peeling,
or during manual peel, strongly depends on the plasticity and elasticity of
the face stock material (Fig. 2.12).
Figure 2.12 presents the dependence of the peel on the face stock
material for tackified PSAs dispersions with different tackifier levels. As can
be seen from Fig. 2.12, the value of the peel is quite different for a soft,
plastic face stock material. Adhesive formulations display quite different
peel adhesion values when coated on different face stock materials.
Thus face stock flexibility and plasticity are the main bulk
characteristics influencing the rheology and the peel. From release force
measurements, it is well known that the peel force from a silicone release
liner depends on the peel angle p and the peeling rate :
Peel force ¼ f ðp , Þ ð2:51Þ
The peel angle is really a function of the flexibility fFS of the face stock:
p ¼ f ðflexibilityÞ ð2:52Þ
Thus:

Peel force ¼ f ðp , fFS Þ ð2:53Þ

The peel angle is normalized (90 or 180 ) but the real peel angle depends
on the flexibility of the material (Fig. 2.13), and its deviation from the
theoretical angle is different for 90 and 180 peel tests.
62 Chapter 2

Figure 2.12 Dependence of 180 peel adhesion force on the nature of the face
stock material. 1) PVC; 2) aluminum.

Figure 2.13 Theoretical and real peel angle during 180 peel measurement.
Rheology of Pressure-Sensitive Adhesives 63

For peel testing at 80 the delaminated length A of the peeled bond is
given as a function of the modulus of the face stock Ef, the modulus of the
adhesive Ea, the thickness of the adhesive coating ha, and the thickness of
the face stock hf [130]:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
A 4
ðEf =Ea Þ  h2a  hf ð2:54Þ

A more detailed description of the influence of the face stock elasticity


(stiffness) on the peel is given in Chap. 6. Stress transfer to the adhesive and
from the adhesive is influenced by its plasticity. The experimental data from
Fig. 2.14 illustrate this behavior.
Both phenomena—face stock stiffness and elasticity (resistance to
elongation)—were studied by Gent and Kaang [68]. When adhesive labels
(or tapes) are pulled away from a rigid substrate, the force required depends
upon both the strength of adhesion and the resistance of the label to
stretching [68]. The pull-off force F may be correlated with the work of
detachment Ga and the effective tensile modulus E of the tape and its

Figure 2.14 The influence of the elasticity/plasticity of the face stock on the peel.
A paper label material (A) is being peeled through a 180 angle from a substrate,
while the peel stress is being transmitted via different materials (B). B1: paper (peel
force 7.4 N/25 mm); B2: PET (7.7); B3: soft PVC (6.9); B4: LDPE (6.9).
64 Chapter 2

thickness t. The pull force is found to be neither proportional to Ga as might


at first be expected, nor is it proportional to (EGa)1/2 as found in many
linearly elastic systems. Instead it is found to be proportional to (EG3a )1/2,
a result which emerges directly from the analysis as a consequence of the
particular relation between the force F and the corresponding elastic
displacement of the label where no further debonding occurs. It is to be
noted that the values of detachment energy Ga obtained from pull-off
experiments are generally lower than those obtained from peeling
experiments. It was stated that:

F  ð8Ga =3Þ3=4 ðE  tÞ1=4 ð2:55Þ

That is,

F ¼ f ðt1=4 Þ ð2:56Þ

A further prediction of the theory is that the product F (where is the


detachment angle) will be independent of the stiffness of the tape (label).
Differences in Ga from pull-off and from 90 peel measurements are
attributed to additional energy losses (by peeling) due to the severe bending
deformations imposed on the label. A possible cause is nonlinear elastic
behavior of the label in tension. The effective tensile modulus E at small
strains is greater than at large ones. The simplifying assumption of linearly
elastic behavior is quite inadequate for film labels or for tapes which
undergo large deformation as well as plastic yield. The work to bend the
label away from the substrate may be important too. In some circumstances
this contribution can be both large and strongly dependent upon the
magnitude of the peel angle [131]. Later the shaft loaded blister test SLBT of
Wan and Mai [132] was used to determine the apparent strain energy release
rate (G), for a PSA tape in order to examine the influence of the solid state
carrier material [133]. Samples consisting of stacked films (either 1, 2, or 4
plies thick) were tested to determine if the backing’s tensile rigidity had any
effect on the measured G. The G values measured from the SLBT were
compared to values obtained using the pull-off test [68]. For high energy
aluminum substrates (i.e., for high adhesion values) results from a 90 peel
test were an order of magnitude greater than the SLBT, probably due to the
plastic deformation of the peel arm. A measure of the fracture toughness
of an adhesive bond is the applied or apparent strain energy release G. The
expressions used to calculate G are typically derived using linear elastic
fracture mechanics (LEFM), that is by assuming that the adherent and
adhesive are loaded within the elastic range, except locally at the crack tip,
Rheology of Pressure-Sensitive Adhesives 65

where plastic deformation occurs. In practice adhesives behave viscoelas-


tically or elastic-plastically, and under high stress yielding may occur. For
thin films the effects of plastic deformation and yielding may be significant
given their small load bearing capacity, associated with the small thickness,
and the relatively strong adhesion [134].
Energy transfer from the face stock to the adhesive depends on the
anchorage of the adhesive layer and on the existence of an energy absorbent
layer on the face stock materials. Both are surface dependent characteristics,
thus the surface of the face stock material will influence the peel of PSAs
(especially for removable and nonpolar adherent applications). Indirectly the
anchorage of the PSAs and the existence of different adhesive layers on the
surface or in the face stock material depend on the migration of the adhesive
or adhesive components in the face stock material. This phenomenon
depends on the surface and bulk properties of the face stock material.
Influence of the Substrate. The nature of the substrate influences the
peel adhesion, and thus the removability of the PSA label. Certain PSAs may
be classified (depending upon the substrate) as removable or permanent
adhesives. The most important parameters of the substrate that influence the
removability of the PSA label are its flexibility, its elasticity/plasticity
balance, and the anchorage of the adhesive onto the face stock. The peel does
not depend on the deformability of the adhesive, but on the flexibility of the
face stock and substrate [130].
The deformability of the adherents influences the stress distribution in
the bond [56]. If the adherents are not deformable, the adhesive would be
stressed uniformly. If the solid components of the joint are deformable the
stress distribution varies along the joint. Mathematically this phenomenon
may be described by a factor n, where n is the ratio between maximal and
mean stress ( max, ):
pffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
n ¼ max = ¼ =2 coth =2 ð2:57Þ

where  is given by the following correlation:

 ¼ ðG  l 2 Þ=ðEabÞ ð2:58Þ

where a is the thickness of the adherent, b is the thickness of the adhesive, G


is the modulus of the adhesive, E is the modulus of the adherent, and l is the
overlap length. The influence of the deformability of the adherents on the
stress distribution by peeling has also been demonstrated [135]. The adhesive
is under tension at the interface and under compression just behind it. The
tensile strength is a function of the stiffness of the combined adherents and
66 Chapter 2

adhesive, as well as of the thickness of the system. The compressive load


appears greatest at 180 peel between flexible and rigid adherents, less at 90
peel under the same circumstances, and the peel force is the lowest when
measured between two flexible adherents.

Face Stock Material and Shear Strength. There are no experimental


data available about the influence of the bulk properties of the face stock
material on the shear resistance. Differences in shear strength of PSAs coated
on different face stock materials are due to the surface of these materials.

Influence of the Face Stock on the Converting Properties of the


Pressure-Sensitive Laminate
The most important converting properties of pressure-sensitive labels are the
printability and the cuttability. The rheology of the adhesive strongly influ-
ences the cuttability of the label. Next the dependence of the cuttability on
the bulk properties of the face stock material will be analyzed, and in
particular the influence of the rheological characteristics of the face stock
material. During the die-cutting operation, the mechanical strength of
the face stock and release liner materials influences the distribution of
the stresses from the cutting edge to the adhesive layer. Generally, this
mechanical resistance is a function of the mechanical properties of the bulk
material, and of the dimensions of the face stock and release liner:

Rc ¼ f ðK1  K2 Þ ð2:59Þ

where Rc is the mechanical resistance of the face stock material, K1 denotes


material characteristics and K2 material dimensions. In some cases (sheet
material) a certain number of superposed laminate sheets are cut during the
same operation; thus the distribution of the force may be influenced by
laminate slip (Fig. 2.15). Therefore the laminate strength will depend on the
motion of the laminate layers (face stock and release liner respectively), with
respect to one another (i.e., on the surface characteristics of the laminate:
face stock, release liner) K3:

Rc ¼ f ðK1  K2  K3 Þ ð2:60Þ

where K3 is a surface-dependent factor influencing the laminate slip. As can


be seen from Fig. 2.16, the motion and cold flow of the adhesive will be a
function of stress magnitude and distribution, but smearing of the cutting
knife and of the cut (side, front) surface also depends on the porosity of this
surface.
Rheology of Pressure-Sensitive Adhesives 67

Figure 2.15 The motion of the laminate layers during cutting.

Figure 2.16 Flow and smearing of the adhesive during cutting. (A) through (E) are
the different steps during the cutting operation.
68 Chapter 2

Thus the cutting behavior and the cuttability C of the laminate will
depend on the mechanical properties of the bulk material (face stock, release
liner) K1, on the surface properties of these materials K3, and on the
dimensions (influencing the real value of the forces and the cut area) of the
solid components of the laminate K2:

C ¼ f ðK1  K2  K3 Þ ð2:61Þ

In order to identify the most important mechanical properties of the


bulk material which may influence the cuttability, the nature of the forces
and deformation during cutting have to be examined. Different kinds of
stresses are present in a material, as demonstrated by creasing and folding
paper board (Fig. 2.17) [136,137].
One can conclude that the same stresses (i.e., tensile, compression, and
shear) are working when cutting a pressure-sensitive laminate. In fact
creasing the face stock constitutes the first operation during cutting.
Material characteristics related to tensile, compression, or shear resistance
of the face stock material will influence the cuttability. The distribution of
the stresses, the uniformity of the force acting on the laminate, and the
compression stress character of the first cutting step are determined by the
rigidity of the face material (Fig. 2.18).

Figure 2.17 Stress distribution during creasing (cutting) of the face stock (label)
material. 1) Tensile, 2) compression, 3) shear stresses.
Rheology of Pressure-Sensitive Adhesives 69

Figure 2.18 Stress distribution during cutting as a function of the nature (rigidity)
of the solid components of the laminate.

The rigidity (stiffness) of the material is a function of its modulus. An


incompressible liquid PSA may support a compressive stress well. In the first
step of the cutting operation the force is acting perpendicularly on the whole
laminate. The situation changes only through the concentration of the local
stresses, and because of the weakness of the solid material, giving rise to the
flexure of the face and shearing of the adhesive. Hence the stiffness of the
solid laminate components influences the stress distribution and the creep
resistance of the adhesive. The value of the modulus of elasticity E of the
face stock material (or/and release liner) influences the cuttability of the
material. Thus:

C ¼ f ðEÞ ð2:62Þ

On the other hand, assuming that the cutting edge (knife) is acting on
a partially supported material, flexure of the material occurs first (i.e., the
flexural modulus of the material influences the cuttability). In a first
approximation the cuttability C depends on the tensile, compression,
flexural, and shear moduli (El, Ec, Ef, Esh) of the material and the value of
these moduli may or may not differ from one another, as a function of the
70 Chapter 2

materials used (e.g., paper, plastic, or metal) and of the practical conditions:
C ¼ f ðEi Þ ð2:63Þ

Cuttability also depends on the material dimensions. It should be


mentioned that the stress distribution dependence on the thickness of the
face stock material may vary according to the nature of the applied stress. It
may be stated that the cuttability improves with material thickness, that is:
C ¼ f ðEi , hÞ ð2:64Þ

where h is the material thickness and  is the exponent taking into account the
improvement of the cuttability by the increased cut surface for a voluminous
(thick) laminate component. For hygroscopic materials like paper, bulk
properties may depend on the environmental humidity. Because of the fiber-
like structure of the material, the surface texture may also change as a
function of the humidity. Surface texture influences the smoothness (friction)
of the surface and the smoothness influences the cuttability. The surface
friction of papers increases with increasing air humidity [138]. On the other
hand, because of the hygroscopic nature of the face paper, there is a con-
tinuous transfer of water between the adhesive and face material, influencing
the composite character of the adhesive and its rheology (water-based
adhesives). A more detailed analysis of the parameters influencing the cut-
tability will be presented in Chap. 7. Here the influence of the bulk properties
of the face stock material on the distribution of the stresses acting on the
laminate, and of the rheology of PSAs on the laminate will be discussed.

Influence of Surface Properties of the Carrier Material


Coatability is characterized by wetting out and anchorage of the coated
adhesive on the carrier material. Both are influenced by surface polarity
and porosity. Figure 2.11 showed that the surface properties of the face
stock material influence the adhesive and converting properties of the
pressure-sensitive laminate, either directly or indirectly. The direct influence
is based on the anchorage phenomenon at the surface of the face stock.
Indirect influences are reflected by the porosity related adhesive penetration
in the face stock and the changes in the coating weight caused by
penetration. Special problems are encountered with nonpolar plastics where
no chemical affinity exists between the polar PSA and the nonpolar,
nonporous carrier surface. For textured fiber-like porous surfaces adhesive
migration and breakthroughs may improve anchorage but denaturate the
top surface of the face stock and at the same time change the coating weight.
Fillers or surface-active agents built into the carrier material of protective
films or tapes can act as release enhancing agents, but they disturb the
Rheology of Pressure-Sensitive Adhesives 71

adhesive contact. Fillers can modify the relaxation spectra of the polymers
[139]. On the other hand for special PSPs (e.g., self-adhesive films) migration
of an adhesive component from the plastic carrier (e.g., polybutylene) due to
its incompatibility ensures the adhesive properties of the product [140].
Influence of Face Stock on Migration. It is difficult to store labelstock
for any length of time, due to possible bleeding and/or oozing problems. A
lot of problems exist with high gloss label material bleedthrough. Adhesive
bleedthrough often interferes with the printing. Migration (penetration or
bleeding) is due to the diffusion of the adhesive into and through the face
stock material. This phenomenon depends on the properties of the adhesive
and of the face stock material, on the coating technology, and on the
processing (storage, converting) conditions:

Migration ¼ f ðadhesive, face stock, coating technology, processingÞ


ð2:65Þ

Migration of the adhesive in the face stock material or of the additives from
the face stock material in the adhesive produces changes in the composition
of both materials and in their mechanical/rheological properties.
In many health care and other applications it is desirable to have a
PSA film coated onto a porous web. Hot-melt PSAs offer a relatively easy
route to coating PSAs onto fabric and nonwoven webs with a minimum of
adhesive penetration into the web. The penetration of the adhesive in the
porous web can be avoided with a release liner used as an intermediate web
surface (transfer coating) for some processes, but during storage of the
product diffusion in the porous web may occur. Evidently, migration
(penetration) of the adhesive into the face stock depends on the adhesive’s
nature, but also on the face stock porosity. The dimensions of the pores
strongly influence the penetration. It was demonstrated that pore diameters
of 0.01–0.05 mm allow good penetration [141]. Theoretically, penetration
under pressure and capillary penetration in paper are different phenomena
[142]. Penetration through pressure Pp would depend on the pressure p, the
radius of the pores r, and the viscosity :

Pp ¼ f ð p, r, Þ ð2:66Þ

Capillary penetration Pc (mainly water penetration) depends on the


humidity of the paper H and its pH value:

Pc ¼ f ðH, pHÞ ð2:67Þ


72 Chapter 2

Porosity influences positively the anchorage of the adhesive through


migration. Where pores, crevices, and capillaries are accessible in the surface
of the adherent the adhesive will penetrate to some extent, and so increase
adhesion. The capillary rise h is a function of the porosity [143]:

h ¼ k  LV cos =R ð2:68Þ

where R is the equivalent radius of the capillary, k is a constant, and  is the


density of the liquid; h reaches a maximum when:

1
LV ¼ ðc þ 1=bÞ ð2:69Þ
2

where  c is the surface tension of the liquid. One can conclude that the
dimensions (radius) of the pores, the wettability of the material (surface
tension), the humidity of the paper, and the pH of the paper are major factors
influencing the migration. This is also confirmed by the following correlation

dL=dt ¼ ð  cos  rÞ=ð4  LÞ ð2:70Þ

where dL/dt denotes the penetration rate,  is the viscosity of the liquid, r is
the capillary radius,  is the surface tension of the liquid, is the wetting
angle of the capillary wall, and L is the penetration length. Generally, the
penetration rate of the adhesive in the face stock depends on the capillary
radius (i.e., on the paper density) [144].
Influence of Face Stock Porosity on Coating Weight. It is well known in
the field of cold-seal adhesives that blocking is a function of the adhesive
coating weight and of the face material. For a porous web (paper) a coating
weight of about 6 g/m2 is recommended; nonporous face stock (film or
aluminum) needs no more than 3 g/m2. Similarly the dry coating weight (the
weight of the adhesive applied per unit surface area) can vary substantially
depending upon the porosity and irregularity of the face and of the substrate
surface to which the face material is to be adhered. For instance higher PSA
loadings are preferred for adhering porous, irregular ceramic tiles to porous
surfaces, while lower adhesive loadings are required to manufacture tapes,
films, and other articles from relatively nonporous, smooth surface materials
such as synthetic polymer films and sheets. When the adhesive is applied to
nonporous polymeric or metallic face materials, intended for adhesion to
nonporous polymeric or metallic surfaces, adhesive coating weights of about
7–70 g of dry adhesive per m2 of treated surface are generally adequate. A
good adhesion for tapes manufactured from continuous polymeric materials
can usually be achieved with dry adhesive coating weights of about 15–30 g
Rheology of Pressure-Sensitive Adhesives 73

of adhesive per m2, while coating weights of about 30–70 g/m2 of adhesive
are usually employed for the manufacture of paper-backed tapes, such as
masking tapes. In order to avoid a decrease of the coating weight due to the
porosity of the face stock material or to prevent plasticizer migration, it is
also possible to apply a barrier coating to the paper [145].
Influence of the Face Stock on the Adhesive Choice. Face stock
sensitivity towards PSAs requires an adequate choice of the adhesive system.
Shrinkage of soft PVC face stock requires the use of shrinkage resistant
adhesive systems [146]. To this end crosslinked acrylics are generally emplo-
yed. On the other hand, plasticizers in the vinyl film can have a pronounced
effect on the adhesive properties during the aging process [147]; plasticizer
resistant adhesives have to be selected. These examples show that the face
stock nature and its sensitivity towards certain adhesives may influence the
choice of the adhesive and thus the rheology of the adhesive layer.
Influence of the Face Stock on the Coating Technology. In the first
instance the mechanical and chemical characteristics of the face stock
materials determine the choice of the coating technology, the coating
process (direct coating/transfer coating), use of cold/hot adhesives (solvent-
based/hot-melt) or chemically aggressive or inert adhesive systems (solvent-
based/water-based), and coating machines (metering device, drying system,
web transport system, etc.). The influence of the face stock material on the
choice of the coating technology is illustrated in Fig. 2.19.
As can be seen in Fig. 2.19 the web is usually supported when making
paper-based laminates via transfer coating, using a gravure roll as a metering
device, and solvent-based, water-based, or hot-melt PSAs. For film coating a
Meyer bar and/or gravure roll is suggested, usually using water-based PSAs.
Bonding (or anchorage) supposes the coating of a substrate (or face
stock) material with the liquid adhesive, that is the spreading and wet-out of
the adhesive on the surface of the solid material. After wetting the surface
the adhesive penetrates into the cavities of the solid material and physical
and chemical contact points are formed. Both phenomena, bonding and
debonding, depend on the quality of the surface.
Wetting Out as a Function of the Face Stock Surface. Earlier the
impor-tance of the wetting out, and the parameters influencing it were
discussed, considering in particular the liquid component (adhesive). Next the
characteristics of the solid components influencing the wet-out are examined.
It was shown earlier that the bulk properties of the face stock materials on the
one side, and the surface properties on the other side are determinant for the
properties of the pressure-sensitive laminate. The surface quality influences
the wetting out of the adhesive. Wetting the solid surface by the adhesive is an
74 Chapter 2

Figure 2.19 The influence of the face stock material on the coating technology.
Dependence of the coating technique (direct/transfer), of the nature of the PSA
(water-based, solvent-based, hot-melt), of the coating device, and of the drying
technology used, on the nature of the face stock materials (paper/film).

initial necessary condition for the anchorage of this adhesive to the solid
surface (e.g., face stock material) [54], and is a function of the surface tension
of the solid and liquid. Liquids, and for that matter solids, at the boundary
between two phases (liquid/paper or liquid/solid) possess properties which
are different from those within the mass of the liquid itself. This difference
results from an imbalance between the attractive faces of the surface layer of
molecules and the molecules contained within the liquid (or solid) [148]. The
molecules on the surface layer compensate for charge imbalance by pulling
themselves close together, thus creating the phenomenon of surface tension.
In the most extreme case when the surface tension of the adhesive differs
greatly from the initial surface tension of the film or paper (face stock or liner)
fish eyes and cratering will result (when the face stock surface tension is less
than the adhesive surface tension).
To obtain optimal wet-out the surface tension of the adhesive must be
equal to or less than the critical surface tension of the face stock or liner
(a contact angle of 0 ) (Fig. 2.6). As can be seen in Fig. 2.6, a high surface
tension liquid beads up on nonpolar, low energy surfaces [109]. Table 2.2
Rheology of Pressure-Sensitive Adhesives 75

Table 2.2 Surface Energy Values of Common Face Stock/Substrate Materials

Material Surface Energy, mN/m References

Aluminum 45.0 109,148


Cellulose acetate 39.0 151
Cellulose triacetate 29.0 142–154
Glass 47.0 109,148
Iron 46.0 148
Polyamide 36.0 155
40.0–46.0 156
46.0 151
33.0–46 148
Polyamide (PA-6) 42.0–48.0 53,109
(PA-11) 43.0 151–154
Polybutadiene 32.0 151
Poly(butyl methacrylate) 26.0 154
Polycarbonate 46.0 154
Polydimethylsiloxane 14.1 100,152,153,
Poly(ethylene) 29.0–31.0 53,152–156
(HDPE) 33.0 52,151–156
(LDPE) 31.0 151
Poly(ethyleneterephthalate) 37.0–43.0 109, 157
P(EVAc) 33.0 151
Poly(hexafluoropropylene) 16.0 53,149
Poly(isobutylene) 27.0 151
Polyisoprene 31.0 151
Poly(methyl acrylate) 39.0 151
Poly(methyl methacrylate) 29.8 152,154
33.0–44.0 109,141
38.0 151
Poly(propylene) 28.0–30.0 53
29.0 151,155
30.0 127,151,151
31.0 148
Poly(styrene) 33.0–43.0 53
53,101, 152–154
33.0 156
30.0–35.0 151
33.0–35.0 141
Poly(tetrafluoroethylene) 18.5 35,151–153
Poly(trifluoroethylene) 22.0 151
Poly(vinyl acetate) 25.0 151
24.9 152–154
(continues)
76 Chapter 2

Table 2.2 Continued

Material Surface Energy, mN/m References

Poly(vinyl alcohol) 37.0 151


Poly(vinyl chloride) 35.0–41.0 53, 157
36.0 147
39.5 148
33.0–38.0 148
39.0–40.0 53,109
41.9 151
Hard 35.0–37.0 151
Soft 29.0–31 151
Poly(vinylidene chloride) 40.0 53
Silicones 20.0 152
18.0–22.0 153
Steel, stainless 45.0 109
Steel, chrome plated 60.0 109

lists the surface energies of materials commonly used as face stock and
substrate.
Situations can also arise where the adhesive is absorbed on the surface
of the face stock (e.g., film) by chemisorption, with a zero contact angle
(good wet-out). However, the mass of the liquid does not bond to this
chemisorbed layer [148].
Except for hot-melt PSAs (and some warm coated masticated rubber-
based formulations used for medical tapes [149]), PSAs are coated onto the
face stock (or release liner) as liquids. The wetting out of liquid PSAs
depends on the quality of the face stock surface, on the surface tension,
and on the roughness of the surface. The dependence of the wetting out on
the roughness is a general phenomenon. The wet-out time on cardboard
(for an aqueous dispersion) may vary between 0.1–0.4 sec, depending on the
roughness of the surface [142]. Moreover, Bikermann [150] stated that
adhesion is due to the inherent roughness of all surfaces. Molecular forces
of attraction cause an adhesive to wet and spread on the surface. But once
achieved the mechanical coupling between the adhesive and the inherently
rough surface is more than enough to account for the bond strength.

Influence of Carrier on the Anchorage of PSAs. Interfacial adhesion is


a prerequisite for bonding and debonding. Stable bonds are formed when
there is a strong interaction between the adhesive and the face stock (or
substrate). The different types of possible interactions include van der
Rheology of Pressure-Sensitive Adhesives 77

Waals’ forces, polar interactions, electrostatic interactions, and chemical


bonding. Effective bonding between the face material and the adhesive
is critical to the performance of PSAs; given an adhesive with internal
integrity, the adhesive properties exhibited by that system will be pro-
portional to the affinity of the adhesive for the face stock material [64].
A useful approach for evaluating the compatibility of the adhesive and face
stock is to measure the free energy of these surfaces. Theoreticians have
formulated adhesion models based on the thermodynamic or reversible
work of adhesion WA. This is the change of free energy when materials are
brought into contact, which represents the amount of energy released under
reversible or equilibrium conditions to disrupt the interface. The expression
is given by the following equation:

WA ¼ SV  LV  SL ð2:71Þ

In practical terms the face stock material and the adhesive should have
comparable surface energy properties.
Solvent-based PSAs consist of a polymer dissolved in a solvent. The
solvent-based PSA adhesion or anchorage on plastic films essentially
concerns the study of how polymers and solvents interact. In order to
explain the phenomena which occur during the coating of solvent-based
adhesives the nature of this interaction between polymers and solvents is to
be examined. Solvent molecules are attracted to other solvent molecules.
Similarly, polymer chains possess attractive forces within a chain, and from
one chain to another. A polymer dissolves in a solvent when the net effects
of the solvent/solvent and polymer/polymer attractive forces are reduced.
Hildebrandt [158] defined a solubility parameter  as a measure of the
internal attractive forces of the solvent, defined as heat of evaporation EV
per unit molecular volume VM;  is a measure of the energy density of the
solvents:

2 ¼ EV =VM ð2:72Þ

Hansen et al. [159,160] proposed an alternative to the single valued


solubility parameter, namely:

2 ¼ 2d þ 2p þ 2h ð2:73Þ

where d is the attractive molecular force from temporary dipole formations,


p is the attractive force from dipole/dipole interactions, and h is the
attractive force from hydrogen bonding. This concept allowed for the
weighted effects of the highly polar and hydrogen bonding solvents.
78 Chapter 2

Polymers, unlike solvents, have a range of acceptable solubility parameters


due to the effects of molecular weight distribution, areas of crystallinity,
different monomer units on the same chains, etc. Hansen [159,160]
developed a method for rating the solubility parameter ranges for polymers
using the three-component parameter system. One can then define the
solubility parameter range of a polymer in terms of a three-dimensional
volume. Then any solvent or solvent blend with a solubility parameter that
falls within the described volume will dissolve the polymer. For our
purposes, an adhesive/solvent blend that falls within the solubility parameter
area of a plastic film will develop good adhesion or anchorage. For
simplicity, one can make a good approximation of the adhesive/substrate or
face material interaction using a two-dimensional graph.

Carrier Surface and Removability. The nature of the face stock


influences the peel and thus the removability of a PSA label. Due to the strong
influence of the carrier material some PSAs may be classified (depending
upon the face stock) as removable or permanent adhesive. The main criteria
for remova-bility are tack, peel, cohesion, and anchorage. The most impor-
tant parameters of the face stock influencing the removability of the PSA
label are its flexibility and the anchorage (adhesion) of the adhesive to the face
stock. The influence of the carrier surface on the anchorage was discussed
above. The peel does not depend on the deformability of the adhesive only,
but also on the flexibility of the face stock [127,44]. The influence of the bulk
properties of the face stock on the peel, and the removability will be examined
in Chap. 6. The influence of the surface properties of the face stock on the
removability of PSAs will be discussed next.
An adequate anchorage of the PSAs on the carrier imparts good
removability from the adherent. Some special adhesives display built-in
release from untreated polypropylene. This means that the anchorage on
treated polypropylene is higher than on the untreated one, and thus no need
for release coating exists when coated on the corona-treated side of
polypropylene tape [157]. For tapes and protective films the back side of the
product can work with supplemental release coating or, with an appropriate
carrier material with low level of adhesivity (e.g., polyolefins) as an uncoated
liner [161]. Biaxially oriented multilayer PP films for adhesive coatings have
been manufactured to improve their adhesion to the adhesive coating by
mixing the PP with particular resins. The resins add up to 25% by weight.
The special importance of the anchorage as the main parameter for
removability is illustrated for protective films. As discussed in detail in [27]
the use of a primer improves the anchorage of the adhesive in such a manner
that for primed coatings a higher coating weight may be used, allowing the
same removability.
Rheology of Pressure-Sensitive Adhesives 79

Energy dissipation during debonding depends on the interfacial


adhesion [109]. Interfacial adhesion is a function of the surface quality. In
the case of removable adhesives high rate peel forces must be dampened by
high energy dissipation (i.e., good anchorage of the adhesive on the face
material is needed). Adhesion to paper and stainless steel is greater than the
anchorage to polyester film; therefore the bond fails due to a breakdown
within the adhesive layer when paper is used; when polyester film is used it
peels away leaving the adhesive layer virtually intact on the steel plate. On
the other hand, peel characteristics change on plasticized PVC [162]. A
thermodynamic approach considers adhesion between two materials as the
result of two different interfacial interactions, namely a dispersive (Id) one
and a polar (Ip) one [163]. The reversible adhesive work Wh is described as
the sum of these two types of physical interactions:

Wh ¼ Id þ Ip ð2:74Þ

In addition to these physical interactions there exist some chemical


interactions as well. Both physical and chemical interactions depend on
the quality, polarity, and chemical affinity of the face stock (or adherent)
material. In general a peel energy measurement yields a greater value than
the corresponding reversible work. On the other hand [163] peel energy
depends both on macroscopic factors of energy dissipation f(R) and on the
molecular factors of energy dissipation g (Mc):

PF ¼ Wh  gðMcÞ  f ðRÞ ð2:75Þ

where PF is the peel energy and Wh is the reversible work of adhesion.


Regarding the role of the macroscopic factors, the dependence of the
dissipation energy (peel) on the polymer thickness has been shown. Polymer
thickness implies coating weight, and as shown earlier, surface porosity
influences the coating weight. These results refer to hot-melt PSAs, but one
can admit their general validity for the PSA domain.
The influence of the anchorage on the peel is illustrated by the use of
polyester (PET) as standard face stock material for PSA tests. Paper differs
according to the manufacturing conditions and storage conditions. There-
fore a dimensionally-stable material (e.g., PET) is used as standard face
stock material in laboratory tests. Unfortunately, it is difficult to wet [164]
and anchorage onto PET remains poor. Thus it was observed in many cases
[165] that if the adhesive to be tested was applied on a polyester film, the film
peels away, leaving the adhesive layer initially intact on the steel plate;
however, if paper is used the bond fails due to breakdown within the
80 Chapter 2

Table 2.3 Dependence of the Peel on the Face Stock Material

180 Peel Adhesion on PMMA


at Room Temperature, N/25 mm

Coating Weight, g/m2 White Transparent

2.0 0.89 0.62


2.5 1.04 0.87
3.0 1.19 1.02
3.5 1.34 1.17
4.0 1.50 1.32
4.5 1.65 1.47
5.0 1.81 1.62
Face stock: 70-mm, blown, low density polyethylene film, white and transparent.
Adhesive: Crosslinked water-based acrylic PSA.

adhesive layer (adhesive transfer). Generally, peel tests using PET face stock
give lower values than the same tests using paper. The data from Table 2.3
illustrate the influence of the face stock material on the peel (removability).
For removable PSAs the adhesive must adhere better to the face stock
material than to the substrate. This adherence or anchorage may generally
be improved using primers [166–168,27]. The influence of the substrate
surface quality on the peel was studied indirectly as the separation energy
dependence on the surface tension of the substrate [16].
Converting Characteristics of the Pressure-Sensitive Laminate. It is
evident that the surface properties of the face stock material influence its
converting characteristics, as well as the converting characteristics of the
pressure-sensitive laminate. Convertability of the face stock material means
in a first approximation its coatability (printability). Convertability of the
pressure-sensitive laminate includes its printing ability, confectioning
characteristics (cuttability, die-cuttability, perforating, embossing, and
folding and winding) and its (end-use related) dispensing and labeling ability.
The distribution of forces during the cutting operation depends on the
slip of the laminate layers. The surface quality and smoothness of the face
stock surface influence the cuttability [see Eq. (2.60), parameter K3]. On the
other hand surface smoothness influences the anchorage of the adhesive
(i.e., its flow between the ‘‘walls’’ of the laminate). Cuttability increases with
the anchorage of the adhesive and decreases with the smoothness of the face
stock.

C ¼ f ð1=Sm
a
Þ ð2:76Þ
Rheology of Pressure-Sensitive Adhesives 81

where Sm denotes smoothness and a is an exponent taking into account the


dependence of the cuttability on external slip surfaces (between various
laminates) and internal flow surfaces (enclosed within a laminate).

Face Stock Geometry (Label Thickness) and Pressure-Sensitive


Properties
From labeling practice it is apparent that the label geometry influences the
adhesive properties of the label. Generally it is difficult to find suitable PSAs
for labels to be applied around sharp corners [169]. The label thickness
influences the peel properties, whereas peel adhesion depends on the face
thickness [170]. The influence of the thickness of the face stock material
is given by the change in flexibility. The flexural resistance depends on the
width b and thickness h of the sample (moment of inertia). Flexural
resistance influences the peel angle. In practice, however, the peel angle is
different from the angle theoretically supposed (90 or 180 ). The face stock
(label) form (width) also influences the peel force [171], and peel adhesion
increases linearly with increasing label width. It must be mentioned that the
global geometry of the pressure-sensitive laminate must be taken into
account for convertability tests (i.e., the whole thickness of the adhesive/
solid component sandwich). Similar to the formula given by Whitsitt [172]
for corrugated board construction, where the stiffness S is a function of the
elastic modulus of the liners, t is the liner thickness, and H is the combined
board caliper, one can write:

S ¼ ðE  tH 2 Þ=2 ð2:77Þ

and thus the global laminate thickness influences the stiffness related
properties (peel, tack, cuttability) of pressure-sensitive laminates.
Thickness changes of 4–5% for paper may give 15% changes of the
flexural resistance [173]. There exists an exponential dependence between
thickness and the flexural resistance (exponent ¼ 2.5). The flexural resistance
is not a material characteristic, but a result of the construction of the
laminate. The effect of the face stock thickness on the peel was studied [174];
peel adhesion increased up to a limit with the thickness of the face stock
(film). Especially at higher peeling rates the peel value increases with the
thickness of the face stock film used.

3.3 Influence of the Composite Structure


The structure of the PSA label influences the rheology of the bonded
adhesive. Generally the continuous/discontinuous character of the PSAs, the
symmetry of the build-up, and the multilayer structure of the PSA sandwich
82 Chapter 2

determine the adhesive and rheological properties of the bonded PSA. As


shown earlier the pull-off force of a label depends on the elastic modulus E,
width w, and thickness t of the label [Eq. (2.55)]. Experimental values of F
were plotted against N1/4 where N is the number of layers of tape (label)
applied on top of another and pulled away together. The effective label
thickness was proportional to N. In practice if the tape (or film label) face
material stretches so that the detachment angle becomes increasingly large,
then two or more film layers should be applied [68].

REFERENCES

1. W. Hoffmann, Kautschuk, Gummi, Kunststoffe, (9) 777 (1985).


2. S. W. Medina and F. W. Distefano, Adhesives Age, (2) 18 (1989).
3. A. Zosel, Colloid Polymer Sci., 263, 541 (1985).
4. Adhesives Age, (9) 37 (1988).
5. A. S. Lodge, Elastic Liquids, Academic Press, London and New York,
1964, p. 72.
6. D. Satas, Adhesives Age, (8) 28 (1988).
7. S. G. Chu, Viscoelastic Properties of PSA, in Handbook of Pressure Sensitive
Technology (D. Satas, Ed.), Van Nostrand-Rheinhold Co., New York, 1988.
8. J. Ferry, Viscoelastic Properties of Polymers, 2nd Edition, J. Wiley and Sons,
New York, 1970.
9. A. W. Bamborough, 16th Munich Adhesive and Finishing Seminar, 1991, p. 96.
10. Adhesives Age, (11) 40 (1988).
11. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 1.
12. J. P. Keally and R. E. Zenk, (Minnesota Mining and Manuf. Co., USA),
Canad. Pat., 1224.678/10.07.82 (US Pat. 399350).
13. A. S. Rivlin, Paint Technology, (9) 215 (1944).
14. M. Sheriff, R. W. Knibbs and P. G. Langley, J. Appl. Polymer Sci, 17,
3423 (1973).
15. P. J. Counsell and R. S. Whitehouse, in Development in Adhesives (W. C. Wake,
Ed.), Vol. 1, Applied Science Publishers, London, 1977, p. 99.
16. A. Zosel, Adhäsion, (3) 17 (1966).
17. J. Johnston, Adhesives Age, (12) 24 (1983).
18. R. Bates, J. Appl. Polymer Sci., 20, 2941 (1976).
19. W. Retting, Kolloid Zeitschrift, (210) 54 (1966).
20. A. Midgley, Adhesives Age, (9) 17 (1986).
21. K. McElrath, Coating, (7) 236 (1989).
22. J. Class and S. G. Chu, Org. Coat. Appl Polymer Sci., Proceed., 48, 126 (1989).
23. C. A. Dahlquist, Tack, in Adhesion Fundamentals and Practice, McLaurin and
Sons Ltd., London, 1966.
24. C. A. Dahlquist, in Handbook of Pressure Sensitive Adhesive Technology
(D. Satas, Ed.), Van Nostrand-Rheinhold Co., New York, 1988, p. 82.
Rheology of Pressure-Sensitive Adhesives 83

25. J. Kendall, F. Foley and S. G. Chu, Adhesives Age, (9) 26 (1986).


26. G. R. Hamed and C. H. Hsieh, J. Polymer Physics, 21, 1415 (1983).
27. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 5.
28. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 318.
29. P. Duncley, Adhäsion, (11) 19 (1989).
30. C. M. Chum, M. C. Ling and R. R. Vargas, (Avery Int. Co., USA), EP
12257927/18.08.87.
31. Kaut., Gummi, Kunststoffe, 40 (8) 34 (1987).
32. Coating, (6) 184 (1969).
33. S. E. Krampe and L. C. Moore, (Minnesota Mining and Manuf. Co., USA),
EP 0202831A2/26.11.86, p. 20.
34. T. Matsumoto, K. Nakmae and J. Chosoake, J. Adhesion Soc. of Japan,
(11)5(1975).
35. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 4.1.3.
36. C. Creton, C. Y. Hui and Y. Y. Lin, ‘‘Viscoelastic Contact Mechanics and
Adhesion of Periodically Roughened Surfaces,’’ Proc. 24th Annual Meeting of
Adh. Soc., Feb. 25–28, 2001, Williamsburg, VA., p. 21.
37. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 274.
38. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.
39. J. H. S. Chang (Merck & Co., Inc., Rahway, NJ), EP 0179628/24.10.1984.
40. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 2.
41. Die Herstellung von Haftklebstoffen, Tl.-l; 2-14d; Dec. 1979, BASF,
Ludwigshafen, Germany.
42. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 275.
43. A. Gent and P. Vondracek, J. Appl. Polymer Sci., 27, 4357 (1982).
44. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 5.
45. G. Fuller and G. J. Lake, Kautschuk, Gummi, Kunststoffe, (11) 1088 (1987).
46. J. A. Miller and E. Von Jakusch, (Minnesota Mining and Manuf. Co., St. Paul,
MN), EP 0306232B1/07.04.1993.
47. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 112.
48. M. Gerace, Adhesives Age, (8) 85 (1983).
49. G. Hombergsmeier, Papier und Kunststoffverarb, (11) 31 (1985).
50. G. Bonneau and M. Baumassy, ‘‘New Tackifying Dispersions for Water-Based
PSA for Labels,’’ 19th Munich Adhesives and Finishing Seminar, 1992, p. 82.
51. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 276.
52. G. Meinel, Papier und Kunststoffverarb, (10) 26 (1985).
84 Chapter 2

53. R. Köhler, Adhäsion, (3) 66 (1972).


54. H. Merkel, Adhäsion, (5) 23 (1982).
55. R. Köhler, Adhäsion, (3) 90 (1970).
56. M. Toyama and T. Ito, Pressure-Sensitive Adhesives, in Polymer Plastics
Technology and Engineering, Vol.2, Marcel Dekker, New York, 1974,
pp. 161–230.
57. L. Woo, ‘‘Study on the Adhesive Performance by Dynamic Mechanical
Techniques.’’ Paper presented at the National Meeting of the American
Chemical Society, Seattle, WA, March, 1983; cited in M. A. Krecenski, J.
F. Johnson and F. C. Temin, JMS Rev. Macromol. Chem., Phys., C26 (1) 143
(1986).
58. R. Mudge, ‘‘Ethylene-Vinylacetate Based, Waterbased PSA,’’ in TECH 12,
Advances in Pressure Sensitive Tape Technology, Technical Seminar
Proceedings, Itasca, IL, May, 1989.
59. T. H. Haddock, (Johnson & Johnson, USA), EP 0130080 Bl/02.01.85.
60. Kautschuk, Gummi, Kunststoffe, (8) 758 (1987).
61. Adhesives Age, (12) 35 (1987).
62. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 7.
63. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 8.
64. M. A. Johnson, Radiation Curing, (8) 10 (1980).
65. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 284.
66. Die Herstellung von Haftklebstoffen, T 1.2.2; 15d, Nov. 1979, BASF,
Ludwigshafen.
67. H. Muller, J. Turk and W. Druschke, BASF, Ludwigshafen, EP 0118726/
02.02.1987.
68. A. Gent and S. Kaang, J. Appl. Polymer Sci., 32, 4689 (1986).
69. Coating, (5) 11 (1971).
70. L. Jacob, ‘‘New Development of Tackifiers for SBS Copolymers,’’ in 19th
Munich Adhesive and Finishing Seminar, 1994, p. 107.
71. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.
72. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 115.
73. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 485.
74. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 3.
75. Fr. Patent 2331607; in F. F. Lau and S. F. Silver (Minnesota Mining and
Manuf. Co., St. Paul, MN), EP 0130087B1/02.01.1985.
76. Coating, (3) 65 (1974).
77. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 184.
Rheology of Pressure-Sensitive Adhesives 85

78. Jpn. Patent A 8231792; in P. Gleichenhagen, E. Behrend and P. Jauchen


(Beiersdorf A. G., Hamburg, Germany), EP 0149135B1/24.07.87.
79. Kautschuk, Gummi, Kunststoffe, (6) 556 (1977).
80. G. Menges and V. Lackner, Kunststoffe, 82 (2) 106 (1992).
81. H. Schuch, ‘‘Anwendung einer Mischungsregel für die dynamische Viskosität
von Polymerschmelzen,’’ German Rheological Society, Annual Meeting,
Erlangen,Germany, 27 April 1987; in Kautschuk, Gummi, Kunststoffe, 40 (9)
860 (1987).
82. C. Donker, R. Ruth and K. van Rijn, ‘‘Hercules MBG 208 Hydrocarbon
Resins: A New Resin for Hot Melt Pressure Sensitive (HMPSA) Tapes,’’
19th Munich Adhesives and Finishing Seminar, 1994, p. 64.
83. Adhesives Age, (10) 24 (1977).
84. Adhäsion, (9) 352 (1966).
85. E. H. Otto, 4th PTS Adhesive Seminar, Munich, Germany, PTS Reports,
Nr. 03, 1984, p. 39.
86. J. B. Bonnet, M. J. Wang, E. Papirer and A. Vidal, Kautschuk, Gummi,
Kunststoffe, (6) 510 (1986).
87. P. A. Mancinelli, J. A. Schlademann, S. C. Feinberg and S. O. Norris,
‘‘Advancement in Acrylic HMPSAs via Macromer-Monomer Technology,’’
PSTC XVII Technical Seminar, May 4, 1986, Woodsfield Schaumburg, IL;
in Coating, (1)12 (1986).
88. Coating, (6) 198 (1984).
89. D. Hadjistamov, Farbe und Lack, (1) 15 (1980).
90. Kautschuk, Gummi, Kunststoffe, (6) 545 (1987).
91. N. Hirai and H. Eyring, J. Polymer Sci., 37, 51 (1959).
92. A. K. Doolittle, J. Appl. Physics, 23, 236 (1958).
93. J. D. Ferry and E. R. Fitzgerald, Proc. 2nd International Conf. Rheology,
Butterworth, London, 1953, p. 140.
94. T. Timm, Kautschuk, Gummi, Kunststoffe, (1) 15 (1986).
95. H. A. Schneider and H. J. Cantow, Polymer Bulletin, (9) 361 (1983).
96. Kautschuk, Gummi, Kunststoffe, (9) 31 (1985).
97. A. N. Gent and A. J. Kinloch, J. Polymer Sci., A-2 (9) 659 (1971).
98. Kautschuk, Gummi, Kunststoffe, (8) 34 (1987).
99. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 24.
100. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 2.
101. K. W. Gerstenberg, Corona Treatment for Wetting and Adhesion on printed
Materials, European Tape and Label Conference, Brussels, p. 173, 1994.
102. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 487.
103. Adhesives Age, (3) 36 (1974).
104. P. Tkaczuk, Adhesives Age, (8) 23 (1988).
105. F. Fikentscher, Coating, (4) 86 (1972).
106. ‘‘Wiederablösbare druckempfindliche Kleber und Schutzfilme,’’ Duro-Tak,
Specialty Adhesives, Product Data, 8/89, p. 9, National Starch and Chemical
B. V., Adhesives Division, Zutphen, The Netherlands.
86 Chapter 2

107. A. Fau and A. Soldat, Silicone Addition Cure Emulsions for Paper Release
Coating, in TECH 12, Advances in Pressure Sensitive Tape Technology;
Technical Seminar Proc., Itasca, IL, 3–5 May, 1989, p. 7.
108. Coating, (9) 340 (1995).
109. A. C. Makati, Tappi J., (6) 147 (1988).
110. D. W. Mahoney and J. W. Hagan, PSTC Proceedings, Pressure Sensitive
Tape Council, Glenview, IL, 1982, p. 82.
111. Coating, (2) 39 (1970).
112. Adhäsion, (10) 273 (1975).
113. J. Hausmann, Adhäsion, (4) 21 (1985).
114. H. Hadert, Coating, (2) 35 (1985).
115. S. E. Orchard, J. Oil, Lab. Chem. Assoc., 44, 618 (1961).
116. C. P. Kurzendörfer, T. Altenschöpfer and H. J. Volker, Henkel Referate,
(20) 33 (1984).
117. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 361.
118. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 37.
119. H. Hardegger, Coating, (6) 193 (1992).
120. J. Voges, Klebstoffrohstoffe, Beschichtungstechnikum, BASF, Ludwigshafen,
Germany, 08.01.1993.
121. H. D. Patermann, Coating, (6) 224 (1988).
122. B. W. McMinn, W. S. Snow and D. T. Bowmann, Adhesives Age, (12) 34
(1995).
123. M. Giri, D. Bousfield and W. N. Unertl, ‘‘Adhesion Hysteresis in Viscoelastic
Contacts,’’ Proc.24th Annual Meeting of Adh. Soc., Feb. 25–28, 2001,
Williamsburg, VA., p. 15.
124. C. K. Hui, J. M. Baney and E. J. Kramer, Langmuir, (14) 6570 (1988).
125. R. A. Schapery, Intl. J. Fracture, (11) 141 (1975).
126. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, pp. 275, 294–296.
127. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 481.
128. R. D. Adams, R. Thomas, F. J. Guild and L. F. Vaughan, ‘‘Thick Adherend
Shear Tests,’’ Proc. 24th Annual Meeting of Adh. Soc., Feb. 25–28, 2001,
Williamsburg, VA., p. 59.
129. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 267.
130. J. Wilken, Ally. Papier Rundschou, (42) 1490 (1986).
131. A. N. Gent and G. R. Hamed, J. Appl. Polymer Sci., 21, 2817 (1977).
132. Kai Tak Wan and Yiu-Wing Mai, International Journal of Fracture, 74,
181, 1995.
133. E. O. Brien, T. C. Ward, S. Guo and D. Dillard, ‘‘Characterizing the Adhesion
of Pressure-Sensitive Tapes Using the Shaft Loaded Blister Test,’’ Proc. 24th
Annual Meeting of Adh. Soc., Feb. 25–28, 2001, Williamsburg, VA., p. 113.
134. Yeh-Hung Lai and D. A. Dillard, J. Adhesion, 56, 1996.
Rheology of Pressure-Sensitive Adhesives 87

135. Adhesives Age, (6) 32 (1986).


136. J. D. Hine, Verpackungsrundschau, Technisch wissenschaftliche Beilage, 9–15
(1964).
137. H. Grosmann, Verpackungsrundschau, (4) 1799 (1986).
138. Ally. Papier Rundschou, (42) 1152 (1988).
139. V. I. Pavlov, T. P. Muravskaya, R. A. Veselovskiy and M. T. Stadnikov,
Kompoz. Polym. Mater., 37,14 (1988).
140. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 346.
141. Adhäsion, (2) 43 (1977).
142. E. Brada, Papier und Kunststoffierarbeitung, (5) 170 (1986).
143. J. Skeist, Handbook of Adhesives, 2nd Edition, Van Nostrand-Rheinhold Co.,
New York, 1977.
144. J. Wilken, Papiertechnische Rundschau, (4) 4043 (1988).
145. E. Park, Paper Technology, (8) 14 (1988).
146. R. Lowman, Finat News, (3) 5 (1987).
147. R. A. Lombardi, Paper, Film and Foil Converter, (3) 76 (1988).
148. J. Pennace, Screen Printing, (7) 65 (1988).
149. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 324.
150. J. J. Bikermann, J. Colloid Sci., (2) 174 (1947).
151. Softal Electronic, Report, (107) 1 (1996).
152. B. Martens, Coating, (6) 187 (1992).
153. H. P. Seng, ‘‘Printing Properties of PP-Tapes,’’ European Tape and Label 146
Conference, 28.04.1993, Brussels, Belgium.
154. Softal Electronic, Report, (102) 1 (1996).
155. Deutsche Papierwirtschaft, (1) IV (1988).
156. U. Zorll, Chem. Ing. Tech., 60 (3) 162 (1988).
157. Solvay & Cie, Beschichtung von Kunststoffolien mit IXAN, WA, Prospect
Br.1002d-B-0, 3-0979.
158. J. Hildebrandt, J. M. Prausovitz and R. L. Scotty, Regular and Related
Solutions, Van Nostrand-Rheinhold Co., New York (1970).
159. C. M. Hansen, J. Paint TechnoL, (33) 104 (1967).
160. C. M. Hansen and K. Skaarup, J. Paint. Technol., (39) 511 (1967).
161. Vorwerk & Sohn GmbH & Co KG, Wuppertal, DE 4040917/02.07.1992.
162. Adhesives Age, (3) 36 (1986).
163. H. K. Müller and W. G. Knauss, Trans. Soc. Rheol, 15, 217(1971).
164. J. Lin, W. Wen and B. Sun, Adhäsion, (12) 21 (1985).
165. British Petrol, Hyvis, Prospect (1985).
166. E. Djagarowa, W. Rainow and W. L. Dimitrow, Plaste u. Kaut., (2) 100
(1970); in Adhäsion, (12) 363 (1970).
167. Adhäsion, (1/2) 27 (1987).
168. Coating, (4) 123 (1988).
169. Cham Tenero, 2nd International Cham-Tenero Meeting for the Pressure-
Sensitive Materials Industry, 1990, Locarno.
88 Chapter 2

170. A. W. Aubrey, G. N. Welding and T.Wong, J. Appl. Chem., (10) 2193 (1969).
171. J. R. Wilken, apr, (5) 122 (1986).
172. W. J. Whitsitt, Tappi J., (12) 163 (1988).
173. G. Renz, apr, (24–25) 960 (1986).
174. J. Sehgal, Fundamental and Practical Aspects of Adhesive Testing, Adhesives
85; Conference Papers, Sept. 10–12, 1985, Atlanta.
3
Physical Basis for the
Viscoelastic Behavior of
Pressure-Sensitive Adhesives

It was shown earlier that PSAs are required to bond rapidly to a variety
of substrates under conditions of low contact pressure and short contact
time. This characteristic feature of a PSA is called tack. During bond
formation contact on a molecular dimension between the adhesive and the
adherent is established, in isolated points of the geometrical contact area,
the number and size of which increase with the contact time through
deformation and flow, as well as by wetting.
Tack is a characteristic of amorphous high polymers above the glass
transition temperature (Tg) only. In terms of physical properties the Tg
represents the temperature range through which the polymer changes from
a hard, glassy state into a liquid, rubber-like state. In general a polymer with
low viscosity (low Tg) will be able to wet a substrate surface and establish
intimate contact with the adherent. An increase in Tg will lead to a stiffer
polymer, decreased wettability, and most likely a decline in adhesive
properties. The adhesive polymer must however possess sufficient cohesive
strength to resist further flow and subsequent deformation. The proper flow
of the adhesive ensures quick coverage of the substrate through wetting, but
a solid-like response to applied stresses. Efficient wetting implies a high
deformability (i.e., a low modulus), whereas a solid-like response to stress
means low deformability (i.e., a high modulus). Thus one can postulate:

Wetting out ¼ f ðE 1 Þ ð3:1Þ

End-use ¼ f ðEÞ ð3:2Þ

89
90 Chapter 3

It can be concluded that the viscoelastic flow properties of PSAs


materials are characterized by the value of the elastic modulus E and the Tg.
The interdependence between the macromolecular structure and adhesive
properties has been studied by means of rheology since the mid-1960s by
Hoechst [1]. The interdependence between Tg and auto-adhesion has been
investigated.
Investigations of the viscoelastic properties of many commercial labels
and tapes showed that Tg and E at the application temperature are the most
important features for a good PSA performance. The Tg may be measured
from the peak of the tan  curve (see Chap. 2). Several methods are used to
determine the glass transition of a polymer material; the most commonly
used ones are mechanical/rheological, calorimetric, dilatometric, and
dielectric methods. The calorimetric method has gained wide acceptance;
the most common instrument used is the differential scanning calorimeter
(DSC).

1 THE ROLE OF THE Tg IN CHARACTERIZING PSAs

Adhesive bonding is only possible if the fluidity of the materials involved


can ensure enough contact areas, in which chain segments would diffuse into
each other. In general, the Tg is one of the most important parameters in
order to determine the minimum usage temperature for polymeric materials;
it is in fact the temperature below which large scale molecular motion does
not occur. Below this temperature polymers become glassy and brittle, and
develop a high modulus; above this temperature the polymer behaves like
a more or less rubber-like elastic fluid. Its tensile strength and modulus
decrease above the Tg (i.e., the solid begins to liquefy).
Polymers for use in adhesives must act like a continuous, tacky, and
elastic layer. A continuous adhesive layer is formed by coalescence; the
real area of contact of an elastic film on a rough surface depends on two
parameters: the elastic modulus at the bonding frequency (G0 ) and the
relaxation properties of the polymer which govern the change of the real
area of contact; tack is the result of a bonding (viscous flow, wetting out,
elastic deformation) and debonding (deformation and failure) process,
where elasticity and viscoelasticity refer to energy storage and dissipation.
One of the contributing mechanisms to polymer adhesion is said to be
diffusion of chain elements across the polymer/substrate interface. The
concept makes sense when polymers in close contact are above their
respective Tg. During the stage of adhesive bonding the molecular mobility
of the PSA polymer contributes to the process of polymer relaxation,
which in turn accounts for the increase of the strength of adhesive joints
Physical Basis for the Viscoelastic Behavior of PSAs 91

with PSA-substrate contact times. All these requirements imply fluidity of


the adhesive at operating temperatures above the Tg.
On the other hand the internal strength of the adhesive (i.e., the
cohesion) must be high. The break energy of the polymers depends on their
molecular motion mechanism. This molecular motion is a function of the Tg
and of the molecular weight. Thus, the Tg and molecular weight values
are critical for the use of a polymer as an adhesive. This statement was
confirmed by Midgley [2] and Foley et al. [3]; they demonstrated that the Tg,
molecular weight, and molecular weight distribution are the basic properties
to characterize elastomer latexes used in adhesives. It will be shown later
that the Tg is a function of molecular weight as well as of other chemical and
macromolecular characteristics. The importance of the Tg as a characteristic
of the molecular motion will be described in a detailed manner in Section 1.2.
Also, it should be stressed that it is not the aim of the book to provide
a detailed discussion of the chemical or macromolecular basis of PSAs,
but rather an understanding of the adhesive process; there are some
fundamental correlations between the physical basis and characteristics of
PSAs which need to be discussed. First the value of the Tg, the factors
influencing the Tg, and the adjustment of the Tg will be reviewed.

1.1 Values of Tg for Adhesives


Generally the desired value of the Tg for coatings (film-like polymers)
depends on their application and end-use temperature, quality criteria, and
working mechanism. Paints and varnishes possess a Tg range of þ3 to
þ16 C — this is a relatively high Tg range — that yields hard polymer
surfaces [3]. Similarly dispersion-based offset printing inks show a minimum
film-forming temperature (MFT) of þ5 to þ10 C (i.e., also a relatively high
Tg range).
It should be mentioned that the required Tg of coatings also depends
on the face surface. Coatings on soft face materials need a lower Tg. As an
example, acrylic binders for absorbent nonwovens have a Tg below 0 C [4].
Adhesive coatings have to be softer than lacquer coatings or printing inks
(their Tg is lower also), but their Tg depends on the substrate in a similar
manner. Contact adhesives display a Tg range of 10 to 0 C. These
adhesives need fluidity only for film and bond forming. They must not
display fluidity in the bonded state. Pressure-sensitive adhesives have to be
liquid at room temperature and within the bond, thus they should have a Tg
of about 15 to 5 C or less. In fact the Tg of some PSA base polymers is
much lower (Tables 3.1 and 3.2).
The Tg of PSAs is generally the result of a formulation of a low-Tg
base elastomer and high-Tg additives (see Chap. 8). As an example, the
92 Chapter 3

Table 3.1 Glass Transition Temperature for Different Adhesive Types

Adhesive Type Tg ( C) References

Press (contact) adhesives 5 to þ15 5


Low energy curing adhesives 40 to þ30 5
Hot-melt adhesives þ5 to þ100 5
Wet adhesives 40 to þ15 5
Acrylic PSA 20 5
Acrylic PSA 40 6
Rubber/resin PSA 55 2
Vinyl acetate-acrylic PSA 35 to þ5 7
Electron beam-cured PSA 25 8, 9
Solvent-based PSA for protective films 20 10
PSA generally 15 to þ5 11
30–70 C below the 12
use temperature

current Tg range of rubber/resin PSAs (15 to 5 C) is a combination


of the low Tg of the rubber—less than 20 C (Table 3.2)—and the high
Tg of the tackifier resin—more than þ10 C (Table 3.3)—as shown in
Table 3.4.
In this case the loading level of both components is almost the same,
or of the same order of magnitude. Another example shows that the
relatively low amounts of high-Tg additives used in the synthesis of low-Tg
base elastomers may have a decisive influence on the Tg of the product.
Plasticizer-free vinyl acetate homopolymer (VAc) dispersions display a Tg of
þ30 C, and build a hard, brittle film at room temperature. This is due (at
least partially) to the protective colloid used. As is known, in this case a
polyvinyl alcohol (PVA) with a Tg of þ70 C is used as a protective colloid.
The high value of the Tg of PVA has a negative influence in other cases of
formulation of adhesive polymers too; PVA used as a thickener lowers the
tack of PSA formulations. The values of the Tg for common PSAs have
changed as a function of the raw material development. Some years ago,
a typical PSA had a Tg of about 40 C, now common acrylic PSAs
possess a Tg range of 40 to 60 C (see Table 3.1); the best performing
styrene-butadiene-rubber latexes used for PSAs are those with a Tg of 35
to 60 C [11].

1.2 Factors Influencing Tg


The Tg is a function of chain mobility and flexibility. Chain mobility
depends on internal interactions, the internal forces between polymer chains
Physical Basis for the Viscoelastic Behavior of PSAs 93

Table 3.2 Glass Transition Temperature Values for Base Materials for PSA Labels

Polymer Tg ( C) References Applications

Polybutadiene  85 13 Rubber/resin PSA


Polyethylhexylacrylate  70 11, 13, 14 Acrylic PSA
 62 15
Polyisobutene  60 16 PSA, sealants
Polyisoprene, low  72 17 PSA
molecular weight  65
Natural rubber  64 18 PSA
 56
Poly(styrene-butadiene)  59 19, 20 —
carboxylated  54
Poly(styrene-butadiene)  55 21, 22 —
 50
Poly(ethylene-butylene) 53 23 UV light curable PSA
Poly(ethylhexylacrylate- 49 24 PSA
acrylic acid)
SBR latex  17 (tan ) 46% ST 25 WB PSA
 18 44% ST 25
 46 25% ST 3
Poly-n-triethylacrylate  54 13 Acrylic PSA
 46 15 —
Poly(ethylene-vinyl  40 26 PSA
acetate-acrylate)
Polyvinylchloride,  40 18 Face stock
plasticized 0
Poly(ethylene-vinyl  35 25 PSA
acetate-dioctyl  25
maleinate)
Poly (SIS)  28 25 EB curable HM on
Kraton D-1320X
basis
Poly(propene-butene) 16 27 Semi-pressure-
sensitive adhesive
Poly(propene-octene)  25 28, 29 PSA
Polyethylacrylate  22 13 Acrylic PSA
 14 15
Polypropylene  11 30 Face stock
amorphous  14 27 Semi-pressure-
sensitive
Poly(ethylene-  10 18 PSA, face stock
vinylacetate) þ 15
(continues)
94 Chapter 3

Table 3.2 Continued

Polymer Tg ( C) References Applications

Polymethylacrylate  10 16 Acrylic PSA


þ16 15
Polyvinylacetate þ29 13 —
Poly(hydroxyethyl þ55 31 Radiation curable PSA
methacrylate)
Polyvinylchloride þ75–105 16 Face stock
þ85 32
Polyacrylnitryl þ 100 16 Acrylic PSA
Polystyrene þ 100, þ110 13 Hot-melt, face stock
Polymethylmethacrylate þ 105 16, 33 Acrylic PSA
Polyacrylic acid þ 106 13 Acrylic PSA
Polymethacrylic acid þ 185 13 Acrylic PSA, thickener

Table 3.3 Glass Transition Temperature and Ring and


Ball Softening Temperature for some Tackifier Resins

R&B softening
Resin type Tg ( C) temperature ( C)

Regalrez 1065 17 65
Regalrez 1078 23 76
Regalrez 1094 33 94
Regalrez 1033 7 33
Escorez 9241 5 —

Table 3.4 Changes in the Tg Due to Tackification—


Aqueous Dispersions

Components

PSA Tackifier resin dispersion


Concentration, parts (w/w) Tg ( C)

100 0 54.7
85 15 53.7
70 30 38.7
55 45 26.4
60 40 17.1
60 40 22.8
50 50 27.8
Physical Basis for the Viscoelastic Behavior of PSAs 95

and segments. Intermolecular forces are a function of the nature and


dimension of the polymer chains (i.e., of chemical composition, structure,
and molecular weight). Equations can be derived that express the
dependence of the second order transition temperature on molecular
weight, plasticization, degree of crosslinking and copolymerization. Such
equations may be reduced in form to equations derivable from the free
volume theory. From the finding that the maximum in debonding force
corresponds to the minimum value of the Cp  Tg product for PSA
polymers, the conclusion was derived that this product (defined as a measure
of heat that has to be expended to provide the polymer transition from
the glassy to the viscoelastic state, and to impart translational mobility to
the polymeric segments, constituted from contributions of free volume and
the energy of cohesion) can be used as a predictive criterion of adhesive
behavior for various polymers [34–36]. This argues in favor of the interplay
of free volume against cohesive energy in polymers as a general determinant
of pressure-sensitive adhesion. The Tg depends mainly on the chemical
composition and structure of the adhesive and the molecular weight of the
base polymer. The chemical composition and structure of the polymers may
be influenced by synthesis and formulation. As a general principle one can
say that reduction of the intermolecular forces improves chain mobility and
therefore lowers the Tg.

Dependence of Tg on Molecular Weight


Concerning the composition dependence of the Tg of compatible polymer
blends, the parameters of an extended Gordon-Taylor equation are not only
polymer specific, but also molecular weight dependent [37]. Chain mobility
decreases with increasing molecular weight. The Fox-Flory equation relates
Tg to the number average molecular weight Mn [38]:

Tg ¼ Tg  C  Mn1 ð3:3Þ

where C denotes a constant. This dependence is not a linear function; the Tg


dependence on molecular weight is more pronounced in the low molecular
range. On the other hand PSA formulations are mainly mixtures of high
molecular base elastomers and lower molecular additives. In some cases
the base elastomer itself behaves like a mixture of two macromolecular
compounds with different MW, due to a special sequence distribution.
Therefore, it is more appropriate to discuss the dependence of the Tg on
the molecular weight for the base elastomer and the lower molecular
formulating additives separately.
96 Chapter 3

Molecular Weight of the Base Elastomers. Because of the significant


influence of the molecular weight on the Tg a relatively low degree of
polymerization (DP) is recommended for polymers used as adhesives (e.g.,
for polyvinyl acetate, the degree of polymerization amounts to several
thousands); for PSAs the molecular weight should be lower. The polymer
molecular weight has a significant effect on the balance of pressure-
sensitive adhesive properties [3]. As discussed in [39] shear is roughly
proportional to molecular weight up to relatively high molecular weights
at which the shear resistance drops off dramatically in some polymers.
Tack is typically high at very low molecular weights and decreases
gradually as the molecular weight is increased. Peel adhesion typically
exhibits a discontinuous behavior, increasing with molecular weight up to
moderate molecular weight levels and then gradually decreasing as the
molecular weight further increases. For polyisobutene a maximum of the
peel resistance has been found in a narrow range of molecular weights [40];
the molecular weight for PSA base polymers is situated in the range of
1000–2000. As shown earlier [41] typically the molecular weight of solvent-
based PSA polymers is limited to about 100,000, with a solids content
ranging between 12 and 25%. At this molecular weight level the cohesive
strength remains marginal and crosslinking appears necessary for better
performance. High molecular weight polymers offer high shear and good
thermal stability. An acrylic acid copolymer with a molecular weight of
70,000–80,000 is proposed for better wet adhesion properties [42].
Polyisoprene used for PSAs has a molecular weight of 4  103–5  106.
The Tg of low molecular weight polyisoprenes differs from that of natural
rubber (Table 3.2). Whereas the molecular weight of solvent-based acrylics
is about 100,000, water-based ones may have molecular weight values as
high as 1,000,000 [5].
It is relatively simple to obtain low or standard molecular weight
values for linear, uncrosslinked polymers. It is more difficult to limit the
value of molecular weight for partially crosslinked natural or synthetic
rubbers used as PSAs. In this case the molecular weight is described by
the assembly of linear and crosslinked polymer chains. As an example, for
PSAs based on carboxylated styrene-butadiene (CSBR) latexes, the weight
percentage of the polymer having a molecular weight of about 300,000 or
more is employed as a relative measure of the effective molecular weight [5].
In such styrene-butadiene copolymers higher gel content is an indication for
higher molecular weight. Above 30% gel, tack and peel resistance decrease
rapidly. The degree of crosslinking, i.e., the length of the mobile chain
sequences may influence macroscopic viscoelasticity. The mechanical
properties of crosslinked polymer networks are determined by crosslinking
density, the molecular weight between junctions (Mc), and the Tg. The
Physical Basis for the Viscoelastic Behavior of PSAs 97

elastic properties of the polymeric network depend on the length of the


polymeric ‘‘bridges.’’
At sufficiently low frequency of deformation, linear chains begin to
disentangle by the so-called ‘‘reptation process’’ and the elastomer flows.
Thus, entanglement molecular weight Me affects the mechanical and
rheological properties of the polymer. The time constant for reptation for
linear molecules is proportional to MW to the third power at constant
temperature. The acrylics are inherently (elastic and) tacky due to their high
entanglement molecular weight (Me), without the need of adding tackifier
[43,44]. A physical model correlating polymer viscoelastic behavior to PSA
performance has been developed [45] to explain the relationship between Tg,
Me, and PSA performance
In a similar manner the minimum length of the soft sequences in
thermoplastic elastomers (and polymer blends) influences the properties
of the elastomer [46]. In such cases, both the global molecular weight
and the molecular weight of the elastomer (or plastomer) chain
sequences influence the glass transition temperature. As will be discussed
later, the dimensions of the polymer sequences affect compatibility also,
and thus delimit the regulating possibilities of the Tg. In conclusion, the
molecular weight, the molecular weight between entanglements (Me), the
molecular weight of the crosslinking bridges (Mc), and the molecular
weight of the polymer sequences (Ms) affect the Tg (and the modulus) of
the polymers.

Molecular Weight of Tackifier Resins. The tackifier must have a


lower molecular weight than the base polymer, but a higher Tg than the
base polymer, and it must be compatible. Tackifier addition can cause an
increase in entanglement molecular weight. For instance for an acrylic
polymer Me increased by tackifying from 30 to 93 kg/mol. For this polymer
the weight average molecular weight (Mw) of the linear polymer was 68 k,
and the average molecular weight between crosslinks (Mc) was 52 k [43].
Tackifying resins used in common adhesive formulations possess a
molecular weight ranging from 300 to 3000 where the Tg strongly depends
on the molecular weight [47–49] (see Chap. 8, Section 2.3 too). The resins’
molecular weight controls the location and sharpness of the loss/storage
peak (tan ) in tackified styrene block copolymers (SBCs). It is admitted
that by tackification (for a given softening point and concentration of
the resin) the higher the increase of tan  the better the compatibility.
According to [50] acid resins exhibit a narrow molecular weight
distribution, ester resins have a broader MWD and somewhat lower Tg
(4–6 C).
98 Chapter 3

Dependence of Tg on Chemical Composition and Structure


The bonding/debonding characteristics of PSAs depend on their fluidity,
that is, on the internal mobility of the polymer chains. Internal mobility is a
function of the internal forces acting on a sterical (volume and structure) or
polarity basis. Thus, in a first instance, polymer mobility may be regulated
by changing the polymer (chain) composition through the choice of sterical/
polarity characteristics of the monomers. These changes are reflected by the
value of the Tg. First the dependence of the Tg on the chemical composition
will be reviewed.
Flexible Main Chain. The proper choice of the base monomers and/
or those acting via polar forces permits adjustment of the Tg. A detailed
discussion of the interdependence of the base monomers will be given in
Chap. 5. The influence of the base monomers on the Tg will be discussed
first, going from general problems (polymers) to special ones (PSA).
Theoretically, the unpolar, low-volume chain of polyethylene (PE) should
display a high mobility, and thus a low Tg. In fact PE has a medium to low
Tg(30 C) but a very high tendency to crystallize. Therefore, it does not
fulfill the criteria for tack, namely to be amorphous as referred to earlier. In
order to avoid crystallinity and ordering of polymer chain segments on an
ethylene basis, sterical hinder appears necessary, namely, that voluminous
(branched) monomer units have to be inserted between ethylene units or
inversely ethylene units should be inserted as ‘‘plasticizer’’ units between
other vinyl, acryl, or other monomer units.
This is illustrated (for common plastics) by the plasticizing effect of
ethylene in copolymers with propylene [30]. In this case the Tg decreases
from 262 K to 214–222 K. A similar, plasticizing effect of ethylene may be
observed for common, non-pressure-sensitive adhesives [51]. The same Tg
decreasing effect for polyvinylacetate is produced by the addition of 32%
butyl acrylate; adding 18% ethylene decreases the Tg to 0 C [52]. The Tg of
polyvinylacetate (PVAc) dispersions usually amounts to þ30 C. Polydibutyl-
maleinate dispersions have a Tg of 5 to þ10 C, and poly(vinylacetate-
ethylene) dispersions have a Tg range of 10 to þ5 C [53]. In the synthesis
of PSAs cases exist where an adequate lowering of the Tg may be obtained
by the use of ethylene only. Vinyl acetate-dioctyl maleate copolymers with a
dioctyl maleate content as high as 50% do not possess enough tack and peel
[54]. The use of a monomer ratio of 15–30 wt% of ethylene with 35–50%
2-ethyl hexyl maleate, dioctyl maleate, or fumarate, introduces sufficient
tack and peel. It should be mentioned that other low-Tg materials than
ethylene-based ones show this crystallizing tendency as well. A striking
example is given by the elastomeric sequences of rubbery block copolymers
(hot-melt PSAs), where PE-like sequences are interrupted by butylene units.
Physical Basis for the Viscoelastic Behavior of PSAs 99

Side Groups. Flexible main polymer chains with a low volume and
unpolar side groups, yield a low Tg. Generally, in order to avoid
intermolecular, segmental contact of the chains and their build-up into
structures, sterical hindrance has to be incorporated. According to the
length, bulk volume, and polarity of the side groups an increase or decrease
of the Tg may be obtained. This effect is used in the synthesis of special
plastomers too. Due to the presence of long chain branching the so-called
metallocene resins (very low density polyethylenes) display a combination of
elastomeric and thermoplastic properties. Such polymers give a tacky film
[55]. The effect of the side groups on the Tg can be illustrated in the field of
the acrylics. Definite trends in physical properties of the polymers can be
observed, when higher molecular weight acrylic esters are used; only a few
acrylic monomers are adequate for PSAs. The most important of these are
ethyl hexyl acrylate and butyl acrylate. As the ratio of low molecular weight
(short-branched) monomers is increased, the following changes in properties
generally occur:
Tackiness increases as Tg is lowered;
Hardness decreases;
Tensile strength decreases;
Elongation at break increases.
According to the above scheme (at least theoretically), it is possible to
classify the base monomers into softer (plasticizer) or stiffer (hardener) ones.
The harder ones are short side chain (bulky or polar) acrylates. As an
example illustrating the effect of the side chain length on the Tg the ethyl
acrylate-methyl methacrylate copolymers should be mentioned: 75% ethyl
acrylate forms polymers with a Tg of 5 C, but the decrease of the soft
monomer concentration to 25% leads to a Tg of þ 80 C. The principal
monomers used in emulsion polymerization can be divided into two main
groups: those producing hard polymers such as vinyl acetate, styrene, and
vinyl chloride, and those capable of softening hard polymers or producing
soft polymers such as ethylene or butyl acrylate [54]. Generally soft polymer
compositions must be hardened by polar monomers (see Chap. 5) which
may interact physically or chemically, such as acrylic acid, glycidyl
methacrylate, N-vinyl-pyrrolidone, methacrylamide, acrylonitrile, and
methacrylic acid [56]. Because of the very low Tg of ethyl hexyl acrylate,
which is a principal monomer for PSAs, it is recommended to copolymerize
with acrylonitrile (e.g., Acronal 80D, 81D), styrene, vinyl acetate, or methyl
methacrylate in order to increase the cohesion (i.e., the Tg).
The level of the stiffening (hard) monomers should be kept as low as
possible. As an example, the usual styrene content of CSBR latexes varies
between about 40 and 70%; with the Tg the values range from about 30 C
100 Chapter 3

to room temperature [11]. Generally, the concentration of the most


commonly used carboxylic acids lies between 3 and 6%. It should be
mentioned that the inclusion of voluminous side groups may be carried out
by homopolymerization also (e.g., in the polymerization of butadiene the
concentration of 1,2-butadiene units influences the Tg). On increasing the
1,2-butadiene content, the Tg increases also because of the hindrance of
the internal rotation and lowering of the chain flexibility. The influence of
the length of the side groups of the monomer on the Tg was studied in a
detailed manner by Zosel [57] for common acrylic esters used as raw
materials for PSAs. A strong dependence between side chain length and the
Tg of the PSAs was demonstrated, namely, the Tg decreases with increasing
side chain length. Olefin-based monomers with neo structure (e.g., vinyl
neodecanoate and vinyl neododecanoate) H2C¼CH-O-CO-C(R1)(R3)-R2 in
which R1þR2þR3 average 8–10 carbons, possess a particularly branched
structure and (compared to acrylic polymers) the homopolymers of these
monomers exhibit unusually high entanglement molecular weight (Me) at
comparable Tg [58].
Polar Monomers. Polar monomers are copolymerized with the base
elastomer in order to improve the mechanical properties (cohesion) of the
adhesives. Their inclusion increases the Tg. This may be explained by
increasing dipole interaction, or enhanced hydrogen bonding. Next the
influence of the small scale parameters of the polymer structure, mainly
the microstructure, the morphology, and the linear/crosslinked character
of the polymer will be reviewed.
Influence of the Microstructure. Many variations can be made in the
build-up of a thermoplastic rubber e.g., molecular weight, styrene content,
the number of blocks in the chain (i.e., two, three, or multiblocks), and
furthermore in their configuration (e.g., a linear polymer, or one that is
branched or radial). Varying the structural parameters has a direct influence
on their physical properties, hardness, and viscosity. As is known the
electron beam-crosslinkable SIS rubber has a branched molecular structure.
Ethylene copolymers are directly synthesized in high pressure equipment,
and the comonomer is randomly distributed along the polymer chain [59].
For copolymers containing large amounts of softening comonomer, the
PE segments are short and the melting point and amount of crystallinity
are depressed, reducing the heat resistance.
Graft or block copolymers are ideal as adhesives because their two-
phase structure provides a broad operating range. They can be designed as
structures containing microphases and separated into a tough, rubbery,
low-Tg phase and a hard reinforcing phase (either a glass or a crystalline
material). In some cases (e.g., release coatings) a reduced segmental mobility
Physical Basis for the Viscoelastic Behavior of PSAs 101

is required. It has been shown that the methylsilicate functioning as a high


release agent (HRA) to modify the viscoelastic properties of polydimethyl-
siloxanes (PDMS) effectively reduces the segmental mobility of the flexible
PDMS backbone [60].
Sequence Distribution and Length. Within the block copolymers used
for hot-melt PSAs there are amorphous polymer segments with Tg higher
than room temperature (RT), and other, amorphous polymer segments with
Tg
RT [61]. The Tg of the thermoplastic component (polystyrene)
influences the stiffness of the whole polymer. Its value may be changed,
thereby modifying the length of the sequence. The Tg of polystyrene is a
function of its molecular weight up to a value of about 20,000, after which
the Tg remains constant at 100 C [13]. As an example, the molecular weight
of the polystyrene block in SIS used for hot-melt PSAs is about 104, and its
Tg is about 85 C [62]. Generally, the sequence length of the elastomer varies
between 500–700 monomer units (with Tg at 70 C), where the styrene
blocks have a length of 200–500 units [63]. According to these macro-
molecular characteristics, in SIS copolymers the polyisoprene (PI) phase
shows a peak for the loss modulus in the temperature range of 40 C [33].
Another peak around 100 C is linked to the softening point of the pressure-
sensitive domains. In a similar manner a separate Tg was observed for the
hard phase (105 C) and for the soft phase (45 C) in acrylic hot-melt PSAs
[64]. In each case the high-Tg endblock will cause a high cohesion/stiffness
at room temperature. The thermal stability improves by tackifying these
elastomers with resins that have a melting point higher than the Tg of the
polystyrene domain. As stated by [65] investigating the effect of the di/tri
ratio and of the temperature, for different materials (81% SIS þ 19% SI)
and (5.7% SIS þ 54.3% SI), a broadened glass transition is observed and
a second relaxation process appears at low frequencies (high temperatures)
while the SI content is increased.
Sequence distribution is important because the additivity of the Tg is
valid for random copolymers only. Kendall, Foley, and Chu [3] showed that
the Tg of a random styrene-butadiene copolymer is approximately related to
the styrene content, through the following equation:
Tg1 ¼ w1 Tg1
1 1
þ w2 Tg2 ð3:4Þ

where w1 and w2 are the weight fractions of each copolymer. The ‘‘inverse’’
additivity of the Tg is very important for adjusting the Tg (see Section 1.3).
A general applicability of the proposed PSA model (based on free volume
considerations [34–36]) is confirmed by the fact that the Tg of such systems
obeys fairly well the Kovacs [66] equation concerning the additivity of glass
transition temperatures, and relating the Tg of the system to the Tg of the
102 Chapter 3

components, the fractional free volume at the glass transition temperature,


and the thermal expansion coefficient at Tg (See Chap. 5).

Influence of Morphology on Tg. The nature and build-up of the


polymer chain influences the polymer morphology. Intramolecular forces
may promote crystallization. Crystallization reduces chain mobility, and
thus increases the Tg. By crystallization or crosslinking through the main
valencies there appears a hindrance of the chain mobility [67]. Crystalline
or crosslinked high polymers are not tacky, even above their Tg. This is
an important characteristic not only for adhesives, but also for face stock
materials, or substrates to be bonded. Different adhesion levels can be
achieved on high-density polyethylene (HDPE), which has high crystallinity,
and low-density polyethylene (LDPE), which has low crystallinity. Because
of its crystallinity, HDPE only displays one third of the contact area of
LDPE. As described in [68], very low-density polyethylene films manufac-
tured by casting (i.e., with a greater amorphous content) and having a low
film thickness (conformability) display self-adhesivity at room temperature.
Thus, different adhesion on polyethylene types with different crystallinity
influences the anchorage of PSAs applied on face stock materials or
substrates made from these polymers. On the other hand, crystallization
phenomena may also influence the adhesive flexibility. As an example one
can refer to the design of raw materials for hot-melt PSAs. Here the
elastomer segment composition in a rubbery block copolymer must be
designed to adequately suppress crystallinity by interrupting the poly-
ethylene-like chains (with a butylene monomer unit) without at the same
time enriching the polymer in butylene, in such a way that the Tg is
substantially raised. The purpose of this restriction is to obtain a saturated
olefin rubber block, with the lowest possible Tg and the best rubbery
characteristics. Another possibility is to insert isoprene units. Each isoprene
monomer addition to the growing chain incorporates two methylene and
one propylene sequence (or one ethylene and one propylene for ethylene
propylene rubber, EPR, or ethylene propylene diene rubber, EPDM,
respectively). Although isoprene can also polymerize through the 3,4
carbons, the resulting hydrogenated structure which has an isopropenyl side
group, does not provide any crystallinity or substantial increase of the Tg. In
a similar manner, the increase of the acrylic acid content (decrease of the
crystallization tendency) improves the peel values, through the lowering
of the Tg [63]. Common amorphous polypropylene shows a crystallization
tendency, and may be used for semi-pressure-sensitive adhesives only. They
lose their performances because of crystallization. The copolymerization of
propylene with hexene or octene allows the synthesis of pressure-sensitive
raw materials (see Chap. 5). Leibler et al. [69] studied side chain fluorinated
Physical Basis for the Viscoelastic Behavior of PSAs 103

acrylic copolymers. This family of polymers organizes at low temperature


into very structured lamellar phases, with in addition, the side chain groups
crystallized on a hexagonal array inside the layers. They organize into
isotropic phases at around 35 C, depending on their composition. When the
temperature is raised and the transition between the crystallized lamellar
and the isotropic phase occurs, the tack energy goes from zero to a typical
value of 50 J/m2, on a temperature range of 2–5 C. The tack energy G then
decreases again as temperature is further increased, like in a conventional
adhesive.
The influence of the sequence build-up (and morphology) on the
rheological characteristics is illustrated by linear and radial block
copolymers. Because of their more ordered structure, the radial copolymers
(SBCs also) generally show a longer linear portion of the rubbery plateau
and lower values of tan . The two tan  peaks characteristics of block
copolymers are more extreme (87 and þ85 C) for a radial styrene block
copolymer than for the linear one (82 and þ75 C). The butadiene used in
these types of block copolymers normally has a loss tangent peak at 90 C,
the styrene one at þ100 C.

Influence of Crosslinking. Steric hindrance by crystallization or


crosslinking reduces chain mobility and thus increases the Tg. Relaxation
in the glass transition region involves cooperative motion of molecular
segments. An approach pioneered by Adam and Gibbs involves the concept
of cooperativity where the configurational entropy at a given temperature
is lower than that expected for independent relaxation of the primitive
molecular segments. Cooperativity is a theoretical concept that describes the
scale over which intramolecular forces are being propagated in a relaxation
event and the strength of coordinated motions among polymer chains
needed for the relaxation to occur.
Fracture toughness characterizes the glassy state. The role of the
network density on glass transition for vinyl ester polymers was studied
by [70]. For vinyl ester polymers the narrow range in molecular weights
between crosslinks, did not make apparent any correlation between the
fracture toughness and molecular weight between crosslinks (Mc). No
correlation was found between fracture toughness and the domain size at Tg.
In the case of crosslinking the effect on Tg depends on the crosslink
density. The crosslink density is a function of the crosslinking method too.
For instance the modulus of polybutadiene networks crosslinked chemically
or by radiation differs [71]. Common electron beam-curing polymers are
not recommended for PSAs because of the high crosslinking density, which
imparts hardness and a tack-free character. For PSAs it is recommended
to have a Tg lower than 20 C and a low crosslink density. Crosslinking
104 Chapter 3

agents in an adhesive influence the Tg and thus the minimum film-forming


temperature. Woodworking glues having trivalent chrome nitrate as the
crosslinking agent (for better water resistance) display a MFT increase of
3–4 C.
Generally, the high temperature performance of hot-melt adhesives is
limited by the Tg of the endblocks of the base polymers. Better temperature
resistance may be achieved by crosslinking the polymers [72].

1.3 Adjustment of Tg
For practical use adhesives must be fluids (i.e., they must be applied at a
temperature above the Tg). An end-use temperature higher than the Tg can
be achieved by melting the adhesive (e.g., hot-melt PSAs) or by changing
(at least temporarily) its Tg. Fluidity may be achieved by other ways, such
as using a volatile plasticizer like a solvent (solvent-based adhesives), or
a dispersing agent (water-based dispersions). These methods permit the
application of the adhesive, but a continuous adhesive layer is formed above
the Tg only. Thus adjusting the Tg appears to be required.
The composition of the adhesive is determined by polymer synthesis
or most frequently through formulation. Polymer synthesis regulates
intermolecular forces by the nature and polarity of the chain building
monomer units. Polymer formulation regulates intermolecular interactions
by changing the internal structure of the adhesive polymer.
Polymer Synthesis
Earlier a short description of the parameters influencing the Tg was given.
The macromolecular and chemical characteristics of polymers (molecular
weight, chemical composition, structure) can be adjusted through the
polymer synthesis, by the choice of the raw materials and of the
polymerization technology. In the case of solvent-based adhesives and
hot-melt PSAs, the final composition of the adhesive layer is that of the
bulk material (i.e., the polymer does not contain important amounts of
other nonpolymeric additives).
In the case of water-based dispersions, the final adhesive composition
may differ from that of the bulk adhesive, by the additives used, and/or the
water included in the adhesive composition. Thus there are differences
between the Tg of the synthesized, uncoated (bulk) polymer and the Tg of
the applied adhesive. Water-based adhesives are almost always formulated
offering the advantage of Tg adjustment by both procedures (changing the
composition of the bulk or of the dispersed adhesive). As a function of the
level of formulating agents, the Tg differs more or less from that of the base
polymer. On the other hand, it was shown that for natural rubber latices,
Physical Basis for the Viscoelastic Behavior of PSAs 105

the Tg of natural rubber latex is effectively indistinguishable from that of


the dry rubber separated from the same latex [73].

Adjusting the Tg by Formulating


In polymer synthesis different copolymerization techniques of various
monomers giving different chain flexibility and intermolecular forces (i.e., a
different macromolecular order) lead to adequate, low-Tg PSAs. Another
possibility to change the macromolecular order is the change of the
intermolecular forces by the change of the ‘‘packing density’’ through the
use of macromolecular or micromolecular diluting agents such as tackifiers
or plasticizers.
Use of Plasticizers. Plasticizers change the degree of order of the
molecules, also changing their entropy and enthalpy. The use of plasticizers
for current thermoplastic materials (e.g., polyvinyl chloride, or PVC) is well
known. Using high amounts of plasticizer, it is possible to drastically reduce
the Tg as an example the addition of 40% plasticizer to PVC decreases its
Tg to 40 C. Such a plasticized PVC used as the face material for pressure-
sensitive labels, possesses a Tg in the same temperature range as the PSAs
applied on it, which contributes to very good anchorage of the PSAs on
the face stock material. Low stiffness and thus poor cuttability are the
drawbacks associated with low Tg. Skin softeners used in transdermal
applications can also plasticize the medical adhesive thus lowering the Tg
and reducing the peel strength. The addition of DMSO reduced the Tg by
19 C [74]. The decrease of the Tg by a micromolecular plasticizer yields
increased plasticity and thus better tack of the adhesive, thus plasticizers
act like tackifiers.

Use of Macromolecular Tackifiers. In a manner different from


plasticizers, the use of macromolecular tackifiers results in a higher Tg. In
this case the better flow properties and improvement of the tack are due to
the decrease of the modulus.
Earlier a special case to forecast the Tg for random styrene-butadiene
copolymers was mentioned. Additivity of the Tg as described for this
polymer is a general phenomenon. Generally the Tg is additive, that is, the
Tg of polymer compounds is a function of the nature of the components
and their ratio. Practically, this is only true for compatible polymers, and
thus polymer compatibility also influences the adjustment of the Tg.
The dynamic mechanical properties of polymers are basically
determined by their mutual solubility. If both are compatible and soluble
in one another, the properties of the mixture are approximately those
resulting from the random copolymer of the same composition. Many
106 Chapter 3

polymer composites, however, form two phases due to their natural


insolubility. In that case the tan  temperature graph will show two peaks,
which identify the Tg of each of the components. As an example,
a good compatibility between natural rubber and glycerine ester of a
hydrogenated colophonium was demonstrated by Zosel [57]; the Tg of the
mixture increases for a higher resin content, but no second Tg is observed in
the composition range of 80–200% rubber. In a similar manner the increase
of the Tg and the absence of distinct melting points was observed for nitryl
rubber-phenolic resin and EPDM-hydrocarbon resin [75]. For compatible
adhesive components (most rubber-resin systems) an acceptable estimate of
the tan  peak temperatures can be made using the Fox equation [37,76,77].
As Schneider and Leikauf [78] have shown the Fox equation [79] results in a
first approximation from the Gordon-Taylor expression, which was deduced
in supposition of additivity of the specific volumes. As Kovacs [80]
demonstrated, the supposition of the specific volume additivity requires that
the respective free volumes be additive. For miscible polymer blends the
Gordon-Taylor [81] equation may also be used [82]:

Tg ¼ ðw1 Tg1 þ w2 Tg2 Þðw1 þ kw2 Þ1 ð3:5Þ

where Tg, Tg1, and Tg2 refer to the peak temperature of tan  for the
blend and the pure components, w1 and w2 are the weight fractions of the
components, and k is a polymer constant. For compatible polymers
deviations from the additivity rule is less than 5 C [83]. It should be
mentioned that different addition rules may give different results. The Tg for
hot-melt PSAs calculated according to the Fox equation or by the mixing
rule may yield different results [83]. On the other hand it was indicated
that many miscible blend systems exhibit phase separation upon heating,
caused by the existence of a lower critical solution temperature [84]. This
incompatibility caused by heating may influence the tack and peel adhesion
of stored PSAs or the fluidity of PSAs during cutting.
Compatibility is a function of the molecular weight and this statement
is very important for the tackification process of PSAs. Generally, higher
softening point resins are less miscible with the base polymer, and any resin
added after a certain saturation level is reached, does not contribute to
raising the Tg of the system. Instead a second phase consisting of pure resin
or a resin-rich blend is formed. The existence of this incompatible phase
becomes noticeable by the appearance of a second glass transition point.
The structure of the low molecular weight resin is very important to
its compatibility with elastomers, and consequently to its effect on the
viscoelastic properties and performance as PSAs. A completely aromatic
Physical Basis for the Viscoelastic Behavior of PSAs 107

resin such as polystyrene exhibits poor compatibility with natural rubber,


but is compatible with styrene-butadiene rubber (SBR). A cyclo-aliphatic
resin such as poly(vinylcyclohexane) is compatible with natural rubber and
is incompatible with styrene-butadiene rubber. An alkyl-aromatic resin,
such as poly-t-butyl styrene is incompatible with both elastomers.

Use of Blends of Viscoelastomers. The most used method for


suppliers to regulate the Tg of adhesive raw materials is the mixing of
viscoelastomers having different glass transition temperatures. As discussed
in [85] the most common solvent-based acrylates for protective films have a
Tg of 20 C. Such formulations generally contain at least two components.
The base polymer is harder (with a Tg of about þ7 C). As suggested in [86] a
nontacky but flexible acrylic copolymer, with a Tg value of 14 C can be
used as a ‘‘cohesion improving agent’’ for PSA formulations. Polyisobutene
is manufactured from oily to hard products with a molecular weight of
1000–200,000. Mixtures of polyisobutene with high and low molecular
weight should be used to regulate the adhesion-cohesion. As discussed in
detail in [87] the so-called formulation by mixing of high polymers includes
also the use of plastomers.

1.4 Correlation Between the Main Adhesive, End-Use, and


Converting Properties of PSAs and Tg
As discussed earlier, the Tg characterizes the rheological behavior of the
polymers and of the adhesive formulation, thus it influences the end-use
and converting properties also.

Adhesive Properties and Tg


Room and elevated temperature cohesive strength, adhesion to various
substrates, flexibility, etc., are all related to the film-forming ability of the
adhesive (i.e., to its Tg) [88]. The viscoelastic properties of the polymer film
determine its adhesion-cohesion balance (i.e., its adhesive and end-use
properties). The value of the glass transition temperature determines the
main adhesive properties. An adequate choice of the Tg allows optimization
of the adhesive and end-use properties for a specific application area.
Generally, low Tg values ensure very high tack, whereas medium ones give
an optimum peel combined with an acceptable cohesion. The holding power
has been shown to be controlled by zero shear viscosity which is related to
Mw, Me, and Tg by the following equation [58]:

log o ¼ 3:4 log Mw  2:4 log Me þ A=ðT  Tg þ 70Þ þ B ð3:6Þ


108 Chapter 3

where A and B are constants. This equation shows that while o is reduced
by Me, it is also increased by the Tg of the polymer (since it is more rigid).
For polymers exhibiting higher Me and glass transition temperature, we can
expect that PSA peel, tack, and holding power can all be improved. The Tg
appears particularly useful when comparing the flexibility of several latexes,
that is, the latex grade with the lower Tg can be expected to impart a softer
bond at a given temperature and to remain flexible at lower temperatures
(see Chap. 8. on the choice of latexes for general/deep freeze labeling).
But care should be exercised when comparing latexes belonging to
different polymer groups, as the nature of the polymer can also influence
the flexibility. On the other hand a more complex situation arises by the
synthesis of the new core shell latexes where particle structure is important.
Here it is possible to identify the composite structure of the latex by
determining the Tg and the MFT. The optical examination of the cross-
section may show a random distribution of the soft and hard polymer
phases, but unusual Tg values indicate the special structure. It must be
stressed that Tg values of adhesive latexes are necessary but not sufficient
to characterize these polymers. Particle characteristics (morphology) are
also important.
Furthermore the Tg is not an absolute measure of an adhesive’s
suitability for use, but it is a good predictor. Once a material reaches
temperatures below its Tg, the viscous component of its viscoelastic
character is eliminated [89]. Therefore it can be explained why silicone-
based PSAs suffer no delamination at 75 C, like acrylics do at 25 C.
The Tg range of silicone-based PSAs is situated at 100 to 125 C, while
organic PSAs possess a Tg range of 40 to 25 C.
The effect of the Tg on the peel adhesion at constant molecular
weight was studied by Kendall et al. [3]. They stated that the peel adhesion
increases with increasing Tg (at constant tackifier resin level) and passes
through a maximum. This maximum occurs at a Tg of about 55 C. In
some instances maximum peel strengths are apparent at constant Tg and
increasing resin level. Schrijvers [40] found that peel and tack increase
with increasing Tg.

Converting Properties and Tg


Converting and laminating properties of the adhesive refer to its ability to be
coated in order to form a continuous film on the face material. The fluidity
required to coat the adhesive may be achieved using different procedures.
Through the use of higher temperatures (hot-melt PSAs) or dispersing
agents the viscosity of the adhesive can be reduced to a convenient level.
Practically, the dispersed (or dissolved) adhesive will form a continuous film
Physical Basis for the Viscoelastic Behavior of PSAs 109

after drying, if its particles are soft enough to be deformed, to exhibit


coalescence only.
Coalescence is only possible above the Tg. Near this temperature the
coalescence is theoretically possible, but the viscosity of the polymer remains
too high to achieve full coalescence. The coalescence depends on the
viscosity of the bulk polymer in the polymer particles [90]:

 ¼ 3  tð    dÞ1 ð3:7Þ

where  is the contact angle of the particles (Fig. 3.1), g is the surface tension
of the dispersion,  is the viscosity of the bulk polymer in the particles, d is
the particle diameter, and t denotes time.
Pressure-sensitive adhesives are very low-Tg polymers in a continuous
fluid state with low viscosity. Thus PSA coalescence above the Tg is given at
all temperatures. Therefore for these materials the definition of a minimum
film-forming temperature has no sense nor importance. For adhesives other
than PSAs the negative influence of the surface active agents (decrease of
the surface tension, see above formula) and the much higher viscosity in the
vicinity of the Tg leads to a higher MFT value than the Tg. Thus between the
glass transition temperature of these adhesives and their MFT, discrepancies
exist [91], that is:

Tg 6¼ MFT ð3:8Þ

Figure 3.1 Coalescence of polymer particles.


110 Chapter 3

These discrepancies are due not only to the dispersion nature of the
adhesive, but also to the test methods used for measuring the MFT;
the MFT depends on particle size also. Excellent film-forming properties are
obtained with small particle size dispersions. On the other hand, the particle
size dependence of the MFT, makes it sensitive towards the face stock
porosity. The MFT measured on porous surfaces (real conditions) is higher
than the standard value measured on aluminum [92].

End-Use Properties and Tg


It is well known that different adherents require different adhesives (i.e.,
harder or softer formulations). Adhesives for vinyl substrates have a Tg
value of 10 C in order to obtain an optimum bonding [5]. For adhesion
to wooden plates a Tg of 3 C is required. Steel can be bonded optimally
with an adhesive exhibiting a Tg of þ7 C. The bonding technology will also
influence the Tg range. Wet laminating needs Tg values between 40 and
þ15 C, bonding by pressure requires a Tg range of 5 to þ15 C.
Bond forming at high temperatures is possible for adhesives with a Tg
range of þ5 to þ 100 C, and low energy curing (UV, electron beam) requires
a Tg range of 40 to þ30 C. As discussed earlier, PSAs generally display
a Tg range lower than 20 C.

2 ROLE OF THE MODULUS IN CHARACTERIZING PSAs

The Tg is an important characteristic of PSAs, allowing the selection of raw


materials for PSA applications. Its value defines the tack of PSAs; a low Tg
is a prerequisite for tacky materials. On the other hand the Tg alone does
not permit us to obtain a real image of the adhesive performance. As an
example, Zosel [57] showed that poly(ethyl hexyl acrylate) (PEHA) and
polyisobutylene (PIB) have about the same Tg values (64 and 60 C,
respectively), but PIB is less tacky because its modulus is 20–300 times
higher than that of the PEHA. On the other hand hydroxyethylmethacrylate
(with a Tg of þ55 C) and ethyl hexyl acrylate have the same influence
on the adhesive properties of radiation cured formulations. Thus it can
be concluded that knowledge of both Tg and modulus is necessary for
characterizing PSAs. The interdependence between the creep compliance
and tack was demonstrated by Dahlquist [93,94].
As discussed earlier (see Chap. 2) PSAs may be described as
viscoelastic fluids, with the plastic, viscous part of the system characterized
by viscosity (viscous flow occurring above the Tg) and the elastic
deformation governed by the modulus. The Tg and modulus both
characterize the adhesive flow and its mechanical response. During
Physical Basis for the Viscoelastic Behavior of PSAs 111

application of PSAs bonding occurs first; bonding implies adhesive


deformability and wetting of the substrate. Deformability by wetting
includes the building of contact areas, anchorage, and penetration. Thus:

Bonding ¼ f ðfluidityÞ ¼ f ð1=EÞ ð3:9Þ

Anchorage ¼ f ðfluidity, 1=fluidity Þ ¼ f ½ð1=EÞ , E  ð3:10Þ

Penetration ¼ f ðfluidityÞ ¼ f ð1=EÞ ð3:11Þ

Good wettability to the adherend relates to a low elastic modulus (G1).


Fukuzawa and Uekita [95] examined the modulus of various pressure-
sensitive products and their components, and the adhesive performances
related to modulus. For instance they found G1 values of 1.42–2.04E þ 5
for labels, 1.3–1.68E þ 5 for Kraft paper, 1.42–2.09E þ 7 for OPP, and
6.60E þ 4 for cellophane. Ideally the most important property of an
adhesive is the bond strength, or the debonding resistance. The bond
strength increases with the increase of the modulus:

Bond strength ¼ f ðEÞ ð3:12Þ

This general statement was made for adhesives, using more or less complex
mathematical correlations. In Chap. 2, Eq. (2.15) showed that the flow limit
is correlated to the value of the modulus [96]. A formal physical significance
of the  b/"b ratio (ultimate strength-break elongation), is an apparent
ultimate tensile modulus at polymer fracture. PSA polymers obey also
Dahlquist’s criterion of tack, which establishes that PSA elasticity requires a
modulus value of the order of 105 Pa. The fact that the  b/"b ratio is
2–9  105 Pa for several PSA polymers allows us to appreciate the physical
meaning of Dahlquist’s phenomenological criterion [97]. At the most
fundamental, molecular level, Dahlquist’s criterion of tack specifies the
ratio between cohesive interaction energy and free volume (see above, the
correlation between Tg and free volume) within pressure-sensitive polymers,
since the cohesive toughness and free volume are the fundamental charac-
teristics of any polymer materials.
Earlier studies [45,98–100] have shown that the modulus G0 value at
low frequency can be related to the wetting and creep behavior (bonding) of
the adhesive and the modulus at high frequency (100 rad/sec) can be related
to the peel or quick stick (debonding). Using the relationship developed by
Chu [99], general correlations were developed between tape properties
112 Chapter 3

(release and peel adhesion). The peel strength is proportional to the ratio
of G00 and G0 at the respective debonding and bonding frequencies [58]:

P / 1  G00 ð!1 Þ=G0 ð!2 Þ ð3:13Þ

Since PSA bonding typically occurs at the plateau region of the viscoelastic
curve, the bonding is facilitated by lowering the G0 at this region. G0 can be
effectively lowered by increasing the Me of the polymer. PSA debonding,
however, happens at a much higher frequency, typically between
100–12,000 rad/sec. The loss modulus at this frequency is in the transition
region of the viscoelastic curve and its value is highly dependent on the Tg
of the polymer. Increasing the Tg results in higher G00 in this region.
As discussed in the next chapter concerning contact physics and
mechanics, debonding of pressure-sensitive materials is characterized on
a macroscopical (visible) scale by cavitation build-up and fibrillation. The
driving force for the cavitation process in a predominantly elastic medium is
the release of elastic energy [101,102]. If this idea is extended to a viscoelastic
material, one expects the critical stress for unstable growth of cavities to be
proportional to the real G0 of the dynamic shear modulus of the adhesive.
The maximum stress is the parameter most directly related to the cavitation
process.
Pressure-sensitive adhesives intended for similar applications also
exhibit similar rheological properties [103]. The correlation between
adhesive properties and dynamic shear storage modulus is quite good.
The above given short examination of the dependence of bonding and
debonding on the modulus clarifies the role of the modulus as the main
parameter influencing the adhesive properties. Detailed theoretical consid-
erations about the influence of the modulus on the rheology of the adhesion
were provided in Chap. 2. Chapter 6 describes the dependence of adhesive
characteristics on the modulus.
Dynamic mechanical analysis (DMA) currently is used in many
application areas, especially in the adhesive industry. The relationship
between rheological properties, as measured by DMA, and end-product
performance was defined for many types of adhesives, particularly PSAs (see
0
Chap. 1, Section 1). The storage modulus G is a measurement of hardness
[Eq. (2.9)]. At G’ ¼ E þ 0.8 Pa and above, the hot-melt PSA is approaching
the glassy region. A typical value of G0 for a PSA at 23 C is between 1 E þ 0.5
and 2 E þ 0.4 Pa [104]. The loss modulus G00 is associated with energy
dissipation; the greater the value related to G0 , the more flexible the PSAs. Tan
 is the balance of the viscous/elastic behavior; in practice it can be correlated
to the cohesive strength of the adhesive; for hot-melt tan  ¼ 1 is a limiting
Physical Basis for the Viscoelastic Behavior of PSAs 113

value. Above this value the hot-melt PSA can flow without induced
deformation; below this value the hot-melt PSA has inherent cohesion.

2.1 Factors Influencing the Modulus


Generally the modulus is a material characteristic, but for materials in a
composite structure it may depend on the structure and component
characteristics of the composite. An example is that the elastic modulus of
epoxy coatings is higher than that of the uncoated material [92].

Material Characteristics
For common plastics the modulus is not a constant as it depends on the
molecular weight distribution (MWD) and on the manufacturing process
[48]. This is generally valid for adhesives and PSAs in particular. Here the
modulus is a function of the molecular weight and its distribution, the
chemical composition, and the structure. It should be noted that in certain
cases, when evaluating adhesive characteristics, the creep compliance (1/E),
not the modulus, and its dependence on the polymer need to be considered
(see Chap. 2, Section 2). As an example, special PSAs (e.g., skin adhesives)
use grafted polymers; the number and composition of the attached
polymeric moieties on the polymer, and the inherent viscosity (molecular
weight) of the polymer are manipulated to provide an adhesive composition
with a creep compliance value greater than 1.2  105 Pa1 [6].

Modulus and Molecular Weight. Molecular weight is a parameter


influencing the most important practical properties such as tensile strength,
elongation, modulus, Tg, viscosity, green strength, wettability, pot life,
chemical resistance, and processability of the adhesives. For instance, it
has been shown for PIB, that tack and adhesive fracture energy decrease
monotonically for MW > 45 kg/mol [105,106]. Schlademan and Bryson [107]
examined the effect of MW and aromatic content of the resin, for tackified
formulations. As stated, the number average molecular weight showed a
strong correlation with shear adhesion values. On the other hand, the
molecular weight effect on the molten viscosity was small. It was difficult to
determine whether the effect is primarily due to a direct viscosity (modulus)
influence, an indirect compatibility (entropy) influence, or both.
There is a plateau of the modulus corresponding to rubbery/elastic
behavior of the material as a function of the temperature. The existence of
this plateau and its position (Tg) depend on the molecular weight. If the
molecular weight and the molecular weight between entanglements of the
polymer do not attain a critical value, no rubbery, elastic plateau appears.
114 Chapter 3

That means that for low molecular weight polymers a ‘‘pressure-sensitive’’


bonding may appear, but pressure-sensitive debonding properties do not
exist.
Generally PSAs are polymer blends, thus the influence of the
molecular weight on the modulus must be taken into account from the
point of view of the mutual solubility — compatibility — (see Section 1.3
and Chap. 8, Section 2.3).
Considering the tackified elastomer as a polymer solution in the resin,
and the polymer solution as ideal (zero excluded volume effect), the plateau
modulus ðG0n Þ decreases as a function of the volume fraction of the polymer
(Vp), the density of the blend ( ), and the molecular weight between the
entanglements (Me) according to the following equation:

G0n ¼ ð RT =Me Þ  Vp2 ð3:14Þ

where R is the universal gas constant. The gas constant appears in the
correlation due to the hypothesis that the ideal rubber behaves like an ideal
gaseous conformation network. The role of the Me for the Tg was discussed
earlier. Me can be calculated from the plateau modulus. For instance, as
found by [108], in the rubbery plateau region, the storage moduli G0 (!, T) of
all poly-n-butylacrylate polymers (having different molecular weight) were
constant at about G0n  0.15 MPa, thus the average molecular weight
between entanglements was  22 kg/mol. In principle the failure of the fibril
may occur by: disentanglement of the polymer chain, fracture of the
polymer chain, or debonding of the fibril from the adherend. In the first
process the maximum strain of the fibril "max is controlled by the product
 dd"/dt where  d is the terminal relaxation time of the polymer ( / M3 and
decreases with increasing temperature), and d"/dt is the local strain rate
(d"/dt / Vdeb) [108]. This means that both the fibril stress  fibril and "max
increase continuously as VdebM3 increases because disentanglement becomes
more difficult. If the failure mechanism is not disentanglement but a
debonding process, the critical stress  sustainable by the fibril-adherend
interface is the rupture criterion. One expects  r to be constant in this regime;
"max decreases with VdebM3 since "max /  r/VdebM3. An entanglement model
can be used to estimate the critical molecular weight using the characteristic
ratio of the polymer, the number of bonds per monomer, and the monomer
molecular weight.
A classical example of the modulus dependence on molecular weight
is given by rubber/resin formulations. Here nonmilled rubber from dried
natural latex shows a larger plateau modulus and a higher elastic modulus
(G0 ) than milled natural rubber. Milled rubber possesses a lower molecular
Physical Basis for the Viscoelastic Behavior of PSAs 115

weight because of the mechano-thermal degradation of the elastomer. For a


CSBR latex, the increase of the molecular weight is accompanied by an
extension of the plateau region of the dynamic modulus in the master curve
towards higher temperatures [62]. Although the modulus increases with
increasing molecular weight, its adjustment by the molecular weight remains
limited. It was shown [2] that the maximum of the modulus is situated at
about 500–1000 atoms in the main polymer chain. Longer molecules do not
give higher strength because their various segments act independently of
each other. On the other hand, an excessive increase of the MW is not
recommended because of the poor wetting out (low flow) and decrease of
the ability of the adhesive to desorb energy (adhesive break). It was shown
earlier that Tg related parameters (chain mobility) limit the molecular weight
as well.
Another parameter is the molecular weight distribution, which may be
characterized by dynamical rheology (i.e., the rheology is very sensitive
towards MWD) [109]. This sensitivity is also valid for PSAs; in this case the
MWD influences the loss tan  peak. McElrath [110] demonstrated that the
molecular weight (and its distribution) of a tackifier influences the peak of
the loss modulus and its location as a function of the frequency. For
tackified formulations the loss tan  peak is sharper for those resins with a
narrow MWD. It is very broad and diffuse for the broad MWD obtained by
blending two resin fractions [33]. In a diagram representing tan  peaks for a
resin-rich phase (on a log scale) the broad peak (large MWD, resin mixtures)
does not equal zero. At zero, the loss modulus equals the storage modulus.
Above zero there is more energy loss than storage. On the other hand, for
tack it is important to dissipate energy during the debonding step in order to
achieve high tack. Energy dissipation is higher for a broad MWD (i.e., for
resin mixtures); this consideration is very important for the choice of the
tackifier resins. Commercial acrylic adhesives like PEHA have a very broad
spectrum of relaxation times. In such products low MW fractions provide
the fast relaxation times necessary to achieve a large real contact area within
a short contact time (typically 1 s). As shown in [108] in the case of
monodisperse poly-n-butylacrylates these low MW fractions are absent, and
the contact area depends mostly on the elastic properties of the polymer at
bonding frequencies. One can conclude that cavitation by debonding of
PSAs is indeed governed by the viscoelastic properties of the adhesive and
the low molecular fractions in PSAs are important for the formation
of the contact. Experimental results have shown that a large spectrum of
relaxation times is a necessary condition for the performance of PSAs.
If short relaxation times are absent (i.e., no low MW fraction) the contact
with the substrate is not achieved within short contact times, while if long
relaxation times are absent, a fibrillar and adhesive fracture cannot be
116 Chapter 3

achieved. It is to be noted that the role of the relaxation times is taken into
account in the modified form of the Kaelble equation (for the peel) which
can be presented in the form [111]:

P ¼ b  l ½ða  N    DÞ=12RT f2 ð3:15Þ

where  f is the critical tensile stress of polymer at fibril failure, D is the self-
diffusion coefficient of the polymer segment, b is the width and l the
thickness of the adhesive layer, a is the size of the diffusing polymer segment,
N is the Avogadro number, and  is a segmental relaxation time.
The influence of the molecular weight and MWD on the modulus
should be taken into account for rubber blends too. For instance, it was
shown [112] that mastication (used for natural rubber) will have a relatively
small effect on long chain molecules and on high molecular weight elements.
If they survive mastication to some extent, then their relative contribution
to the elasticity will increase. Therefore in mixtures with fractions having
different molecular weights, the influence of the high molecular weight
factors is more pronounced. For tackified blends the molecular weight of the
resin is also significant for the compatibility. In this case an optimum of
the viscoelastic properties is linked to molecular weight below 1000 with a
narrow MWD. The dependence of the modulus on the mutual solubility was
confirmed by Kendall, Foley, and Chu [3]. However, in certain cases the use
of low modulus polymers is dependent on coating technology. Low viscosity
requirements imply low molecular elastomers as raw materials for hot-melt
PSAs; therefore the toughness of hot-melt PSAs is also low [113].
Opportunities to adjust the molecular weight of the base polymers for
adhesives allow the adjustment of the modulus as well. Theoretically the
molecular weight may be modified by increasing the polymerization degree
for the linear polymer, or by crosslinking. It was shown earlier that
increasing the polymerization degree is limited because of the decrease of
the flow properties (for processing and end-use purposes). Thus an ade-
quate adjustment of the molecular weight should be achieved through
crosslinking.

Chemical Composition. The modulus is a function of the chemical


composition; polymers based on different monomers possess different
modulus values. For instance, of block copolymers with different rubbery
blocks (but similar block molecular weights and about 30% styrene), the
SEBS had the highest modulus [114]. On the other hand minor
compositional aspects such as sequence distribution also influence the
modulus (see later).
Physical Basis for the Viscoelastic Behavior of PSAs 117

Zosel [57] tested the dependence of the modulus of acrylates on the


side chain length. Polymethyl-, polyethyl-, n-butyl-, i-butyl-, and ethyl hexyl
acrylate were studied in order to estimate the influence of the chain length
(branching) on the Tg and E. Shorter side chains in acrylates impart higher
modulus values, the E ¼ f (temperature) plot being moved to the right
(higher Tg values) with a higher plateau. In vinylacetate-butylacrylate
copolymers the maximum of the curves (peel force-peel rate) corresponds
to the change in mode of failure from cohesive (at lower peel rates and a
VAc content <50%) to adhesive (at higher debonding rates and a higher
VAc concentration) as a result of stiffening by vinylacetate [115].
According to the theory of rubber elasticity the shear modulus GA
depends on the number of crosslinks Vc.

GA ¼ Vc  R  T ð3:16Þ

where R and T are the universal gas constant and the temperature,
respectively. For tackified PSAs the addition of resin, the dilution of
crosslinks, and the destruction of the elastomeric network decrease the
modulus, but the modulus increases again at higher resin loadings. The
influence of the crosslinking through molecular weight increase and
natural aging was discussed earlier. Here another special effect of the
crosslinking on the modulus, due to the lack of interaction between gel
and tackifier, will be examined. Rubber shows a highly reversible
deformability combined with a low modulus (the modulus of unfilled
rubber is 103 lower than that of plastics). Additionally rubber exhibits a
technologically important property, namely the ability to undergo self-
reinforcement [116] through formation of strain-induced crystalline
structures; this imparts excellent tensile properties to rubber. Self-
reinforcement by crystallization is a well-known possibility to improve
the properties (modulus) of rubber-like synthetic polymers used as raw
materials for hot-melt PSAs. In this case soft polyethylene butylene (PEB)
segments (matrix) are reinforced by polystyrene segments at the end of the
chain; this gives physical crosslinking [117]. Chemical crosslinking is
characteristic for CSBR latexes, as they consist of a mixture of polymer
species with linear chains, branched chains, and a crosslinked network.
Many natural rubber solvent-based adhesive formulations also contain a
crosslinking agent. The effect of crosslinking may be illustrated by the
increase of G0 (elastic modulus). The adhesive with more crosslinking
possesses a higher modulus and higher shear holding power.
Classical rubber elasticity theory predicts that the plateau modulus G0n
for a diluted polymer is proportional to the square of the volume fraction of
118 Chapter 3

the polymer. Scaling law interpretations by De Gennes [118] suggests that


G0n should be proportional to the 2.25 power of the polymer volume fraction
[see correlations (3.16) and (3.18) also]. Plots of SBR blends indicate higher
values of 2.45–2.65 for the exponent. A possible explanation is that the SBR
contains gels (see Chap. 5) which do not participate in the dilution by the
resin. Therefore the resin concentration in the amorphous phase is higher
than expected, resulting in an apparent higher power for the polymer
volume fraction, and as a practical consequence for SBR tackification in a
higher tack. The influence of the tackifier concentration on the modulus was
studied in [39]. It was shown that the minimum G0 value (at 25 C) occurs
at about 40–60% resin loading for rubber/resin systems. For a detailed
examination of the modulus dependence on tackifiers, the reader is referred
to Chap. 8.

Dependence of Modulus on Sequence Distribution. The sequence


distribution exerts a strong influence on the modulus of block copolymers
used as raw materials for hot-melt adhesives. By changing the sequence
length, volume, and polarity, it is possible to modify the modulus, as well as
the adhesion/cohesion balance of the polymers; here polar styrene blocks
reinforce these polymers [117]. It should be mentioned that a similar
hardening effect of certain monomers was observed in some statistical
copolymers as well [119]. For partially hydrolyzed ethylene vinyl acetate
(EVAc) copolymers the dependence of the modulus on the distribution of
the hydroxy groups was shown [120].
Theoretically SBCs have a lower modulus value (106 Pa) in
comparison to unvulcanized rubber (107 Pa). Therefore, in principle,
their tackification seems to be easier. Really, because of the segmented
construction of such elastomers, their tackification (as the increase of the
tack) concerns the diene sequences only, i.e., only the ‘‘half ’’ of the
molecule can be tackified. However, the most important practical use of
sequence regulation occurs in the synthesis of thermoplastic rubbers used
as raw materials for hot-melt PSAs. These materials are comparatively
low molecular weight polybutadiene and polyisoprene rubbers, inter-
twined with short polystyrene chains. When they are cooled from the
melt the (incompatible) polystyrene end blocks separate out and gather
to form small hard domains of polystyrene. Since the polystyrene blocks
are covalently bonded to the rubber, phase separation is restricted; this
forces the copolymer into a special microphase morphology [121] (see
Chap. 5). Special sequence distribution allows increasing of the
molecular weight without the loss of the tack. Thus segmented EVAc
copolymers with very high molecular weight are adequate raw materials
for hot-melt PSAs [122].
Physical Basis for the Viscoelastic Behavior of PSAs 119

Softer, more tacky adhesives are obtained with polymers with a low
polystyrene content such as Cariflex TR 1107 [123]. Some applications,
however, require a high shear resistance (i.e., a higher polystyrene content).
For lower viscosity hot-melt PSAs, which are easier to process at lower
temperatures, SIS block copolymers with high melt flow rates containing a
relatively high proportion of diblocks were developed. These low modulus
polymers also promote better die-cuttability of PSAs used in label stock
applications. It can be expected that in the future SIS polymers will be
designed with even more variations in the diblock/triblock ratio, which will
enable the manufacture of still softer hot-melt PSAs with a more aggressive
tack. The higher modulus polystyrene-polybutadiene-polystyrene (SBS)
block copolymers are mainly used in nontacky adhesives. For tacky hot-
melt PSAs with better shear, low MW, low viscosity SBS grades with a high
polystyrene content are used [2].
The size and morphology of the polystyrene domain in styrene
block copolymers influence the cohesive strength of the system. However,
if one increases the polystyrene domain by increasing the polystyrene
content of the copolymer the storage plateau modulus increases, thus
making the polymer considerably more difficult to tackify [124]. On the
other hand it is well known that in tape applications the diblocks act as
an elastically ineffective diluent in the formulated adhesive. A high
diblock content limits the cohesive strength of the base polymer of the
adhesive system. The tensile strength and the modulus decrease as the
diblock content increases. The diblock has a molecular weight which is
half that of the triblock. Increasing the SI content is equivalent to
replacing one triblock with two diblocks. This loss of connectivity causes
a loss of elastic strands. The measured  max in probe tests decreases
as the amount of diblock increases. In [65] the shear properties of
styrenic block copolymers were measured in the linear viscoelastic regime
(small strain) on a parallel plate rheometer. As demonstrated, differences
due to the change of the SI content occur at 5  103 to 5  101 rad/sec,
where an increase in the SI content leads to a decrease in the elastic
modulus G0 , and to an increase in the loss modulus G00 . In this range
the blends are more dissipative if there is more SI. Tensile tests show
that decreasing the SI content or increasing the crosshead velocity leads
to a higher stress for all extensions and to a more pronounced strain
hardening. For label stock adhesives the presence of diblocks is currently
considered desirable for obtaining the die-cutting and skeleton waste
stripping performance. The optimum level of required diblock depends
upon the type and speed of the converting machine, the label stock face
material used, together with the other components in the adhesive
formulation.
120 Chapter 3

Modulus and Tg
The tensile strength and elastic modulus decrease above Tg [125]. This
general observation has a special implication for polymer blends. For
certain polymer blends a phase separation temperature exists, the value of
which depends on the Tg. A different dynamic mechanical response is
observed below and above this phase separation temperature (Tc). For
temperatures below Tc a thermorheologically simple behavior is observed
and the time/temperature (frequency/temperature) superposition principle
can be applied [126]. For these temperatures master curves are obtained
according to:

Gred ðTo , !Þ ¼ GðT, !Þ  To  o  Tg1 ð3:17Þ

where Gred is the modulus (G0ðxÞ or G00ðxÞ ) reduced to the reference temperature
To. For polymer blends where no strong interactions occur between the
polymers (i.e., one of the components acts as a low molecular weight
solvent) the plateau modulus Gon2 is given by the following equation:

Gon2 ¼ Gon2 ð¼1Þ  2:25


A ð3:18Þ

where A is the polymer volume fraction. For phase-separated mixtures (and


chemically crosslinked systems) an !1/2 power law was observed (usually
G0  !2 and G00  !).
Filler Effect on the Modulus. The use of fillers with elastomers yields
increased stiffness, modulus, and rupture energy [127]. In the presence of
reinforcing fillers the elasticity modulus of elastomers at small and moderate
deformation increases in the first approximation according to the Guth-
Smallwood equation:

E ¼ Eo ð1 þ 2:5 þ 2 Þ ð3:19Þ

where Eo is the modulus of elasticity of the unfilled elastomers and  is


the volume fraction of the filler. This effect of the filler is due to a
hydrodynamic viscosity increase, an increased stiffness, and a lower Tg of
part of the rubber phase at the interface with the filler, given by surface
interaction. A covalent bonding of network segment to filler surface, and a
tightening of the short chains between aggregates must also be considered.
For filled elastomers the mechanical work of deformation (the change of
entropy and internal energy) should include the parameter . The filler
particles, which may be as small as 1.5  105 mm (about 30 polymer chain
Physical Basis for the Viscoelastic Behavior of PSAs 121

diameters), may also interface with the motion of the polymer molecules
[128]. Generally the modulus increases by increasing the concentration and
activity of the fillers [129]. Fillers increase the modulus proportional to
their concentration [130]. The relative modulus increase was calculated as
a function of the concentration. These correlations are calculated for fillers
acting as chemically inert substances. Proactive fillers (such as metallic
oxides or salts) may interact with the base polymer, actually crosslinking
it. This addition of some oxide to carboxylated compounds dramatically
reduces the pressure-sensitive nature of the adhesive [131]. On the other
hand, tackifier resins act partially as fillers. Exhibiting a surface free energy
close to that of the elastomer, the filler is perfectly wetted by the organic
matrix and, as a consequence, good dispersions (i.e., lower viscosities) are
obtained [127].
In the case of PSAs fillers can have a negative effect on the tack, but
they are sometimes added to provide additional functionality. As
demonstrated in [132], a filler volume fraction of 0.5 substantially stiffens
the adhesive, while also reducing the overall deformation. An approximate
characteristic value of the energy release can be obtained assuming that
the nonlinear effects arise from the presence of a plastic displacement
in addition to the elastic displacement (by tack) corresponding to linear
elasticity. The overall adhesion energy is much larger than Gest (energy
release rate), indicating that most of the energy is dissipated by bulk
deformation of the adhesive, with a relatively small fraction available
for crack propagation. Filler particles reduce the energy release rate
required for crack propagation, but have a more complicated effect on
the deformability of the adhesive, and on the resultant work of adhesion.
The overall extension of the adhesive initially increased with increasing
filler concentration, but was reduced substantially at the highest
concentration.
Crystalline polymers are two-phase systems also consisting of both
crystalline and amorphous regions. They can be considered as networks in
which crystallites are the solid filler and act as multifunctional crosslinks.
Such structures exist in paper, they may arise by strain-induced crystal-
lization in special elastomers, or in plastomers used as carrier materials,
as discussed in detail in Chap. 2, describing the rheology of the carrier
material [133].
As is known, multiphase rubbery polymers can be considered as elastic
networks or filled systems. The stress-strain behavior of such polymers
can be adequately modeled by the Mooney-Rivlin equation as an elastic
network (considering the trapped entanglements as finite crosslink
junctions) or using the Guth-Gold equation relating the stress level
(modulus) to the hydrodynamic effects of the styrene blocks, considered
122 Chapter 3

as a filler. In this case the plateau modulus is given by the equation [see
correlation (3.18) also]:

Gon ¼ ð =Me Þ  RTð1 þ 2:5 þ 14:12 Þ ð3:20Þ

According to [134] after drying an emulsion-based PSA the adhesive


behaves as a filled polymer melt. Even though emulsion particles are
internally crosslinked to get the appropriate viscoelastic behavior, the
particles only interact via entangled dangling ends at the interface between
particles.

Crosslinking Effect on the Modulus. To increase the peel energy it is


common to add a slight degree of crosslinking. By manipulating the gel/sol
ratio a higher performance adhesive can be designed. On the other hand,
crosslinking is generally influenced by the initial molecular weight of the
polymer. Depending on the network mesh size, additive molecules (e.g.,
tackifier) are either trapped (low Mc) or not trapped (high Mc) in the
network. Such segregation is important for the modulation of the adhesive
properties. Most polymers have a built-in crosslinking, in other cases
postsynthesis crosslinking is carried out. For instance, research has shown
that emulsion polymerization of acrylates carried out to complete
conversion, produces significant amounts of microgels inside the particles
due to chain transfer to the polymer via hydrogen abstraction [135]. As
stated by Van Holde and Williams [136] high gel content could compensate
for low molecular weight. This was not true of the peel behavior. As
observed, adhesion failure was favored by high molecular weight, high gel
content, and high peel rate. Cohesive failure leaving tacky adhesive on the
substrate was favored by low molecular weight, low gel content, and low
peel speed. When the molecular weight was low, but the gel content was
high, a thin residue on the adherend was observed. The creep compliance
both at short times and at long times is governed by a combination of the
initial molecular weight and the gel content. The compliance decreases and
the viscosity increases as the molecular weight increases or the gel content
increases. The effect of gel content becomes less important at the high
molecular weights. The molecular weight of the soluble portion of partially
crosslinked adhesives is always dominated by the initial MW, even at high
gel content. Creep resistance can be achieved either through molecular
weight increases or gel content increases and to a certain extent one can
compensate for low initial molecular weight with greater crosslinking as far
as creep resistance is concerned.
Physical Basis for the Viscoelastic Behavior of PSAs 123

Dependence of the Modulus on the Tackifier Resins. It was shown


earlier that the use of tackifier resins lowers the plateau modulus of adhesive
blends. This effect depends on the nature and concentration of the tackifier.
The dependence of the modulus on the tackifier concentration was shown
by Sherrif et al. [63] and Kamagata and Kosaka [137]. Zosel [57]
demonstrated that the modulus reaches a minimum for 50% resin in the
mixture (natural rubber/glycerine ester of colophonium). Ullmann and
Sweet [138] investigated silicone PSAs. As stated, a minimum concentration
of resin is required to tackify the polymer so that it exhibits PSA properties,
however too much resin will result in poor tack, and loss of adhesion. As the
level of the resin was increased, the storage modulus G0 increased in the
entire frequency range. Good adhesion for tapes is achieved when G0 is low
at low frequency rates, i.e., 0.1 rad/sec, and a relatively high slope exists for
G0 as the frequency is increased [139]. Increasing the modulus G0 results in
adhesives with higher adhesion and lower tack. Adhesives with lower creep
compliance (higher G0 ) will have better cohesive strength (creep resistance).
However, if the modulus is too high, the adhesive may not rapidly wet
the substrate.
As discussed earlier the molecular weight and MWD of the tackifier
resin strongly influence the modulus of the mixture. According to the filler
theory, the modulus of a diluted polymer (e.g., thermoplastic rubber in resin
and oil) is given by:

Gon ¼ 22 ð =Me Þ  RTð1 þ 2:5 þ 14:12 Þ ð3:21Þ

where  is the fraction of the polymer in the polydiene phase and 2 is the
fraction of the polystyrene block in the entire composition [see Eq. (3.20)
also].
The modulus depends on the plasticizer too. Low molecular weight
fractions can play the role of the plasticizer also. As stated in [140] the
fraction of the polymer that has the molecular weight less than twice the
entanglement molecular weight (Me) will act as a plasticizer.

Dependence of the Modulus on Other Factors


The modulus of adhesive joints depends on the adhesive’s thickness [141]. In
fact increasing the thickness of the adhesive layer decreases its modulus.
This behavior may be explained by the special multilayer structure of the
adhesive and the influence of the surface of the solid state components.
The modulus of the laminate varies close to the interface with the adhesive.
The width of this zone depends on the adhesive; for certain adhesives it
varies between 0.04–0.2 mm [142].
124 Chapter 3

It must be noted that the investigations on the influence of the


modulus on the joint resistance, the dependence of the modulus, and
adhesive joint resistance on the adhesive layer thickness were mostly carried
out for thick-wall sandwiches and adhesives other than PSAs. However, it
should be noted that the thickness (coating weight) of the adhesive would
also exert a special influence on the adhesive resistance for PSA laminates.
Increasing the coating weight increases the thickness of the mobile middle
adhesive layer only (i.e., the cold flow of the adhesive). According to the
paradox of Griffith [143] a fiber-like material possesses a much higher
strength than the same material in another form. The paradox of tensioning
or stressed length influences the effect of the thickness of the adhesive and of
the primer also. Another parameter influencing the modulus of plastics is
their degree of orientation [104]. Actually there are no data available about
a similar modulus dependence for PSAs. One should note the dependence
of the modulus on the composite structure of the laminate or laminate
components. Special attention must be paid to the influence of humidity on
the modulus of the paper.
The modulus of the paper depends on its chemical composition,
temperature, physical structure, and moisture content. When analyzing the
elastic properties of paper, it is necessary to take into account the fact that
paper is composed of both crystalline and amorphous cellulose [141], the
latter undergoing softening as the moisture content increases. The effect of
moisture on the modulus of cellulosic materials is also affected by
temperature. A moisture/temperature equivalence exists, that is, for lower
moisture contents the softening effect occurs at higher temperatures. The
decrease in relative modulus with increasing moisture content is different for
papers of different crystallinity. Theoretically the modulus is a function of
the number of hydrogen bonds (n) absorbing the applied strain:

E ¼ k  n1=3 ð3:22Þ

where k is the average value of the force constant and E is the elastic
modulus of the paper. The decrease of the effective number of the hydrogen
bonds results from moisture absorption, resulting in the decrease of E.

2.2 Adjustment of the Modulus


On the basis of the interdependence between modulus and adhesion-
cohesion balance of PSAs there is a need to adjust the modulus value. As
discussed earlier, the modulus is a polymer characteristic and depends on
its chemical and macromolecular properties; thus the modulus should be
Physical Basis for the Viscoelastic Behavior of PSAs 125

modified by changing these properties. In a similar manner to the Tg it is


possible to change the modulus of the adhesives during the polymer
synthesis or through formulation.

Synthesis
The choice of the monomers and the control of the sequence distribution
and polymer structure allow an adequate regulation of the modulus. As an
example, the Tg and modulus of the acrylic latexes can be adjusted by
polymerizing monomeric acrylates at different rates [3]. On the other hand
the choice of the adequate rubbery sequence for block copolymers for hot-
melt PSAs allows the modification of the polymer stiffness. Jacob [144]
pointed out that although more economical, there are only a few cases
where the use of SBS block copolymers for hot-melt PSAs applications
appears appropriate. This may be explained by the high modulus of these
compounds compared to SIS, and by the higher stiffness and limited
tackifying ability (compatibility).

Formulation
On the basis of a few adhesive groups/polymers, a large number of PSAs are
compounded according to the very different requirements concerning the
adhesion/cohesion balance. This is possible through the use of formulation.
Formulation refers to the design of the adhesive properties through
compounding. Designing of the adhesive properties involves mainly
tackification (i.e., the increase of tack and peel) or, in a few cases, the
improvement of the shear and cohesion (see Chaps. 6 and 8).
Tackification imparts better flow properties and molecular mobility.
This is achieved by the use of diluting agents, either micromolecular
(plasticizer) or macromolecular (tackifying resins). Thermoplastic rubbers
are unique in the way they bring together both rubber and plastic [145].
They comprise two inherently incompatible, but chemically connected
segments which attempt to separate at service temperatures; this results
in their having a morphological structure in which thermoplastic end
blocks such as polystyrene form domains holding together a rubbery
network (e.g., polybutadiene; see earlier). Materials with fine basic
morphologies can thus be produced.
In addition to these opportunities provided by changing the basic
polymer structure, further modification of the material properties is possible
during compounding. For example, by using low molecular weight
compounding ingredients which demonstrate preferential compatibility
with, or solvency for one or the other of the two phases in the base
polymer, it is possible to change the effective ratio of the thermoplastic to
126 Chapter 3

rubber phase. More specifically it is often possible to move from a


nonresilient morphology to a soft material by increasing the volume of the
rubbery phase through the addition of a suitable oil, or by decreasing the
amount of the polystyrene compatible aromatic resin in an SBS copolymer/
oil/resin hot-melt adhesive formulation. The plasticizing agents typically
decrease both the dynamic viscosity and the elastic modulus of the adhesive
matrix.

Crosslinking
Crosslinking yields a higher modulus and higher shear resistance. A
noncrosslinked soft acrylic polymer has a storage modulus of about 103 Pa,
and this value may be increased to 104 Pa by irradiation [146]. In the range
of the crosslinked adhesives those with the highest modulus are the most
crosslinked and have the highest shear and holding power [147].
Theoretically, on the basis of crosslinking data it should be possible to
calculate the shear modulus of the polymer. Practically, in a few cases it
is possible with the aid of crosslinking data to predict the gel point and
the shear modulus for polyurethanes [148].
The theoretical basis for the dependence of the modulus on the
crosslinking is given by the dependence of the modulus on the polymer
chain network, and on the number of crosslinked sites [119]. For small
deformations (according to Fox and Flory [38], deformation means
transition from a real network to a phantom network) there is a correlation
between the modulus and the number of crosslinked points (entanglement):

G ¼ 2C1 þ 2C2 ¼ ðvCH  vC Te ÞkTðhr2 i=hr2o iÞ ð3:23Þ

where G is the modulus, 2C1 is a function of crosslinking density, ve is the


number of network points in the uncrosslinked polymer (polymer-specific
content), and Te is the trapping factor (i.e., the number of rigid network
points); it is a function of molecular weight distribution and chemical
crosslinking density. hr2 i is the mean square end-to-end distance of a
network in the undistorted state and hr2o i is the mean square end-to-end
distance of the corresponding free chains. Ideally, the chain length of the
crosslinking agents also influences the modulus [149]; longer-chain cross-
linking agents allow more elongation.

2.3 Modulus Values


The value of the modulus at a temperature above Tg characterizes the
viscoelastic behavior of PSAs raw materials. The location of the rubbery
Physical Basis for the Viscoelastic Behavior of PSAs 127

plateau is very important. In order to obtain suitable adhesive properties


at temperatures above Tg (room temperature or less), a low-E plateau is
needed. The dependence of the modulus on the temperature and stress rate
is well known. Thus it becomes very difficult to compare different polymers
on the basis of single E data only. However, these values provide a first
indication of the possible application field of the polymers. Pressure-
sensitive adhesives are soft and mobile under certain conditions. Even the
most crosslinked PSAs do not match any property of a thermoset plastic
as the value of their modulus is low, and hence their films are not meant
to be used without a face material. However, in the last decade carrierless
pressure-sensitive products were developed also. Such products are
described in detail in [150]. First PVC and butyl rubber were used for
carrierless products. Their plasticizing allows a substantial decrease of the
modulus and its regulation. Later special acrylic adhesives were developed.
They are used mostly as mounting and sealing tapes. Their modulus is
increased by filling (with continuous or discrete fillers) and crosslinking.
Therefore, for such products, an adequate peel value can be obtained
with high coating weight only. Table 3.5 summarizes some modulus values
for polymers used for PSAs or related materials.
A maximum tack implies a modulus of 1  105 to 2  104 Pa [151]. For
a good wet-out a modulus of about 3  105 Pa is required [152]. For different
kinds of stresses different modulus values must be taken into account.
Generally cold flow is characterized by the creep modulus Ec:

Ec ðtÞ ¼ =ðtÞ ð3:24Þ

where  is the stress (constant in time) and e(t) is the time dependent strain:

E ¼ f ðT, stressÞ ð3:25Þ

Generally for small deformations (" < 0.8%) the stress dependence of the
deformation may be eliminated. Pressure-sensitive adhesives are subject to
higher deformation levels, thus the stress dependence of the strain cannot be
ignored and the value of the creep modulus is not a constant. On the other
hand, for high deformation rates (e.g., peel or cutting) the dynamic modulus
must be considered:

E 0 ¼ 2G0 ð1 þ vÞ ð3:26Þ
128 Chapter 3

Table 3.5 Modulus Values for Materials Used for PSA Laminates

Material Modulus Value (N/mm2)

Nature Characteristics MD Mean TD Ref.

Butadiene-styrene Bulk — 7–11 — 153


copolymer
copolymer, block, radial — 5 —
copolymer, block, linear — 25 —
Cellulose acetate Film 24–45 — — 154
Ethylene-vinylacetate Bulk, grafted — 280 — 122
copolymer Bulk, 50–70% VA — 200–1600 —
130–700 155
Isoprene-styrene — 2–8 —
copolymer, block, linear
Natural rubber Bulk — 0.9 — 153
Paper — — 2920–3700 — 156
Polycarbonate Film — 2100–2400 — 116
Polyethylene LD, film — 200–500 — 116
HD, film — 700–1400 — 116
MD, film 28–35 —
Polyethyleneterephthalate Film, oriented 50–130 — 155
Polypropylene BOPP, film 1400 — 2200 157
BOPP, film 2400 — 4000 157
BOPP, film 1500 — 2500 157
BOPP, film 2000 — 2500 157
BOPP, film 1200 — 2100 158
Polystyrene Film 3200–3250 116
Polyvinylchloride Hard, unplasticized, 4000 159
film
Plasticized, film 67 — 61 —
Rubber Unfilled 1–5 116

where v is Poisson’s ratio. As discussed earlier (Chap. 2, Section 1.2),


Dahlquist [93] has taken the 1/E value (creep compliance) as an index for
PSAs, showing that for general PSAs this value should exceed 1  l06 Pa1
but a maximum tack requires a lower modulus (i.e., a higher creep
compliance) [93,94]. The creep compliance for PSAs for medical and
surgical applications should be about 1.2  10 5 Pa 1 , preferably
1.3  105 Pa1 [13]; the higher the creep compliance, the bigger the adhesive
residue left on the substrate. The fundamentals of creep compliance as they
relate to polymeric materials and in particular to viscoelastic polymers are
covered in [110].
Physical Basis for the Viscoelastic Behavior of PSAs 129

3 CONTACT PHYSICS

As stated in Chap. 1, Section 1 the adhesive properties of PSAs require a


viscoelastic non-Newtonian flow behavior. The rheology of such materials
is characterized by the time and temperature dependent modulus, and its
components. In application practice such behavior is manifested by the
time, stress, and temperature dependent bonding (e.g., dwell time) and
debonding (e.g., adhesive or cohesive break of the joint as a function of
the stress rate) of the adhesive. To understand the nature of this special
bonding/debonding process, contact physics and mechanics were applied
for its study. The scope of such investigations is to correlate the end-use
adhesive characteristics (e.g., peel strength, tack, and shear resistance) with
the data obtained by rheological and mechanical measurements and
calculations (see Chap. 6 also). Classical mechanics offers well-known test
and calculation methods which can be extrapolated. Thus, simplifying the
problems of pressure-sensitive adhesive bonding and debonding to those of
fracture mechanics on the one hand, and correlating the properties of the
stressed finite polymer elements with the their macromolecular build-up on
the other hand, it is possible (at least theoretically) to explain the practical
behavior of PSAs.
The viscoelastic contact problem can be divided in two phases:
advancing contact or the bonding phase, and receding contact or the
debonding phase. In the bonding phase the contact radius increases with
increasing time, whereas during receding contact, the contact radius
decreases with increasing time. Contact problems were studied first for the
most simple case of elastic materials; later the results of such investigations
were extrapolated for viscoelastic materials.

3.1 Contact Mechanics for Elastic Materials


The study of the bonding phase for elastic materials started some time ago.
For such materials the lack of creep allows the examination of the bonding/
debonding as a reversible process where stress and strain occur simulta-
neously. Derjaguin was the first to consider the effect of elastic contact
deformation on adhesion by assuming that the particle acts as a Hertzian
indenter. The driving force for elastic contact is the work of adhesion.
Maugis and Pollock [160] generalized the adhesion model to include plastic
deformation and predicted that the contact area was determined by the
hardness H of the deformed material.

F þ Fa ¼ a2 H ð3:27Þ
130 Chapter 3

where F is the external force and Fa is the adhesion force between particle
and surface. The adhesion energy for the contact mechanics data is obtained
using the following equation:

G ¼ fða3 KÞ=ðR  PÞg2 =6a3 K ð3:28Þ

given by Maugis [161].


The adhesion to elastomeric substrates generally can be described
quite well by the Johnson-Kendall-Roberts (JKR) theory [162,163].
According to the JKR theory [161,164] the contact radius a between
contacting spherically symmetric bodies under an applied load is given by:

a3 ¼ ðR=KÞfP þ 3WR þ ½6WRP þ ð3WRÞ2 1=2 g ð3:29Þ

where W is the work of adhesion, P is the applied (contact) load, and Ri, Ei,
and i are the radius of curvature, modulus, and Poisson’s ratio of the ith
cylinder (lenses or sphere) respectively.

3
1=K ¼ ð1  21 Þ=E1 þ ðð1  22 Þ=E2 Þ ð3:30Þ
4

In the above correlations the modulus of elasticity is the material charac-


teristic which is determinant for the deformation of the items in contact.

G ¼ ða3 K=R  pÞ=6a3 K ð3:31Þ

The effective adhesion energy G can be calculated using Eq. (3.31), when the
crack propagates at the interface and gross displacements are purely elastic
[165]. This parameter is similar to the fracture mechanics concept of strain
energy release rate. A desirable feature of a contact mechanical experiment
is the very low crack propagation speed. Another feature is its capability to
reveal both thermodynamic and kinetic information through loading and
unloading cycles.

3.2 Contact Mechanics for Viscoelastic Materials


For viscoelastic materials, from the energetical point of view, contact build-
up and debonding are not fully reversible. The deformation energy is
(partially) dissipated by viscous flow. Such flow depends on the time,
therefore there is a delay between the stress and the strain. In practice,
contacting needs a (dwell) time. Creep appears during bonding/debonding.
Physical Basis for the Viscoelastic Behavior of PSAs 131

Ideally the time dependence of the viscous flow may be a linear


function. Generally it is not. On the other hand (in practice) bonding and
debonding occur at different stress rates, i.e., the time-temperature
dependent material characteristics differ also. Debonding is strongly affected
by the hydrodynamics of liquids and fracture mechanics, manifested
through microscopical deformation of the material under flow (cavitation
and fibrillation, see later), which depend on the experimental conditions too.
Bonding seems to be more easily described by rheological parameters. The
contact mechanics problem for a viscoelastic material including the effects
of adhesion has not been solved.
The basis of viscoelastic adhesion to solid surfaces has been lain by
Creton and Leibler [166]. The viscoelastic nature is accounted for by using a
time dependent modulus E(t) of the adhesive. Hui et al. [167] extended the
model to a rough surface in which the surface topology is modeled by a
Gaussian distribution of spherical asperities. In both cases for wetting,
elastic rather than viscous flow dominated. Elmendorp et al. [134] postulate,
that for PSAs the actual wetting mechanism for slow processes is viscous
wetting rather than viscoelastic deformation.
The most comprehensive theory for the bonding phase is due to
Schapery [168] and Hui et al. [169]. In this theory the rate of contact growth
is related to a reference stress intensity factor, which depends on the loading
history and material viscoelasticity. This theory was confirmed by numerical
simulations [170]. Unfortunately this theory cannot be applied to the
debonding phase, so a comprehensive theory for adhesion hysteresis is still
lacking. Viscoelasticity leads to lower adhesion energy from loading and
higher adhesion energy from unloading than the equilibrum value [171].
Johnson [172] proposed a theory for adhesion hysteresis. This theory
assumes that the rate of loading and unloading are sufficiently slow, so that
the bulk material is in its relaxed state. Barquins and Maugis [173] have
approached this problem from the viewpoint of fracture mechanics for the
special case that viscoelastic losses are localized at the crack tip. Viscoelastic
response is assumed to be proportional to the thermodynamic work of
adhesion, and creep was completely ignored. The theory of Ting [174] and
experimental results suggest creep effects can be significant. Using the
Maxwell and Voigt/Kelvin model, the most striking feature of both models
is that the time of maximum contact radius tpeak does not coincide with the
time of maximum load as it does for elastic materials. Instead maximum
radius always occurs after the maximum load. This delay occurs because the
material continues to respond by creep even during the unloading cycle.
The best sensitivity to creep is obtained if the loading cycle is as short as
possible and the unloading time comparable to the relaxation time. Using
FEM, Lin and Hui [170] showed that Johnson’s solution works well for slow
132 Chapter 3

loading/unloading rates, but fails to capture adhesion hysteresis at moderate


or high rates of loading. In addition Johnson’s theory cannot be used if the
material exhibits long time fluid behavior. Such material behavior is often
used to fit experimental data [175,176]. Lin and Hui [177] investigated
materials with creep compliance of the form:

CðtÞ ¼ Co þ C1 tm 0 < m41 ð3:32Þ

where t is time, C1 is a material constant, and Co is the instantaneous


compliance. The special case of m ¼ 1 corresponds to a linear viscous
material. Using a cohesive zone theory, it was possible to simulate adhesion
hysteresis for two linear viscous spheres in contact. Johnson [172] has shown
that the relevant time scales for the two processes are usually very different.
It has been found [178], that the creep effect should dominate the viscoelastic
response of the most viscoelastic materials with effective moduli smaller
than about 10–100 MPa and W (work of adhesion) < 100 mJ/m2. As
discussed in Chap. 1, Section 2, according to Dahlquist, for PSAs the
compression modulus should not be much higher than 105 Pa (i.e., 0.1 MPa).
That means, that the creep effect dominates the viscoelastic bonding/
debonding behavior of common PSAs.
Although in some cases there is a direct connection between contact
mechanics and peel strength data (see Chap. 6), and the extrapolation of the
contact mechanics data (at low rate) matches the peel data (at high rate), the
above described theoretical approaches (in their actual status) cannot
describe the behavior of PSAs. Supplementary problems arise from the time
dependent character of the peel tests (e.g., adhesive or cohesive debonding),
and in many cases other approaches (than the very frequently used indenter-
based bonding-debonding) give better results, showing that the fundamental
hypothesis concerning the mechanics of the joint is not valid.

3.3 Micromechanical Considerations


As mentioned earlier, in the past few years the parameters of pressure
sensitivity were completed by criteria observed during the microscopic
examination of the adhesive joint. Such characteristics refer to the visible
(physical) modulus of the debonding of strained polymer joints, to the
hydrodynamics of material flow, and to the fracture mechanics. Rheologi-
cally they are related to the modulus of the material.
Adhesion between materials can be considered to be the result of two
contributing factors: the chemical work of adhesion (Wa) which is related to
Physical Basis for the Viscoelastic Behavior of PSAs 133

the bonding across the interface, and the energy dissipation due to the
deformation of the two materials (), generally expressed by:

Gc ¼ Wa ð1 þ Þ ð3:33Þ

where Gc is the strain energy release, a quantitative measure of the adhesion


[179]. The work of adhesion W is composed of contributions from a
thermodynamic component Wo and a viscoelastic component Wo [(R,T)]
[180,181]:

W ¼ Wo þ Wo ½ðR, TÞ ð3:34Þ

The work of adhesion is the energy associated with either formation or


disruption of an interface under equilibrium conditions, which by definition,
allows no dissipation of energy. The work of interfacial fracture is the
energy of disruption of the interface between two solids. Since separation
of two solids cannot be done at equilibrium, some dissipation must occur.
The dissipation is of two kinds: a surface dissipation in the vicinity of the
interface due to the separation, and a second type of dissipation due to the
bulk deformation of the viscoelastic adhesive. The thermodynamic work
of adhesion represents the bond strength (or the intimate wetting of
an adhesive on a substrate), and results from chemical bonds, physical
interactions, acid-base interactions, or other adhesive mechanisms. The
work of adhesion also shows temperature and time dependence, thus
reflecting the viscoelastic properties of the adhesive.
For energy dissipation a special role is played by the deformation,
microscopical flow (strain) behavior of the adhesive polymers. Cracks in soft
materials or at the interface between a soft and a stiff material sometimes
blunt rather than propagate. Delamination of PSAs is accompanied by
cavitation in the PSA, and the formation of an extensive cohesive zone
behind the debond tip. The presence of such large scale bridging may
provide additional energy dissipation and increased resistance to delami-
nation. The transition from propagation to blunting leads to a substantial
improvement in the performance of pressure-sensitive adhesives. According
to [182] blunting in a soft elastic solid has essentially to do with large
deformations. Blunting is predicted when maximum cohesive stress and
Young’s modulus are of similar magnitude.
Taking the example of an indenter which is removed from the
adhesive (e.g., during probe tack, or loop tack), generally the micro-
scopical phenomena of flow related to debonding include cavitation
build-up, build-up of air fingers, and fibrillation (Fig. 3.2). As the probe
134 Chapter 3

Figure 3.2 Debonding by cavitation and fibrillation.

withdraws, the film must sustain a negative hydrostatic pressure and there
is a strong driving force to reduce the degree of confinement of the film.
This reduction of the confinement occurs through the formation of
Saffmann-Taylor air fingers coming from the outside of the probe. These
air fingers cover rapidly the entire contact surface, coalesce, and no
adhesive remains attached to the probe [183]. In practice, cavitation and
fibrillation strongly depend on the geometry of the experimental devices.
When a flat punch is pulled away from the adhesive layer, the contact
area remains constant until cavities form at the punch/adhesive interface.
These voids grow rapidly to an equilibrium size at which point the
loading energy is used to extend an array of fibrils. For flat punches the
hydrostatic stress within the adhesive is large enough to lead to nucleation
and growth of voids. In most commercial PSAs this fibril drawing
process is responsible for the high adhesion energies typically measured.
For spherical punches the most energetically favorable debonding mode
is crack propagation from the contact perimeter.
As the punch is pulled away from the adhesive, mechanical energy is
stored in the adhesive materials as it is deformed. The energy release rate
quantifies the amount of energy which is recovered when an incremental
area of the adhesive detaches from the punch. Part of this energy is stored
as surface energy, in the new surface which is created, but most of it is
Physical Basis for the Viscoelastic Behavior of PSAs 135

dissipated irreversibly within the adhesive in a region near the crack tip.
For a linearly elastic material the values of G are given by:

G ¼ ½ðP0  PÞ2 =4a dC=da ð3:35Þ

where a is the contact radius, P is the applied load, P0 is the load corre-
sponding to a given contact radius in the absence of adhesive interaction,
and C is the compliance of the adhesive layer. As the thickness of the
adhesive layer (h) decreases, its compliance decreases as well, as does its
derivative and the resultant driving force for crack propagation. For very
large values of a/h the compliance of the adhesive layer is inversely
proportional to the bulk modulus of the material.
Experimentally cavitation is observed for values of the average tensile
stress that are comparable to Young’s modulus of the elastic layer
( avg/E ¼ 1). A critical value of G referred to here as Go must be exceeded
in order for a finite crack velocity to be observed. The minimum possible
value of Go is the thermodynamic work of adhesion, which is typically of the
order of 0.05 J/m2 for polymers.
It is to be noted, that such phenomena related to flow instability
possess a general character. They can be observed in the flow of other highly
viscous or viscoelastic fluids also. For instance, the build-up of cavitations
is well known to specialists in the processing of molten plastic films (blown
films, bubble instability); fibrillation is also observed by stretching and
splitting of polymer films (film/filament transition) and from the spinning
of high polymers. In these technologies such phenomena appear if the
strain rate is higher than a critical value (see Chap. 4 also).
Over the last few years appreciable progress has been achieved in the
quantitative description of the micromechanics of PSA debonding [184,185].
These works consider the nucleation of cavities within the PSA polymer and
the extension of fibrils as the major factors leading to the dissipation of
applied energy [186]. For strong adhesives the peel edge propagates as a
process of cavitations and fibrillation [187]. Such failure has been observed in
tack tests [185,188,189] and also in a series of 180 peel tests where the cover
sheet is polyethylene terephthalate and the adhesive is a SIS. Since the peel
edge can be regarded as the tip of an advancing crack, the adhesive ahead of
the advancing peel front is subjected to very large hydrostatic tension due to
the lateral constraint imposed by the substrate and the cover sheet.
The resistance in PSA layers appears to be dominated by the
cavitation behavior of the adhesive. This observation is apparent from
stress-separation tests, where the strongly rate-sensitive peak stress
(cavitation stress,  c) dictates the area under the stress-separation curve.
The cavitation behavior of PSAs can be understood in terms of the
136 Chapter 3

well-established mechanics model [190]. For strong adhesives, the interface


does not debond so that these cavities can grow vertically perpendicular
to the interface. New cavities will form as long as the average spacings
between cavities are large enough to provide the necessary lateral constraint
as pointed out [188]. As more cavities are formed, the spacing between them
is reduced to a level below which the hydrostatic stresses are too low to
further nucleate. There is a lateral hydrostatic stress on the fibrils causing
their spinning, like a tensioned sample under tensile strength, but there is
another vertical hydrostatic pressure caused by compression of the carrier
also. Fibrillation then proceeds by the concurrent vertical growth of these
cavities. The hydrostatic pressure is responsible for the formation of cavities
which grow in the film and progressively become elongated in the tensile
direction to form a fibrillar structure. For an acrylate PSA the cavities
initially appear at the interface between the probe and the film [191]. The
fibrils are modeled as elastic strings. It is assumed that the fibrils break
down at a critical length. According to [65] the strain hardening
observed in the fibrillation part of the probe test curves for styrene block
copolymers can be correlated to tensile tests.
For a material obeying the simple kinetic theory of rubber elasticity,
the critical pressure at which a cavity will grow without limit (tear to form
an internal crack, or fibrillate) has been shown to be proportional to the
initial elastic modulus of the material. For instance (for an SIS-SI
copolymer), if the cavitation stress obtained from stress-separation tests is
plotted as a function of strain rate, the stress exhibits the same trend as the
frequency dependent shear storage modulus (G0 ). These data indicate that
the cavitation stress in PSAs is indeed proportional to the elastic modulus.
Thus, the cavitation criterion of Gent and Lindley [192], according to which
the cavitation in elastic rubbers should occur when   3G0 , is quantitatively
satisfied. However, although energy dissipation through cavitation and
fibrillation is mainly modulus dependent, experimental data demonstrate,
that it depends also on the adherend surface [183,193]. Therefore the
chemical work of adhesion will influence the Gc too, i.e., a pure, mechano-
energetic description of debonding is not possible. As discussed in [194] PIB,
a base polymer for PSAs does not fulfill the Dahlquist criterion and does not
show the fibrillation phenomenon typical for PSAs. According to [69], from
the energetic point of view, cavitation and fibrillation are not equivalent.
On the other hand the debonding force computed from fibrillation
depends on the geometry of fibrillation too [193]:
Z "max
G ¼ ½aðL  eÞ= emax  ð", "0 , TÞd" ð3:36Þ
Physical Basis for the Viscoelastic Behavior of PSAs 137

where "0 is the elongation rate, "max is the maximum elongation of the
filament, l is the periodicity of the filaments, a is the filament width, L is
the filament height, e is the adhesive thickness, and  is the stress given by
Lodge’s model. Thus, one can conclude, that in many PSA systems the
relation between the microscopic structure of a material, its physical
response to large deformations in tack experiment, and its adhesive
properties is not completely elucidated [195,196]. In particular the link
between bond formation and the tack energy, or between the tack energy
and the microscopic processes occurring during separation, such as
cavitation and fibrillation have been the subject of many studies and
still raise a lot of questions.

4 THE ROLE OF OTHER PHYSICAL PARAMETERS


IN CHARACTERIZING PSAs

As discussed earlier the glass transition temperature and the modulus are the
main parameters in characterizing the versatility of PSA raw materials to be
used as viscoelastic compounds having pressure-sensitivity. Other param-
eters may contribute to the investigation of such materials also. Knowledge
from applied physics (mechanics) or special domains of polymer physics
(e.g., compatibility) may help in characterizing PSAs.
The mechanical properties of pressure-sensitive products depend on
the chemical basis and rheological behavior of the macromolecular
compounds. Glass transition temperature, viscoelastic response, and yield
behavior of crosslinked systems are explained by extending the statistical
mechanical theory of physical aging, taking into account the transition of
the WLF dependence to an Arrhenius dependence of the relaxation time
in the vicinity of Tg. As discussed in detail in [197] for classical PSPs
the mechanical properties of the nonadhesive carrier material and of the
adhesive are of different orders of magnitude. The carrier is designed to
resist stresses without deformation; the adhesive is designed to allow a high
degree of deformation without being destroyed. For the main applications,
the plastic carrier material was considered in a first approximation as
nondeformable. In the past decade new pressure-sensitive products were
developed using thin carrier materials (e.g., protective, separation, or
security films) or carrier materials having self-adhesive properties. For such
products the mechanical properties of the carrier strongly influence
pressure-sensitivity. Testing of the mechanical properties is probably the
most perfected area of materials science. Therefore such testing was
extended for the domain of adhesives and adhesive raw materials although it
allows a comparison of order of magnitude only. Although the correlation
138 Chapter 3

between mechanical and adhesive properties is not universal, cohesion


measurements show a good agreement with tensile strength (see Chap. 10
too). For instance, a styrene-isoprene (SI) diblock copolymer possesses a
tensile strength of 20 psi whereas a fully coupled SIS triblock exhibits 300 psi
tensile strength. A shear holding time of 387 days was obtained for a
pure triblock polymer in comparison with 86 days for a commercial one,
illustrating that unentangled and uncrosslinked molecules do not contribute
to the mechanical properties of the polymer [198]. The mechanical
properties were discussed in detail in our book concerning pressure-sensitive
products [197] (i.e., the adhesive-carrier assembly), and some special aspects
are described in Chap. 6. The role of the compatibility of polymers
will be discussed in Chap. 9.

REFERENCES

1. H. Gramberg, Adhäsion, (3) 97 (1966).


2. A. Midgley, Adhes. Age, (9) 17 (1986).
3. J. Kendall, F. Foley and S. G. Chu, Adhes. Age, (9) 26 (1986).
4. F.X. Chancler, J.G. Brodnyan and D. G. Strong, Resin Rev., (3) 12 (1972).
5. Adhäsion, (4)139 (1967).
6. S.E. Krampe and L.C. Moore (Minnesota Mining and Manuf. Co., USA), EP
0202831A2/26.11.86, p. 20.
7. M.C. Bricker and S.T. Gentry, Adhes. Age, (7) 30 (1994).
8. C. Donker, R. Luth, and K. van Rijn, ‘‘Hercules MBG 208 Hydrocarbon
Resin: A New Resin for Hot Melt Pressure-Sensitive Tapes,’’ 19th Munich
Adhesive and Converting Seminar, 1994, Munich, p. 64.
9. K.H. Schumacher and T. Sanborn, ‘‘UV-Curable Acrylic Hot-Melt for
Pressure-Sensitive Adhesives—Raising Hotmelts to a New Level of
Performance,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 165.
10. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 381.
11. A. Zawilinski, Adhes. Age, (9) 29 (1984).
12. M.C. Chang, C.L. Mao and R.R. Vargas, Canad. Pat., 1225792/18.08.1987.
13. J.P. Keally and R.E. Zenk (Minnesota Mining and Manuf. Co., USA), Canad.
Pat. 1224.678/10.07.82 (USPat. 399350).
14. R. Milker and Z. Czech, ‘‘Solvent Based Pressure Sensitive Adhesives,’’ 16th
Munich Adhesive and Finishing Seminar, Munich, 1991, p. 134.
15. W. Hoffmann, Kautschuk, Gummi, Kunststoffe, (9) 777 (1985).
16. N. Willenbacher and A.E. O’Connor, ‘‘Effect of Molecular Weight and
Temperature on the Tack of Model PSAs from Polyisobutene,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 378.
17. T.R. Mecker, ‘‘Low Molecular Weight, Isoprene Based Polymers—Modifiers for
Hot-Melts,’’ TAPPI Hot-Melt Symposium, 1984; in Coating, (11) 310 (1984).
Physical Basis for the Viscoelastic Behavior of PSAs 139

18. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,


Marcel Dekker, New York, 1999, p. 80.
19. Revinex 31F40, Doverstrand, Technical Information, Lat. 051, May, 1987,
Doverstrand Ltd, Harlow, UK.
20. Polysar Lattices in Pressure Sensitive Adhesive Applications, Arnhem, The
Netherlands, 03.1985.
21. G.R. Hamed and C.H. Hsieh, Rubber Chem. Technol., 58, 1038 (1985).
22. Kautschuk Latices, Kunststoff Dispersionen, Latex Chemikalien, Dispercoll
S31, Technische Informationsblätter, Nr.7143, April, 1984, Bayer AG,
Leverkusen.
23. P. Ford, European Adh. Seal, (9) 2 (1997).
24. Adhes. Age, (7) 53 (1994).
25. A. Zosel, Colloid Polymer Sci., 263, 541 (1985).
26. W.M. Stratton, Adhes. Age., (6) 21 (1985).
27. B.W. Foster, ‘‘A New Generation of Polyolefin Based Hot Melt Adhesives,’’
TAPPI Hot-Melt Syposium, June, 7–10, Monterey, CA.
28. T. Wakabayashi and S. Sugii (Minnesota Mining and Manuf. Co., St. Paul,
MN), EP 0285430B1/05.10.1988.
29. G.C. Allen, J.B. Pellon and M.P. Hughes (El Paso Products Co.), EP 251771/
07.01.1988.
30. S.N. Gan, D.R. Burfield and K. Soga, Macromolecules, 18, 2684 (1985).
31. W.C. Perkins, Radiation Curing, (8) 4 (1980).
32. Coating, (1) 12 (1984).
33. Adhes. Age, (9) 37 (1988).
34. M.M. Feldstein, A.E. Chalykh, A.A. Chalykh, G. Fleischer and R.A. Siegel,
Polym. Mater. Sci., Eng., 81, 467 (1999).
35. N. Tanaka, Polymer, 19, 770 (1978).
36. M.M. Feldstein, V.E. Igonin, T.E. Grokhovskaya and N.A. Platé, Proceed.
Intern. Symp. Control., Release Bioactive Mater., 24, 445 (1997).
37. M.J. Brenden, H.A. Schneider and M.J. Cantow, Polymer, (1) 78 (1988).
38. T.G. Fox and J.P. Flory, Bull. Am. Phys. Soc., (1) 123 (1956).
39. D. Satas, Handbook of Pressure Sensitive Technology, Van Nostrand-
Rheinhold Co., New York, 1982.
40. L.M. Schrijvers, Coating, (3) 70 (1991).
41. Adhes. Age, (11) 28 (1983).
42. C.T. Albright, EP 009087B1/23.12.1987.
43. S.D. Tobing and A. Klein, ‘‘Synthesis and Structure Property Studies in
Acrylic Pressure-Sensitive Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001, Williamsburg, VA, p. 131.
44. A. Zosel, Int. J. Adhesion and Adhesive, 18, 265 (1998).
45. H.W.H. Yang, J. Appl. Polymer Sci., 55, 645 (1995).
46. C.D. Han, J. Appl. Polymer Sci., 35 (1) 167 (1990).
47. Coating, (7) 186 (1984).
48. L. Jacob, ‘‘New Developments of Tackifier Resins for SBS Block Copolymers,’’
11th Munich Adhesive and Finishing Seminar, 1986, Munich, p. 87.
140 Chapter 3

49. M.F. Tse and K.O. McElrath, ‘‘Block Copolymer Tackifying Interactions as
Related to Adhesive Performance,’’ European Tape and Label Conference,
19th April, 1989, Brussels, Belgium.
50. G. Bonneau and M. Baumassy, ‘‘New Tackifying Dispersions for Water-
Based PSA for Labels,’’ 19th Munich Adhesive and Finishing Seminar, 1994,
Munich, p. 82.
51. E. Merz, Double Liaison — Chimie des Peintures, (226) 263 (1974).
52. R.P. Mudge (National Starch Chem. Co., USA), EP 0225541/11.12.85.
53. C.M. Chum, M.C. Ling and R.R. Vargas (Avery Int. Co., USA), EP 1225792/
18.08.1987.
54. M. Bowtell, Adhes. Age, (12) 37 (1986).
55. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.1.
56. P.K. Dahl, R. Murphy and G.N. Babu, Org. Coatings, Appl. Polymer Sci.
Proc., (48) 131 (1983).
57. A. Zosel, Adhäsion, (3) 17 (1966).
58. H.W. Yang, ‘‘Effect of Polymer Structural Parameters on PSAs,’’ Proc. of the
22nd Annual Meeting of the Adhesion Society, Panama City Beach, Florida,
Febr., 21–24, 1999, p. 63.
59. G. Prejean, ‘‘High Performance Adhesives from Ethylene Copolymers via
Grafting,’’ 15th Munich Adhesive and Finishing Seminar, 1990, Munich, p. 102.
60. T. Cosgrove, I. Weatherhead, M.J. Turner, R.G. Schmidt, G.V. Gordon
and J.P. Hannington, ‘‘NMR Spin-Spin Relaxation Studies of Reinforced
Polydimethylsiloxane Melts,’’ Polym. Prepr, 39, 545 (1998).
61. G. Holden, E.T. Bishop and N.R. Legge, J. Polymer Sci., (26) 37 (1969).
62. A.S. Rivlin, Paint Technology, (9) 215 (1944).
63. M. Sheriff, R.W. Knibbs and P.G. Langley, J. Appl. Polymer Sci., 17, 3423
(1973).
64. Adhes. Age, (11) 48 (1985).
65. A. Roos and C. Creton, ‘‘Adhesion of PSA Based on Styrenic Block
Copolymers,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 371.
66. M.M. Feldstein, Polymer, 43, 7719 (2001).
67. J. Class and S. G. Chu, Org. Coat. Appl. Polymer Sci. Proc., 48, 126 (1989).
68. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.1.3.
69. G. De Crevoisiere, P. Fabre and L. Leibler, ‘‘Fluorinated Copolymers: an
Example of Temperature Switchable Adhesives,’’ Proc. of the 23th Annual
Meeting of the Adhesion Society’’, Myrtle Beach, South Carolina, Febr., 20–23,
2000, p. 110.
70. L. Shan, C.G. Robertson, N.E. Berghese, E. Burts, J.S. Riffle and T.C. Ward,
‘‘Influence of Vinyl Ester Network Structure on Mechanical Response,’’ Proc.
of the 22nd Annual Meeting of the Adhesion Society, Panama City Beach,
Florida, Febr., 21–24,1999, p. 67.
71. M.I. Aranguren and C.W. Macosko, Macromolecules, 21 (8) 2484 (1988).
Physical Basis for the Viscoelastic Behavior of PSAs 141

72. E.G. Ewing and J.C. Erickson, Tappi J., (6) 158 (1988).
73. D.R. Burfield, Polymer Comm., (6) 178 (1983).
74. S. Trenor, A. Suggs and B. Love, ‘‘An Examination of How Skin Surfactants
Influence a Model PIB PSA for Transdermal Drug Delivery’’, Proc. 24th
Annual Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 144.
75. P. Penczek and B. Kujawa-Penczek, Coating, (6) 232 (1991).
76. H.A. Schneider, J. Thermal Analysis and Calorimetry, 56, 983 (1999).
77. H.A. Schneider, ‘‘Glass Transition,’’ in Polymeric Materials Encyclopedia
(Ed. J.C. Salamone), CRC Press, Boca Raton, 1996, p. 2777.
78. H.A. Schneider and B. Leikauf, Thermochim. Acta, 112, 123 (1987).
79. T.G. Fox, Am. Phys. Soc., (1) 123 (1965).
80. A.J. Kovacs, Fortschr. Hochpolym. -Forsch., 3, 394 (1963).
81. G.R. Hamed and C.H. Hsieh, J. Polymer Physics, 21, 1415 (1983).
82. J.M. Gordon, G.B. Rouse, J.H. Gibbs and W. M. Risen Jr., J. Chem. Phys.,
(66) 4971 (1977).
83. A. Gent and P. Vondracek, J. Appl. Polymer Sci., 27, 4357 (1982).
84. G. Fuller and G.J. Lake, Kautschuk, Gummi, Kunststoffe, (11) 1088 (1987).
85. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.1.1.
86. Ucecryl LRA, August, 1991, 3723/28125, UCB Chemicals, Drogenbos,
Belgium.
87. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter
2.1.1.
88. G. Hombergsmeier, Papier u. Kunststoffverarb., (11) 31 (1985).
89. L.A. Sobieski and T.J. Tangney, Adhes. Age, (12) 23 (1988).
90. Solvay and Cie, Beschichtung von Kunststoffen mit IXAN, WA, Prospect Br
1002d - B - 0,3 - 0979.
91. H.D. Cogan, Off. Digest, (33) 365 (1961).
92. Adhäsion, (5) 76 (1984).
93. C.A. Dahlquist, ‘‘Tack,’’ in Adhesion Fundamentals and Practice, McLaurin
and Sons Ltd., London, 1966.
94. C.A. Dahlquist, in Handbook of Pressure Sensitive Adhesive Technology
(D. Satas, Ed.), Van Nostrand-Rheinhold Co., New York, 1988, p. 82.
95. K. Fukuzawa and T. Uekita, ‘‘New Methods of Evaluation for Pressure-
Sensitive Adhesive,’’ Proc. of the 22nd Annual Meeting of the Adhesion Society,
Panama City Beach, FL, Febr., 21–24, 1999, p. 78.
96. Adhäsion, (5) 100 (1984).
97. M.M. Feldstein, N.A. Platé, A.E. Chalykh and G.W. Cleary, ‘‘General
Approach to the Molecular Design of Hydrophylic Pressure-Sensitive
Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 292.
98. E.J. Chang, Adhesion, 34, 189 (1991).
99. S.G. Chu, in Handbook of Pressure Sensitive Adhesive Technology (Ed.
D. Satas and T.R.I. Warwick), Van Nostrand-Rheinhold, New York, 1989.
100. J.B. Class and S.G. Chu, J. Appl. Polym. Sci., 30, 815 (1985).
142 Chapter 3

101. A. Zosel, J. Adhesion, 30, 135 (1989).


102. D.H. Kaelble, Trans. Soc. Rheol., 15, 275 (1971).
103. D. Satas, Adhes. Age, (8) 28 (1988).
104. D.P. Bamborough and P.H. Dunckley, Adhes. Age, (11) 20 (1990).
105. M.A. Krecenski and J.F. Johnson, Polym. Eng. Sci., 29 (1) 36 (1989).
106. C. Creton, in Processing of Polymers (H.E.H. Meijer, Ed.), VCH, Weinheim,
1997, Chap. 15.
107. J.A. Schlademan and J.G. Bryson, Proceedings of the 21st Annual Meeting
of the Adhesion Society, Savannah, GA, February, 1998.
108. H. Lakrout, C. Creton, D. Ahn and K. Shull, ‘‘Adhesion of Monodisperse
Acrylic Polymer Melts to Solid Surfaces,’’ Proc. of the 23rd Annual Meeting
of the Adhesion Society, Myrtle Beach, South Carolina, Febr., 20–23, 2000, p.
46.
109. H. Wu, Polymer Mater. Sci., Eng. Proceed., (54) 239 (1986).
110. K. McElrath, Coating, (7) 236 (1989).
111. M.M. Feldstein, A.E. Chalykh, A.A. Chalykh, G. Fleischer and R.A Siegel,
Polym. Mater Sci., Eng., 81, 467 (1999).
112. S. Cartasegna, Kautschuk, Gummi, Kunststoffe, (12) 1188 (1986).
113. H. Kehler and W.M. Kulickje, Chem. Eng. Tech., (10) 802 (1986).
114. K. Sato, Rubber Chem. Technol., 56, 942 (1987).
115. A.A. Chalykh, A.E. Chalykh, V. Yu Stepanenko and M.M. Feldstein,
‘‘Viscoelastic Deformations and the Strength of PSA Joints Under Peeling,’’
Proc. of the 23rd Annual Meeting of the Adhesion Society,’’ Myrtle Beach,
South Carolina, Febr., 20–23, 2000, p. 110.
116. U. Eisele, Kautschuk, Gummi, Kunststoffe, (6) 539 (1987).
117. D. Krugers, Kautschuk, Gummi, Kunststoffe, (6) 549 (1988).
118. P.G. De Gennes, Macromolecules, (9) 587 (1976).
119. L.V. Sokolova, O.A. Nikolaeva and V.A. Sersnev, Vysokomol. Soyed., A27,
1297 (1985); in Kautschuk, Gummi, Kunststoffe, (6) 563 (1986).
120. T. Matsumoto, K. Nakmae and J. Chosoake, J. Adhesion Soc. of Japan, (11) 5
(1975).
121. R.A. Fletscher, ‘‘The Temperature Factor in Compounding Thermoplastic
Rubber Based HMA,’’ 11th Munich Adhesive and Finishing Seminar,
Munich, 1986, p. 36.
122. A. Koch and C.L. Gueris, apr, (16) 40 (1986).
123. D. De Jaeger, ‘‘Developments in Styrene Block Copolymers,’’ 11th Munich
Adhesive and Finishing Seminar, 1986, Munich, p. 96.
124. K.E. Johnson and Q. Luvinh, ‘‘Dexco Triblock Copolymers for the Adhesive
Industry,’’ European Tape and Label Conference, Exxon, 1989, Brussels,
Part 19.
125. S.W. Medina and F.W. Distefano, Adhes. Age, (2) 18 (1989).
126. R. Stadler, L. De Lucca Freitas, V. Kriegel and S. Klotz, Polymer, (9) 1643
(1988).
127. J.B. Bonnet, M.J. Wang, E. Papirer and A. Vidal, Kautschuk, Gummi,
Kunststoffe, (6) 510 (1986).
Physical Basis for the Viscoelastic Behavior of PSAs 143

128. G.L. Schneeberger, Adhes. Age, (4) 21 (1974).


129. B. Poltersdorf and D. Schwambach, Kautschuk, Gummi, Kunststoffe, (1)
43 (1988).
130. Adhäsion, (7) 242 (1972).
131. Adhes. Age, (10) 26 (1977).
132. P.R. Drzal and K.R. Shull, ‘‘Adhesive Properties of Model, Filled
Elastomeric Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.,
25–28, 2001, Williamsburg, VA, p. 168.
133. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 3.1.1
134. J.J. Elmendorp, D.J. Anderson and H. De Koning, ‘‘Dynamic Wetting Effects
in Pressure-Sensitive Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001,Williamsburg, VA, p. 267.
135. K.E. van Holde and J.W.Williams, J. Polymer Sci., 11, 243 (1953).
136. P.A. Lovell and T.H. Shah, Polym. Commun, 32, 2 (1998).
137. R. Kamagata and K. Kosaka, J. Appl. Polymer Sci., (15) 183 (1971).
138. K.L. Ullmann and R.P. Sweet, ‘‘Silicone PSAs and Rheological Testing,’’
Proc. of the 22nd Annual Meeting of the Adhesion Society, Panama City Beach,
FL, Febr., 21–24, 1999, p. 410.
139. A. Pocius and C. Dahlquist, Adhesion and Adhesives, ACS Audio Courses
(1986).
140. M. Coleman, J. Graf and P. Painter, Specific Interaction and Miscibility of
Polymer Blends, Technomic Publishing Co., Lancaster, PA, 1991.
141. L. Dorn and G. Moniatis, Adhäsion, (11) 32 (1967).
142. M. Schlimmer, Adhäsion, (4) 8 (1987).
143. A.A. Griffith, Phil. Trans. Roy. Soc., (Lond.), A221, 163 (1920); in D.W. van
Krevelen, Kautschuk, Gummi, Kunststoffe, 37 (4) 295 (1984).
144. L. Jacob, ‘‘HMPSA: a Well Performing, Economic and Environ-
mentally Friendly Technology to Manufacture Tapes,’’ Afera 93, 1993,
Dresden.
145. D.L. Bull, ‘‘Thermoplastic Rubbers, the Utilisation of Their Dual Structure
in Compounding,’’ Shell Chemicals, Thermoplastic Rubbers, Technical
Manual, TR 8-9.
146. G. Auchter, J. Barwich, G. Rehmer and H. Jäger, Adhes. Age, (7) 20 (1994).
147. R. Köhler, Adhäsion, (3) 66 (1972).
148. F. De Candia and R. Russo, J. Thermal Analysis, 30, 1325 (1985).
149. H.Y. Tang and J.E. Mark, Macromolecules, (17) 2616 (1984).
150. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 8.3.
151. Adhes. Age, (12) 35 (1987).
152. R. Kohler, Adhäsion, (3) 90 (1970).
153. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 134.
154. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 127.
144 Chapter 3

155. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,


Marcel Dekker, New York, 1999, p. 180.
156. H.E. Kramer, apr, (31) 843 (1988).
157. Mobil Plastics, Europe, MA 657/06/90.
158. Courtaulds Films, Technical information, 01.04.1991.
159. P. Hammerschmidt, apr, (4) 190 (1986).
160. D. Maugis and H.M. Pollock, Acta Metall, 32,1323 (1984).
161. D. Maugis in Adhesive Bonding (L.K.Lee, Ed.), Plenum Press, New York,
1991, p. 303.
162. K.L. Johnson, K.Kendall and A.D.Roberts, Proc. Roy. Soc., London, Ser. A
324, 301 (1971).
163. K. Kendall and J.C. Padget, Int. J. Adhesion, Adhesives, 2, 149 (1982).
164. J.A. Emerson, J. Benkoski, V. Miller and R.A. Pearson, ‘‘Work of Adhesion
Measurements of Silicone Networks Using Contact Mechanics,’’ Proc. of the
22nd Annual Meeting of the Adhesion Society, Panama City Beach, FL, Febr.,
21–24, 1999, p. 143.
165. K.L. Johnson, Contact Mechanics, Cambridge University Press, Cambridge,
UK, 1985.
166. C. Creton and L. Leibler, J. Polymer Sci. B, Polym. Phys., 34, 545 (1996).
167. C.Y. Hui, Y. Lin and J.M. Baney, J. Polymer Sci. B; Polymer Phys., 38, 1485
(2000).
168. R.A. Schapery, Int. J. Fracture, 25,1 95 (1984).
169. C.Y. Hui, J.M. Baney and E.J. Kramer, Langmuir, 14, 6570 (1995).
170. Y.Y. Lin and C.Y. Hui, submitted to J. Polym. Sci., B, Polym Phys., 2001;
in Y. Lin and C.Y. Hui, ‘‘Detailed Simulations of Viscoelastic Spheres in
Contact,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 153.
171. K.R. Shull, D. Ahn and C.L. Mowery, Langmuir, 13, 1799 (1997).
172. K.L. Johnson, in Microstructure and Microtribology of Polymer Surfaces,
(Eds., V.V. Tsukruk and K.J. Wahl), Chapter 2, American Chemical Society,
Washington D.C., 1999.
173. D. Maugis and M. Barquins, J. Adhesion, 13, 53 (1981).
174. T.C.T. Ting, J. Appl. Mech., 33, 845 (1966).
175. A. Falsafi, P. Deprez, F.S. Bates and M. Tirell, J. Rheol., 41, 1349 (1997).
176. S. Mazur, R. Beckerbauer and J. Buckholz, Langmuir, 13 (16) 4287
(1997).
177. Y. Lin and C.Y. Hui, ‘‘Detailed Simulations of Viscoelastic Spheres in
Contact,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 153.
178. W.N. Unertl, ‘‘Creep Effects in Nanometer Scale Contacts to Linear
Viscoelastic Materials,’’ Proc. of the 22nd Annual Meeting of the Adhesion
Society, Panama City Beach, FL, Febr., 21–24, 1999, p. 17.
179. Brian J. McAdams and R.A. Pearson, ‘‘Adhesion and Deformation at the
Underfill/Polyimide Interface,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001, Williamsburg, VA, p. 104.
Physical Basis for the Viscoelastic Behavior of PSAs 145

180. A.E. Eichstadt, I.V. Farr, J.E. Mc Grath and T.C. Ward, ‘‘The Temperature
and Rate Dependence of the Energy Dissipation in Polyimides,’’ Proc. of the
22nd Annual Meeting of the Adhesion Society, Panama City Beach, FL, Febr.,
21–24, 1999, p. 235.
181. A.N. Gent and J. Schultz, J. Adhesion, (3), 281 (1972).
182. A. Jagota, C.Y. Hui, ‘‘Crack Blunting in a Soft Elastic Material,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 312.
183. G. Josse, O. Poizat and C. Creton, ‘‘Probe Tack Tests of PSAs on Silicone
Release Coatings,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society,’’
Myrtle Beach, South Carolina, Febr., 20–23, 2000, p. 36.
184. A.J. Crosby, K.R. Shull, H. Lakrout and C. Creton, J. Appl. Phys., 88 (5)
2956 (2000).
185. C. Creton and H. Lakrout, J. Polym. Sci., Polym. Phys. Ed., 38, 965 (2000).
186. M.M. Feldstein, T.A. Borodulina, R.Sh. Vartapetian, S.V. Kotomin, V.G.
Kulichikhin, D. Geschke and A.E. Chalykh, ‘‘Correlations Between
Activation Energy for Debonding and that for Self-Diffusion in Pressure-
Sensitive Adhesive Hydrogels,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001, Williamsburg, VA, p. 137.
187. Y.Y. Lin and C.Y. Hui, ‘‘Modeling the Failure of an Adhesive Layer in a
Peel Test,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 230.
188. Y.Y. Lin, C.Y. Hui and H.D. Conway, J. Polymer Sci., B, Polym Physics,
38, 2769 (2000).
189. K.R. Shull and A.J. Crosby, J. Eng. Mater. Technol., 119, 211 (1997).
190. A.N. Gent and C. Wang, J. Mater. Sci., 26, 3392 (1991).
191. J.C. Hooker, C. Creton, P. Tordjemann and K.R. Shull, ‘‘Surface Effects on
the Microscopic Adhesion of Styrene-Isoprene-Styrene-Resin PSAs’’, Proc. of
the 22nd Annual Meeting of the Adhesion Society, Panama City Beach,
Florida, Febr., 21–24, 1999, p. 415.
192. A.E. Gent and P.B. Lindley, Proc. Roy. Soc., A 249,195 (1958).
193. C. Verdier and J.M. Piau, ‘‘Understanding Peeling of PSAs by Use of
Visualization,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society,’’
Myrtle Beach, South Carolina, Febr., 20–23, 2000, p. 116.
194. S. Tobing and S.D. Klein, J. Appl. Polym. Sci., 79, 2230 (2001).
195. A. Zosel, J. Adh. Sci. Tech, 11, 1447 (1997).
196. H. Lakrout, C. Creton, D. Ahn and K.R. Schull, Polymer Mater. Sci., Eng.,
81, 458 (1999).
197. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chap. 3.2.1.
198. M.H. Mazurek (Minnesota Mining and Manuf. Co., St. Paul, MN), EP
4693935/March 15, 1987.
4
Comparison of PSAs

There is no special science dedicated to the study of PSAs and there


is no special education geared to this particular field. Therefore specialists
from other related fields work on the complex technical problems involved
in the manufacture and use of PSA laminates. Pressure-sensitive adhesives
are macromolecular compounds and thus one can suppose that a better
understanding of this field requires the input of a polymer scientist or
engineer. In fact certain mathematical correlations or approximations
describing the dependence of the polymer behavior on the experimental
nature and structure are also valid for PSAs. But the problem appears more
complex because of the viscoelastic behavior of PSAs which depends on
the experimental conditions; PSAs may obey correlations valid for plastics
(viscous materials) or rubber (elastic materials).
Doing mostly empirical work, the PSA expert tries to act on a
theoretical basis. This theory is based on the knowledge from processing
plastics and elastomers, or from the more thoroughly investigated field
(relative to PSAs) of general purpose adhesives. Therefore, a better under-
standing of PSA practice requires a quintessence of the similarities/
differences in the properties/rheology of plastomers/elastomers, adhesives,
etc. Generally, making a comparison between the rheological properties
of plastomers/elastomers, PSAs, and of classical adhesives, one can observe
certain similarities and differences. Similarities are given by the (limited)
flow of plastomers/elastomers; the differences arise from the (partially)
ordered structure of these materials.

1 COMPARISON OF PSAs WITH THERMOPLASTICS


AND RUBBER

Similar to plastics and rubber, PSAs are macromolecular compounds; their


properties depend on their chemical composition and macromolecular
147
148 Chapter 4

characteristics. But in a different manner from elastomers, plastomers, or


common adhesives, the physical state and not the chemical composition
plays a determinant role for PSAs. For plastics and rubber the fluid state
is only required for processing purposes. For PSAs flow is a permanent
need. Properties related to flow are the lack of crystallinity, the lack of
orientation, the need for cold flow, the high rate of relaxation, and the
ductility. It should be pointed out that in comparison to the flow of plastics
or rubber during processing, the flow of PSAs occurs isothermally. On the
other hand the PSA fluid is always a composite, containing solids, liquids,
or volatile components coming from the converting process (fillers, solvents,
wetting agents, etc.). The level of additives is never zero, but remains
relatively low (less than 20–30% by weight).
Generally the rheological similarity of PSAs, thermoplastics, and
rubber is observed by examining the practical behavior (e.g., cuttability,
re-adhering behavior, mechanical destruction, etc.) of the material. These
properties, starting from similar practical processing or test conditions
for polymers and from the parameters influencing the polymer behavior are
discussed next.

1.1 Cold Flow


Pressure-sensitive adhesives must exhibit a good resistance to stringing,
edge flow, and bleeding [1]. Even though PSAs are highly viscous, they
nevertheless display cold flow [2]. The rheology of the adhesive determines
its converting and end-use properties. In the industrial practice of laminate-
converting, the storage (as a converting-related phase) and the cuttability
of the laminate are affected by the flow of the adhesive (see Chap. 2, Section
3.1). As far as the end-use properties are concerned the plasticity/elasticity
balance is important. Both storage and cutting may be accompanied by edge
flow (oozing) and bleeding of the adhesive. This phenomenon is due to the
viscous flow of PSAs and occurs also at low (room) temperature.
The cold flow is a general characteristic of plastics. In order to
understand its importance and the parameters influencing it, the flow of
plastics and the similarities and differences between the cold flow of plastics
and that of PSAs need to be examined. The cold flow of plastics is generally
taken into account in the design of plastic elements (pieces) and where
safety factors are required; when long-time, dynamical, or thermal stresses
are expected, lower values of the relevant material properties are used.
The nature of the flow behavior does not change for amorphous plastics
at higher temperatures, therefore their molten state and flow in the
molten state may be taken as a simplified model for the flow of PSAs.
The investigation of the parameters influencing the flow of plastics may yield
Comparison of PSAs 149

information about the flow of PSAs during processing. The influence of


the coating weight, shear rate, and the viscosity on the flow of PSAs will
be reviewed based on the processing of plastics.

Coating Weight and Cold Flow


There are tests in polymer processing to determine the calendering ability
of a polymer; these are based on the shear resistance of the material [3]. Here
the maximal shear rate _ max depends on the linear speed of the cylinder ,
and the thickness of the layer ho:

_max ¼ 2v  h1
o ð4:1Þ

that is

v ¼ _max  ho =2 ð4:2Þ

Assuming a similarity between the material volume squeezed out


during cutting of pressure-sensitive laminates and calendering of plastics
(push-out of the plastic web) one can estimate that the rate of edge bleeding
(i.e., the volume of the adhesive oozing out from the laminate) depends on
the adhesive’s thickness (i.e., on its coating weight) [4], or:

Bleeding ¼ f ðcoating weightÞ ð4:3Þ

For polymers (plastics) the minimum pressure for flow can be esti-
mated from the knowledge of the melt flow index (MFI); the MFI depends
on the temperature and this dependence is a function of the temperature
range used. With knowledge of the geometric parameters of the mold,
temperature, and thermal properties of the melt, the minimum pressure
requirement for flow may be calculated. On the other hand the thermal
properties are polymer dependent (see Chap. 2, Section 1.5) [5]. For rubber
blends there is a flow (creep) limit, tested with a Höppler consistometer [6],
depending on sample volume V, modulus G, and sample thickness hL,
according to the following equation:

 1=2
To ¼ 9h5L G2 =4V 3 ð4:4Þ
150 Chapter 4

Viscosity and Cold Flow


During the extrusion of plastics the volume V of the extruded elastomer
depends on its viscosity s [7]:

V ¼ ½i  ðBH=2Þ  fs  v2   ½i  ðBH 3 =12s Þ  fs  ðdp=dzÞ ð4:5Þ

where i, BH, v, fs, and p are polymer- and machine-dependent parameters.


Assuming a similarity between the material squeezed out by extrusion
and by the cutting of PSAs laminates, one can conclude that:

Oozing ¼ f ðViscosityÞ ð4:6Þ

Thus, on the basis of Eqs. (4.1)–(4.5) describing the flow of plastomers


during processing, one can suppose a similar dependence of the flow of
PSAs on the viscosity, shear rate, and layer thickness; cuttability also
depends on these parameters (see Chap. 2).
Other complex features of the polymer flow also show similarities with
PSAs. An example is given by the melt flow viscosity of polymer blends.
Here an ‘‘interfacial slip viscosity’’ is defined because of the observed
minima in the viscosity. A multilayer structure is assumed for blends where
the steady state shear viscosity of the blend at a constant shear stress may be
written as follows [8]:

1
i ¼ 1 =1 þ 2 2 þ i i ð4:7Þ

where i are the viscosities of the adjacent layers and i are the volume
fractions. Measurements of i were lower by a factor of 10 with respect to
the lowest viscosity of the mixed homopolymers. The minimum in viscosity-
composition plots of the blends disappeared on the addition of selected
block copolymers to the blend. Thus it may be concluded that interfaces
with low frictional resistance may be formed and must be considered
when modeling the behavior of polymer blends. In tackified PSAs similar
viscosity-lowering phenomena occur, which may be partially avoided using
polymeric additives.

Shear Rate as a Parameter for Cold Flow


Some rubber blends that exhibit good processability during pressing do not
display good properties during injection molding because of the different
shear rates applied in both processes. Since the viscosity of these rubber
Comparison of PSAs 151

blends is a function of the shear rate, the tests of processability have to be


carried out at different shear rates according to the real shear conditions.
Molten thermoplastics undergo high shear straining during process-
ing; extrusion of polymer melts is carried out at shear rate values of 102
to 105 s1 [9]. At such a high rate, high shear processing may increase the
polymer temperature and may decrease the polymer viscosity. Experimental
work has developed a mathematical basis for this phenomenon. The
amount of mechanical energy transformed into frictional heat (also called
dissipation D) depends on the viscosity and shear rate [10].

D ¼ f ð  _ 2 Þ ð4:8Þ

where  and _ denote viscosity and shear rate, respectively. The increase of
the temperature of a sheared polymer sample depends on the shear stress  ,
the shear rate _ , the density , the heat capacity Cp, and the shear time t.
In a similar manner it may be assumed that the cutting of pressure-sensitive
laminates and the shear during cutting depend on the same parameters,
that is, that the rate of the cutting and cutting time influence the bleeding
and smearing of the adhesive during cutting, because (at least partially) of the
increase of the temperature. Thus the equation given for the temperature
increase of plastics will be valid for PSAs:

 ¼ ð  _  tÞ=ð  Cp Þ ð4:9Þ

where , _ , Cp, and t are the above-mentioned parameters. In the processing


of plastics the decrease of the viscosity at higher shear rates and the self-
heating of the material are positive phenomena; for PSA-converting
purposes, however, these phenomena constitute a disadvantage [11].

Flow Test of Plastics and PSAs


The shear rate must be examined by the test of the flow properties of the
plastomers. The amount of the material yield V in a plastometer is given by
Eq. (4.10), where the shear rate _ plays an important role [12]:

ð  R3  _  TÞ
V¼ ð4:10Þ
4

where R and T are the universal gas constant and the temperature,
respectively. Unfortunately in laboratory scale tests the melt flow of
thermoplastic polymers is characterized by the melt flow rate (MFR), which
depends on the molecular weight and molecular weight distribution
152 Chapter 4

measured under low shear conditions [13]. Thus no (or only limited)
information about the real molding behavior may be obtained from the melt
flow rate. Extrapolated to the field of PSAs, the MFR corresponds to a
static, long cold flow time. It is interesting to note that like the MFR the
shear of PSAs depends on the molecular weight. This statement is valid
and important for tackified PSAs, where shear resistance may decrease
exponentially with the level of low molecular weight tackifier resin. As with
MFR values, static or low speed dynamical shear tests provide only limited
information due to the high speed shear resistance (i.e., the cuttability of
PSAs). In fact, the melt flow of polymers is both pressure (shear) and
temperature dependent [14]. Therefore the same temperature dependence for
holding power tests, or other low stress rate tests (e.g., Williams-plasticity)
may be assumed for PSAs. A similar temperature dependence of the flow
characteristics may be observed in the evaluation of elastomers (e.g., hot-set
test: tensile strength at high temperature and less than 75% elongation) [15].
Practically, there is a relation between the time/temperature dependence of
the tested polymer characteristics and the processing behavior of the
material. As an example Shenoy [15] demonstrated the dependence of the
minimum pressure P necessary to fill a cavity during processing, on the
polymer characteristics [i.e., f (n), K, C, and n] and the melt flow index
(MFI):
 
Pmin n
¼ f ðÞ K  ð4:11Þ
C 3n R1þn MFIn

where  and K are rheological parameters; for the dependence of MFI on


the temperature, the Arrhenius or Shenoy equation [Eq. (4.11)] should be
used. In a similar manner the dependence of the PSA bleeding (or migration,
especially for hot-melt PSAs) on the polymer rheological characteristics and
temperature should be evaluated.
Generally there are many test methods available to investigate the
temperature dependence of certain properties of plastics and elastomers
(e.g., dependence of the tensile modulus on temperature [16], or of the
critical tear energy [17]), which should be considered for similar measure-
ments on PSAs. There are also similarities in the low temperature range
and the low shear rate tests of plastics/PSAs. In the case of polymer
deformation by elongation (tensile) the viscosity depends on the tensile rate
(draw ratio) [18]:

ðtÞ
c ðtÞ ¼ ¼ tensile strain=elongation rate ð4:12Þ
_
Comparison of PSAs 153

Table 4.1 Test of the Adhesive Flow via Tensile Strength and Static Shear

Values for Different Formulations

Performance Characteristics 1 2 3 4 5

Shear resistance (min) 900 200 50 170 30


Tensile strength (N/mm2) 8.0 3.3 2.9 3.5 1.4
Elongation (%) 400 1020 1150 700 1550

The strain rate is low (below 1 s!) and it enables the slow mechanical
response of the material to be examined [19]. Therefore it may be assumed
that tensile tests occur at similar low shear rates to static shear
measurements (holding power). Such tests would not give any relevant
data about high shear behavior (cutting) of PSAs. Table 4.1 illustrates the
flow (cohesion) behavior of a PSA tested by tensile strength and statical
shear (holding power). As can be seen from Table 4.1, the tensile strength
of the adhesive is proportional to its shear strength but only for certain
formulations.

Rheology of Plastics/PSAs
There is a special chemical basis of PSAs that provides the whole assembly
of rheological properties (see Chap. 2). The main polymer characteristics
(e.g., modulus and viscosity) may be controlled through chemistry. The
basic knowledge of the interdependence of polymer composition/structure
and rheological parameters is known for plastomers/elastomers. Adhesive
chemists have to apply these general methods to the special field of PSAs.
The most important molecular factors influencing the rheological
behavior of polymer melts are the chemical composition of the monomers,
the average molecular weight, the MWD, and the branching of the chains
[20]. These same parameters also influence the rheology and the adhesion-
cohesion balance of PSAs. Other more complex similarities also exist; the Tg
is an example. The chemical composition of the monomer unit provides the
flexibility and fluidity of the adhesive. Because of the volume dependence
of the Tg, substituted high-volume resins provide less tack. The chemical
composition of the monomer makes crosslinking possible. The crosslinking
density influences tensile strength and tension set [21]; in a similar manner
it influences the peel and cohesion of PSAs. In certain tackified rubbers, gel
formed by crosslinking gives higher tack, but generally crosslinking leads
to lower tack (nontackified mixtures). An increase in molecular weight of
plastics may lead to incompatibility in plastic blends. In a similar manner
154 Chapter 4

tackifier resins with a molecular weight higher than 1000 cause incompat-
ibility problems. On the other hand this behavior may explain the higher
tackifying effect of low molecular weight liquid resins.
As discussed earlier (Chap. 3, Section 3), trials were made to correlate
the pressure-sensitive debonding with the visible morphology of the
debonded surfaces, i.e., with the build-up of surface and bulky defects in
the adhesive and the development of fibrillar structures. The fracture of such
structures was compared with the break of macroscopic, plastic (elastic or
viscoelastic) items, using the well-known test and calculation methods
of classical mechanics. The mechanical performances can be related to the
modulus of the material, and to its macromolecular characteristics. Such
fibrillar structures are known from different technologies working with fluid-
like materials (polymer solutions, molten plastics, plastics under creep, etc.).
The ‘‘legging’’ of PSAs was described also. According to [22,23] the spinning
of polymeric fibers, the processing of numerous foodstuffs, and the peel
and tack characteristics of adhesives are all associated with the formation,
stability, and ultimately the longevity of the fluid ‘‘strands.’’ This tendency
to form strands is usually described in terms of tackiness of the fluid or
by concepts such as stringiness. The dynamics of such processes are com-
plicated due to spatially and temporally nonhomogeneous growth of
extensional stresses, the action of capillary forces, and the evaporation of
volatile solvents. Fibrillation as an engineering phenomenon is the result
of material and processing characteristics. Theoretically it can be described
by dimensional analysis. Although such analysis was yet not developed,
trials were made to describe fibrillation as a general phenomenon (including
PSAs also). Tripathi et al. [22] discuss the application of a simple instrument
referred to as a capillary break-up extensional rheometer [24] that can be
used to readily differentiate between the dynamical response of different
PSAs. The device relies on the quantitative observation of the rate of
extensional thinning or ‘‘necking’’ of a thin viscous or viscoelastic fluid
filament. The high resolution measurement of the radial profile provides
a direct indication of the ultimate time to break-up, of the fluid filament.
This critical time is a sensitive function of the rheological properties of
the fluid and the mass transfer characteristics of the solvent, and can be
conveniently reported in terms of a dimensionless processability parameter
(P). Experiments with a new variant of the filament rheometer using
a viscoelastic polymer solution with a nonvolatile solvent showed a good
agreement with numerical simulations of the elasto-capillary thinning
process. The computations indicate that PSA samples with a processability
number of P 3  104 show rapid necking and break-up corresponding
to good processability, whereas PSA samples with P 3  103 are prone
to strand formation and have poor processing characteristics.
Comparison of PSAs 155

As discussed in detail in [25,26] there are special adhesiveless self-


adhesive products (e.g., protective films, separation films, etc.) which exhibit
auto-adhesion due to their chemical composition and to the application
conditions. Such self-adhesive carrier materials are plastic films having
plastic or plasto-elastic character. As described in [27] their self-adhesive
performances are given by improved molecular freedom to freely rotate.
Generally their Tg is higher than the glass transition temperature of
common PSAs but lower than their application temperature. Such materials
are used mainly as removable products, where removability is achieved by
the low level of adhesion and by the energy dissipation during deformation
of the thin polymer film. Developments in film-forming polymers allowed
the increase of the conformability of such films (i.e., better bonding
characteristics), thus self-adhesive carrier materials compete with adhesive
coated ones in application domains requiring higher peel adhesion values.
Actually there is no theoretical approach comprising self-adhesivity
achieved by means of different technologies (e.g., adhesive coated and
adhesiveless products).

1.2 Relaxation Phenomena


For all elastomers, reinforced or not, the stress to be applied to obtain
a given elongation is less important on a second elongation cycle [28].
This so-called stress-softening or Mullins effect [29] results from several
mechanisms (i.e., not only the structure of the elastomeric network, but
also the interactions of the polymer with the solid components of the
laminate). The stress-softening effect should be taken into account for
re-adhering adhesives or for the repeated (periodical) compression-yield-
shear stresses when cutting pressure-sensitive laminates. The dependence of
the peel force and rate for rubbery materials on the frequency of subsequent
bonding and debonding (e.g., reuse, renewed application of the adhesive
label), has also been described as a ‘‘memory effect.’’ An applied label loses
this memory effect after some time or when in a solvent. The memory as a
function of the dwell time effect is associated with some rearrangements
of the molecular structure of the rubber at the interface. This memory effect
is important when re-adhering repositionable labels.
The role of the relaxation phenomena and their influence on the
mechanical properties of plastics is well known. The dependence of
the mechanical properties of plastic blends on the relaxation ability was
demonstrated. Applying cyclical stresses to polymer/polymer mixtures
leads to a dissipation of the energy in the elastic parts of the mixture; the
tensile strength depends on the viscoelasticity and relaxational behavior.
Similar phenomena exist for PSAs where a balance of elastic/viscous parts
156 Chapter 4

(given by a certain formulation) may allow a soft or hard debonding


(see Chap. 2, Section 1.4) as a result of the relaxation. As discussed earlier
(Chap. 3, Section 3) fibrillation supposes the existence of a minimum
molecular weight (higher than Me). The cohesive strength of the fibrils/
lamellae increases with increasing chain relaxation time. The area of contact
of an elastic film on a rough surface depends on the relaxation properties
of the polymer also.

1.3 Mechanical Resistance


The mechanical resistance of plastics, rubber, and adhesives is quite
different. As can be seen from the modulus values (see Chap. 2), the
modulus of the face stock materials (paper or plastics) differs by a factor
of 103 from the modulus of rubber, or by 104 from the modulus of the
adhesive. Rubber is an ideal raw material for PSAs. Rubber displays
auto-adhesion and shows strain-induced crystallization [30]. Amorphous
rubber is very mobile during application (bonding) and strain-crystallized
materials yield high shear. But PSAs are mostly mixtures of resin and rubber
in a 1:1 ratio, thus it is as yet not clear if the effects of the strain-induced
crystallization can be extrapolated to PSAs. A more thorough examination
of the main rheological and mechanical characteristics of plastics/elastomers
and PSAs reveals a lot of differences in their properties. Generally the values
of the modulus E, the tensile strength, and the elasticity are much lower for
PSAs than those of the elastomers (or plastomers) [25]. High modulus values
as encountered with elastomeric block copolymers (e.g., 250–350 kg/cm2),
high elastic recovery values (of about 80%), or high elongation at break
(800–1300%) cannot be obtained for PSAs. Their flow characteristics
are very pronounced. There is a model rubber (DIN 53516) that is used to
characterize rubber and it is possible to characterize the creep under com-
pression of the elastomers in a ‘‘creep unit.’’ If the deformation  < 0.8%
(see Chap. 3, Section 2.3), then for plastics one can write:

E 6¼ f ðÞ ð4:13Þ

If the modulus does not depend on the stress value, the yield, compression,
and flexural modulus are identical,

Et ¼ Ec ¼ Ef

No model PSA exists as deformation values for PSAs are higher than
0.8% and the modulus values for different kinds of stresses vary. On the
Comparison of PSAs 157

other hand, not only the deformation value but also the mechanism of the
deformation and destruction of plastics/elastomers differ from those of
PSAs.
Plastics are viscoelastic materials and do not obey linear elastic
fracture mechanics (LEFM) unless as an approximation where the plastic
zone remains small [20]. For ductile fracture the F integral was suggested
as a fracture criterion for large-scale plasticity. In a similar manner PSAs
undergo ductile fracture; however, in the design of removable PSAs the
formula for the critical fracture toughness Kc given by Griffith [31] should
also be considered:

Kc ¼ c  a1=2  ð4:14Þ

where depends on the geometry of the specimen, a denotes the crack


length, and  c is the critical value of the remotely applied stress at which
the crack begins to grow.
It may be assumed that a similar correlation exists between a specimen
geometry a, ‘‘leg’’ length qa, a critical value of the applied stress  ca, and
critical tear toughness Kca for a removable PSA [32]:

Kca ¼ f ðca  qa  aÞ ð4:15Þ

The test and calculation methods used for the mechanical perfor-
mances of industrial plastics were extrapolated for PSAs also. In some cases
the common mechanical characteristics, e.g., tensile strength, tear strength
etc., are related to the adhesive performances. For instance, by the analysis
of PSA fracture mechanics in the course of peeling, Feldstein et al. [33]
established that the peel force P relates to tensile strain  and stress  by
the following equation:

Z
l
p¼b ðÞ  d ð4:16Þ
2
0

where b is the width and l is the thickness of the adhesive film. At low tensile
rates the adhesive fracture occurs at comparatively small tensile stress
values, but at significantly greater values of elongation, . With the increase
in extension rate, the critical magnitude of tensile stress at fracture ( f)
varies rapidly, approaching a limiting value which corresponds to a
maximum cohesive strength of the PSA film  c. A similar behavior was
found in [34]. As stated, the peel strength is strongly dependent on the
158 Chapter 4

effective rate of peeling. At low rates the elastomer layer could not be
detached cleanly. Instead it split apart leaving rubber behind (cohesive
failure). Under these circumstances the peel strength is the same as the tear
strength measured by tearing apart a thin sheet of the elastomer. A detailed
discussion of the test methods will be given in Chap. 10.

2 COMPARISON BETWEEN PSAs AND


OTHER ADHESIVES

Common adhesives have to be fluid only during the coating process. Similar
to PSAs, the flow of common adhesives during application is obtained
through the aid of a dispersing agent (solvent or dispersing media) or by
melting the bulk adhesive. After coating the adhesive bond is achieved by
the increased viscosity of the adhesive (in comparison with the application/
viscosity) and, in some cases, by the chemical interaction of the adhesive
with the adherend. In special cases after bond forming, the adhesive itself
undergoes a chemical transformation (e.g., crosslinking) that results in a
higher adhesion/cohesion balance with respect to the uncoated adhesive.
This is possible by changing the viscous/elastic balance and increasing
the value of the modulus. In this case the adhesive layer has similar rheo-
logical and mechanical properties to those of the adherend, and therefore
debonding (without material destruction) or rebonding is no longer possible.
The properties of such adhesive joints are similar to plastics or elastomers.
Similarities with the rheology of PSAs are observed for the uncoated, or
coated but unbonded adhesive only. On the other hand, the chemical basis
of the common adhesives is almost the same as that of PSAs. The chemical
composition of PSAs and common adhesives may be the same, although
the role of the built-in fractional groups may be different. The presence of
itaconic and/or acrylic acid in vinyl acetate (VAc) terpolymers imparts
excellent freeze-thaw stability to these emulsions. The same monomers in
PSAs are used to crosslink and to impart special adhesion. The same
statement is valid for the formulation. Plasticizers or tackifiers used in PSAs
in order to achieve tack and peel give polyvinyl acetate (PVAc) adhesives
better bonding ‘‘speed.’’
Classical adhesives and PSAs differ in the reason why one uses
plasticizers and tackifiers. For classical adhesives plasticizers and tackifiers
modify the application viscosity and green tack, whereas for PSAs
they modify tack and peel. The functional groups in classical adhesives
serve mainly as crosslinking sites, while in PSAs they are adhesion
promoters. Thus in hot-seal adhesives the resin improves the adhesion to
difficult substrates, decreases the blocking resistance, and lowers the seal
Comparison of PSAs 159

temperature [35]. Differences in the use of additives for water-based


formulations may be noted also. Thus polymeric colloids for use in PSAs
have a relatively low Tg (i.e., in the range of about 40 to 20 C), while
the colloids for use in laminating adhesives have a Tg value within a range
of 25 to 10 C.
Classical (non-pressure-sensitive) adhesive joints are designed and tested
according to the theory and calculation methods of common mechanics.
Such methods were used for PSAs also. For instance, Brockmann and Geiß
[36] extrapolate the calculation methods used for common adhesive joints to
PSAs, admitting that joint failure occurs after a critical value of the creep.
For the evaluation of this critical load limit, creep measurements have been
carried out using single lapped shear joints. Three stadiums of creep were
observed. It starts with an elastic strain and decelerated creep. When the
acceleration becomes zero, i.e., at the beginning of constant creep velocity,
the deformation turns from viscoelastic-plastic to viscoelastic. The third
creep stadium is defined by a starting acceleration and ends with total failure
of the specimen.

REFERENCES

1. Adhes. Age, (3) 8 (1987).


2. H.G. Koch, Adhäsion, (11) 312 (1976).
3. J. Guillet and V. Verney, Polymer J., (8) 773 (1984).
4. R.R. Lowman, FINAT News, (3) 24 (1987).
5. D.R. Saini and A.V. Shenoy, Plastics and Rubber Proc. Appl., (3) 178 (1983).
6. B. Poltersdorf and D. Schwambach, Kautschuk, Gummi, Kunststoffe, (5) 454
(1988).
7. H.J. Laake, Kautschuk, Gummi, Kunststoffe, (6) 506 (1989).
8. J.L. Jorgensen, L.D. Thomson, K. Rasmussen and K. Soendergard, Internat.
Polymer Process, (3–4) 122 (1988).
9. F. Kurihara and S. Kimura, Polymer J., 17, 863 (1985); in Kautschuk, Gummi,
Kunststoffe, (3) 256 (1985).
10. H.Y. Tang and J.E. Mark, Macromolecules, (17) 2616 (1984).
11. H. Kehler and W.M. Kulickje, Chem. Ing. Techn., (10) 802 (1986).
12. Kautschuk, Gummi, Kunststoffe, (9) 31 (1985).
13. S. Cartasegna, Kautschuk, Gummi, Kunststoffe, (12) 1188 (1986).
14. F. Grajewski, A. Limper and G. Schwarter, Kautschuk, Gummi, Kunststoffe,
(12) 1188 (1986).
15. G. Bolder and H. Meier, Kautschuk, Gummi, Kunststoffe, (8) 715 (1986).
16. Y. K. Kawasaki and N. Watanabe, Kautschuk, Gummi, Kunststoffe, (2) 119
(1987).
17. T. Gehman, Kautschuk, Gummi, Kunststoffe, (6) 493 (1989).
18. T.S. Ng, Kautschuk, Gummi, Kunststoffe, (9) 830 (1986).
160 Chapter 4

19. J.N. Hay, Coaters Scientific Conference Communications, Surface Coating


Evaluation and Performance, 1988.
20. H.M. Laun, Kautschuk, Gummi, Kunststoffe, (6) 554 (1987).
21. G. Bolder and H. Meier, Kautschuk, Gummi, Kunststoffe, (3) 196 (1984).
22. A. Tripathi, G.H. McKinley, and P. Whittingstall, ‘‘Using Filament Stretching
Rheometry to Predict Strand Formation and Processability of Pressure-
Sensitive Adhesives and Other Complex Fluids,’’ Proceedings of the 23rd
Annual Meeting of the Adhesion Society, Myrtle Beach, SC, Febr., 20–23
2000, p.119.
23. Cr.I. Simionescu, N. Asandei and I. Benedek, Rev. Roum. Chim., (7) 1081
(1971).
24. M.I. Kolte and P. Szabo, J. Rheol, 43 (3) 609 (1999).
25. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 3.
26. I. Benedek, E. Frank and G. Nicolaus (Poli-Film Verwaltungs GmbH,
Wipperfürth, Germany), DE 4433626 A1/21.09.1994.
27. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.
28. J.B. Bonnet, M.J. Wang, E. Papirer and A. Vidal, Kautschuk, Gummi,
Kunststoffe, (6) 510 (1986).
29. L. Mullins, J. Rubber Res., (16) 275 (1947).
30. G. Hammed, Rubber Chem. Technol., (5) 576 (1981).
31. A. Griffith, in Coaters Scientific Conference Communications, Surface Coating
Evaluation and Performance (J.N. Hay, Ed.), 1988.
32. R. Dorpelkus, Allgemeine Papier-Rundschau, (16) 456 (1986).
33. A.A. Chalykh, A.E. Chalykh,V. Yu Stepanenko and M.M. Feldstein,
‘‘Viscoelastic Deformations and the Strength of PSA Joints Under Peeling,’’
Proceedings of the 23rd Annual Meeting of the Adhesion Society, Myrtle Beach,
SC, Febr., 20–23, 2000, p. 252.
34. A.N. Gent, G.R. Hamed and W.J. Hung, ‘‘Adhesion of Elastomers: Dwell
Time Effects and Adhesion to Fillers,’’ Proceedings of the 23rd Annual Meeting
of the Adhesion Society, Myrtle Beach, SC, Febr., 20–23, 2000, p. 6.
35. E.A. Theiling, Kunstharznachrichten, (1) 14 (1972).
36. W. Brockmann and P.L. Geiß, J. Adhesion, 63, 253 (1997).
5
Chemical Composition of PSAs

A typical PSA is derived from a film-forming elastomeric material such as


natural rubber, styrene-butadiene, butyl, silicone, nitrile, and acrylic rubber.
The elastomer provides flexibility, ease of tackification, and the desired
bond strength when compounded with compatible tackifiers, pigments,
plasticizers, waxes, and oils. Special polymers were developed as raw
material for PSAs which work with or without tackifiers. Natural rubber
represents 45% of the elastomeric raw materials; block copolymers and
water-based acrylics amount to 40%, and the remaining part is mainly
covered by solvent-based acrylics. Although new elastomers have been
identified as base materials for PSA applications, natural rubber is still the
most preferred polymer.
From a chemical point of view classical PSAs are elastomers that
exhibit viscoelastic properties due to their low Tg (40 to 60 C) and
modulus. Their macromolecular basis is built on long-chain polymers, with
a certain degree of branching and with or without crosslinking. The balance
between plastic and elastomeric properties is governed by the polymer
nature and structure, and molecular weight. For classical rubber-based
PSAs, structure and molecular weight adjustment occurs via crosslinking or
mechano/chemical destruction of the network (e.g., mastication, tackifica-
tion, plasticizing). Synthetic raw materials for PSAs can be tailored in order
to achieve the chemical characteristics required to obtain the desired
adhesive properties. A detailed study of the interdependence between the
chemical basis and physical structure of PSAs was carried out by Krecenski,
Johnson, and Temin [1]. As discussed earlier (Chap. 2) the rheology of
the pressure-sensitive product depends on its solid components and the
interaction between the liquid and solid components too. The chemical
composition of the whole pressure-sensitive construction—polymer carrier
and macromolecular adhesive—was discussed in [2]. A detailed analysis of
the formulation of PSAs is given in [3].

161
162 Chapter 5

Pressure-sensitive adhesives are amorphous, viscoelastic materials, and


their flow behavior is characterized by the viscosity and the modulus (see
Chap. 4). On the other hand a special structure and temperature dependence
of the internal phase transitions serves as the basis for the required rheology.
Since the viscosity, modulus, and glass transition temperature are dependent
on the chemical composition, changes in the composition and/or in the
macromolecular characteristics of the basic elastomers may influence the
rheology of PSAs. From a theoretical point of view it is more accurate to
discuss the influence of the compositional and macromolecular factors on
the properties of PSAs separately. However, it is not within the scope of this
book to present a detailed, purely theoretical approach to PSAs; therefore
the chemical basis (composition and macromolecular characteristics) will be
discussed together and, if relevant, related to adhesive practice.

1 RAW MATERIALS

Early PSAs were based on organic solvent solutions of natural rubber that
were tackified by ‘‘some type of resinous material’’ [1]. These so-called
rubber-resin (RR) adhesives are still widely used. With regard to their
chemical basis and application technology, PSAs may be classified into
dispersed systems and 100% solids. Dispersed systems may include solutions
and emulsions; 100% solids cover hot-melt adhesives and radiation-cured
materials (oligomers and prepolymers). The most important solutions are
solvent-based systems. They may include rubber/resin adhesives, acrylics
(thermoplastic and crosslinked), and silicones. Hot-melt PSAs are based
on (mainly styrene-diene) block copolymers and acrylics. The radiation
(electron beam and UV)-cured materials can be either acrylic- or rubber-
based. Finally emulsions (water-borne adhesives) comprise acrylics, natural
and synthetic rubber latexes, and ethylene vinyl acetate dispersions. The
original compositions were heterogeneous physical mixtures of one or
more elastomers (rubbers) with tackifiers and/or plasticizers. The second
approach, essentially homogeneous polymers based mainly on acrylate
esters, was introduced in the mid-1950s. Both techniques were refined and
their usefulness extended as a result of extensive research and development
on new raw materials and crosslinking methods, and both are currently
available in water-borne and 100% solids form as well as in the original
solvent solution. Deficiencies persist, however, especially as the use of
PSAs expanded into new, more demanding markets. Particular needs exist
for PSAs with good aging characteristics, with broader specific adhesion,
higher cohesion and heat resistance, adherend independent removability,
and reduced plasticizer-induced performance losses when used on flexible
Chemical Composition of PSAs 163

polyvinyl chloride (PVC) face stock materials or special substrates. Some


alternative compositional approaches have shown considerable improve-
ment, especially for silicones, but these are too expensive for general use, as
are urethane-based PSAs.

1.1 Elastomers
A main characteristic of PSAs is bond forming (tack). There are three
fundamental criteria that must be met in order for an elastomer to exhibit
tack [4]. The polymeric chain must come into molecular contact with the
surface, thus the chains must interdiffuse across the interface and become
entangled with one another. After bond formation the elastomer must
display high cohesive strength to resist easy separation upon application of a
force. This property distinguishes a tacky elastomer from a low molecular
weight liquid. This combination of partially contradictory characteristics will
be embodied by a viscous/elastic, rubber-like material. Therefore the most
important component of a PSA formulation is the elastomer. Elastomers are
high polymer materials. They are crosslinked and possess a steel-like elastic
behavior. Rubber-like elasticity is characterized by values of the compression
modulus which do not depend on the temperature and by pronounced
reversible deformability. Elastomers do not exhibit viscous flow. From the
theoretical point of view it is possible to transform an elastomer in a visco-
elastomer by blending it with viscous components. The main part of PSA
formulation is based on this principle. PSAs based on natural or synthetic
elastomers are tackified formulations, i.e., multicomponent systems.
A partially crosslinked elastomer (i.e., natural rubber) was used as a
starting material for PSA formulations. Rubber-like synthetic materials,
with a rubber-like (hydrocarbon-based) chemical composition were devel-
oped later and used with or without natural rubber. The development of
soft and elastic polymer coatings (usually acrylic-based) opened up new
directions in the synthesis of elastomers. Later, multicomponent copoly-
merization of nonpolar monomers was developed, which yielded a new
class of soft and elastic, thermoplastic elastomers. Such compounds with
a segregated structure work like a crosslinked elastomer. Heterocyclic
compounds exhibiting a special bond stability function as the elastomeric
backbone of the formulation (e.g., silicones). Thus there is a broad range of
elastomers available for PSA formulations.

Natural Rubber
Natural rubber was used almost exclusively as the base elastomer for PSAs
in the early stages of PSA manufacturing. Diene-based natural and synthetic
164 Chapter 5

rubbers always contain more or less reactive side groups, or branching, and
thus they always possess some crosslinked structure. However, their solid-
like and elastic behavior is not given (it is influenced only) by such structure,
and they are (at least partially) soluble. Mechano-chemical destruction of
the base elastomer (by mastication or dissolution) allows the modification of
the molecular weight or molecular weight distribution. The good compat-
ibility of natural rubber with different tackifiers or plasticizers allows
an easy adjustment of the adhesion/cohesion balance. In some cases more
cohesion is required; therefore blends of natural rubber with styrenic block
copolymers are preferred.
Natural rubber is a well-known component of traditional rubber/resin
systems, but its use (in latex form) in emulsion-based PSAs for permanent
label stock applications is less widely appreciated [5]. Natural rubber latex
can be used in SBR/tackifier resin systems to modify the balance of adhesive
properties. High molecular weight natural latexes contribute to improved
shear resistance, while the lower molecular weight ones enhance the tack
(especially when natural latex is used at less than 10% of the polymer
content in the formulation). Furthermore it was claimed that when SBR and
natural rubber latexes are blended, a product with better aging character-
istics results. Here the SBR latex crosslinks and hardens under oxidative
aging, whereas the natural rubber latex undergoes chain scission and
softens, whence a balance is achieved [6].
Natural rubber latex-based PSAs offer an advantage over solvent-
based systems (e.g., milled smoked sheet) with regard to the molecular
weight difference between the two. The high molecular weight portion of
natural rubber is insoluble in solvents, therefore it cannot be used in solvent-
based adhesives. Natural rubber must be milled to a Mooney viscosity of
53 or below to obtain complete solubility (at 30% solids) [7]. The natural
rubber from dried latex shows a wider plateau of the modulus and a higher
elastic modulus value at 100 C than the milled natural rubber. The
degradation does not change the tan  peak temperature, but it does reduce
the modulus at high temperatures. This modulus reduction relates to the low
shear performance of solvent-based systems. An adequate rubber latex has a
gel content low enough to allow good quick stick, but high enough to give
good holding power when properly formulated [8]. A detailed discussion
about the use of natural rubber per se (solid state) or in latex form will be
presented in Chap. 8.

Synthetic Elastomers
Batch-to-batch inconsistency and staining are two problem areas and/or
disadvantages of rubber/resin PSAs. Molecular weight and molecular
Chemical Composition of PSAs 165

weight dispersion both influence rubber compatibility with tackifiers and


the adhesion/cohesion balance. Unfortunately the molecular weight and its
distribution depend on the natural rubber quality and on the processing
parameters. Therefore the scattering of the adhesive performance values
remains high for natural rubber-based adhesives. For a better reproduci-
bility synthetic rubber-like products were tested. First pure hydrocarbon-
based polymers (polydienes or diene-styrene copolymers) were synthesized;
later acrylic and vinyl monomers were combined. Natural rubber-like
stereoregulated polydienes, diene-styrene random and block copolymers,
and thermoplastic elastomers on styrene-diene, styrene-olefin, ethylene-
propylene-diene, acrylic, etc. basis have been synthesized and tested. As
discussed in detail in [9], unlike natural rubber, certain of these products
can also be used for noncoated pressure-sensitive products.
Various raw materials (monomers) and polymerization methods
were used to prepare rubber-like products. Synthetic, hydrocarbon-based
elastomers, homo- or copolymers, and random or block copolymers (two-
component or multipolymers), stereoregulated or not, are synthesized and
applied as raw materials for PSAs. Low speed crystallization of polyolefins
permits their use as semi-PSAs [10]. Macromolecular compounds based
on dienes were tested too. Stereoregulated 1,4-cis-polyisoprene is recom-
mended for special formulations [11]. On the other hand mono-olefins like
atactic polypropylene are also used. The best results were obtained by
copolymerizing styrene with mono- or diolefins. The first products were
designed for hot-melt PSAs; later carboxylated styrene-butadiene rubber
(CSBR) dispersions were synthesized (see Chap. 8). In order to improve
the adhesion of polybutadiene-based PSAs, these were modified with
isopropyl dicarboxylate [12,13]. Other synthetic rubbers (butylene- and
isobutylene rubber) were tested as tackifiers. The use of methyl rubber and
polybutylene in adhesives was described in the literature [14]. Development
of special products for specific end-uses, clarification of the structure/
property relationships, and introduction of new products such as styrene-
ethylene-butylene-styrene thermoplastic elastomers (TPE) and crosslinkable
derivatives were documented [15]. As discussed in detail in [9], concerning
their chemical build-up and processing technology, synthetic elastomers
used as raw materials for PSAs can be divided into polymers with either
nonsegregated or segregated structure (i.e., common and thermoplastic
elastomers).

Nonsegregated Elastomers for PSAs. Synthetic rubber-based PSA


formulations were developed as solvent-based, water-based, and 100%
solids recipes. The most widely known synthetic elastomers are the styrene-
diene copolymers, but butyl rubber, polyisobutene, acrylnitrile-diene
166 Chapter 5

copolymers, halogenated diene-diene copolymers, silicones, and polyur-


ethanes are employed too. The first trials to synthesize natural rubber-like
elastomers used dienes (alone or as the main comonomer) for the build-up
of a mostly linear, noncrosslinked polymer. Later different comonomers
were added and ethylene-, propylene-, and diene copolymers were
manufactured having an elastomer- or viscoelastomer-like behavior. Such
synthetic elastomers possess a random sequence distribution.
Styrene-butadiene copolymers were developed for the rubber industry
to achieve hard and abrasion resistant highly filled vulcanizable composites
for technical products. Synthetic styrene-butadiene latexes were first pro-
duced and sold into adhesive end-uses in 1946 [16]. Generally, SBR
latexes have a solids content of 41–54% with a butadiene content of
70–75%. For PSAs the most effective SBR latexes are those with a styrene
level of about 25–35% and a Tg between 60 to 35 C [17]. Low
temperature synthetic butadiene rubbers (GR-S) with 30% styrene are
soft and tacky; high temperature ones (with 45–80% styrene) are harder
and less tacky [16]. The most important properties of SBR latexes used
for PSAs include their molecular weight, molecular weight distribution,
gel content, Tg, and styrene content [18]. Styrene-butadiene rubbers with
23% styrene are proposed for solvent-based PSAs. In such adhesives the
styrene-butadiene rubber exhibits less cohesion and more tack than natural
rubber.
Styrene-butadiene rubber can be made with or without carboxylic
groups. Polar styrene-butadiene latices are manufactured by emulsion
polymerization of styrene, butadiene, and a small amount of a functional
monomer such as vinyl-carboxylic acid (e.g., acrylic, methacrylic, itaconic,
or fumaric acid) either alone or in combination [18]. The functional
monomers stabilize the product. They are usually present at a level of about
1.0–3.0%. When COOH groups are not present in the polymer chain, SBRs
are usually poststabilized with about 1–4% of surfactants; antioxidants and
pH-adjusting agents are also added [16]. The CSBR dispersions possess
small (typically 0.15 mm) particles with narrow particle size dispersion.
Butadiene can polymerize via two different modes of addition, namely the
1,2- or 1,4- configuration. Both configurations contain residual double
bonds capable of further polymer growth, either branched or crosslinked.
The resulting three-dimensional network structure improves the mechanical
properties apparently without affecting the Tg.
The gel content may be adjusted through the ratio of styrene to buta-
diene and the molecular weight. Tack increases with increasing butadiene
content (50–70%), but changes little at higher levels. When measuring peel
adhesion the failure mode is adhesive at lower butadiene levels and cohesive
at high butadiene levels.
Chemical Composition of PSAs 167

Shear strength is higher as the styrene content increases. Above 30%


gel content (70% butadiene content) the tack decreases rapidly. The usual
range of hot carboxylated SBR latexes runs from about 40% styrene to
70% styrene; the normal range for PSAs latexes generally lies below 45%
styrene.
For many systems the Tg is fixed at 55 C for use in combination with
an 80 C softening point resin [19]. The molecular weight and the Tg of
carboxylated butadiene polymers constitute major variables that determine
the pressure-sensitive properties.
Carboxylated SBR latexes consist of a mixture of polymer species
having linear chains, branched chains, and crosslinked materials. The
relative ratio of the species can be controlled by changing the concentration
of molecular weight regulators, and by varying factors such as conversion,
reaction temperature, or polymer particle number [19]. In some cases, for
the characterization of CSBR, the acid equivalent (carboxyl equivalent)
for 100-g latex solids is given [8]. Carboxylated neoprene latexes have the
following advantages [8,20,21]: mechanical stability, resistance to electro-
lytes, crosslinking at room temperature, thermal stability, good polyethylene
(PE) adhesion, high shear, low peel, removability, temperature resistance,
and conformance with FDA regulations (e.g., 175.105 for adhesives).
Aqueous acrylic and modified acrylic PSAs possess certain limitations,
including poor adhesion to polyethylene and polypropylene surfaces.
Attempts to improve their performance by the incorporation of tackifiers
have met with only limited success [18]. Tackified carboxylated styrene-
butadiene rubber latex systems, on the other hand, seem to be able to
overcome the apparent inherent limitations of water-based systems and yet
match the superior performance characteristics of the solvent-based rubber/
resin systems. Until the 1980s, relatively few SBR latexes were available with
the right characteristics necessary for PSAs. The most important factors
in the design of CSBRs for PSAs and their interaction with the other
components of the system were discussed in detail [18]. When using CSBR
latexes it is better to produce a low viscosity latex and control the final
rheology of the adhesive through the addition of thickeners. A comparative
evaluation of the formulation and end-use properties of CSBR and acrylic
PSAs was summarized by Benedek [22] and will be discussed in Chap. 6.
Some years ago contact adhesives on neoprene basis were developed
[21]. Later carboxylated neoprene latexes with better tack for PSAs
applications were synthesized [20]. There is a difference between diene-
styrene and diene-polar comonomer copolymers. Polychloroprene exhibits
strain-induced crystallization (like natural rubber, but uncontrolled).
Styrene-butadiene copolymers do not display such comportment. As a raw
material for PSAs, functionalized chloroprene latices are more important.
168 Chapter 5

The formulation and end-use of carboxylated latexes will be reviewed in


Section 2.9 of Chap. 8.
Polymers and copolymers of C3 and C4 alkenes include a broad
range of plastomers, viscoelastomers, and viscous compounds which can
be used as various components in PSPs. The most important products are
poly(butenes/isobutenes) and amorphous polyalphaolefins. Ideally diene
polymerization supposes 1,4-addition of the diene units with the formation
of a natural rubber-like polydiene. In industrial practice 1,2-polymerization
may occur, i.e., vinyl units may be inserted in the main polymer backbone.
In an extreme situation the polymer backbone is built up from vinyl units
and a low level of diene units. Butyl rubber is such a copolymer of isobuty-
lene and 1–2% isoprene. Butyl rubber may be used as a base elastomer in
PSAs. Its molecular weight and molecular weight distribution influence the
adhesive performance. Generally blends of butyl rubber having different
molecular weight are used.
Polyisobutylene is an elastomer or plasticizer too. Polyisobutylene
(MW 3000) is liquid whereas high molecular weight polyisobutylene (MW
200,000) is a clear solid. The isotactic polymers are to be distinguished from
polymers of isobutene (normally termed polybutene) which are used as oily
additives, and from amorphous, atactic poly(1-butene) polymers which
range from viscous oils to rubbery polymers. The polybutenes used as
plasticizers are isobutene-butene copolymers synthesized from macromole-
cular mono-olefins with low isoparaffin content. Polybutenes are saturated
therefore they exhibit aging resistance.
Amorphous poly(alphaolefins) encompass a group of usually low
molecular weight polyolefins which are obtained by coordination polymer-
ization [23]. Atactic polypropylene (APP) is a byproduct in the production
of isotactic polypropylene, using first-generation Ziegler-Natta catalysts.
The atactic polypropylene obtained was first used as a raw material for
HMPSAs. Later ethylene-propylene copolymers were manufactured. Such
polymers cannot be used as adequate raw materials for HMPSA. Their
tackifying leads to softening, but such products do not have the required
viscoelasticity. The raw materials used for PSAs are the distinctive product
types of APAO made by direct reactor synthesis, namely homopolymers
of propylene, copolymers of propylene and ethylene, and copolymers of
propylene and 1-butadiene [23].
Proper selection of comonomers and synthesis conditions has led to
the development of olefin copolymers which are pressure-sensitive [24,25].
The copolymerization of propene with long-chain olefins leads to amorphous
and tacky products. The best results have been obtained with hexene and
octene as comonomers. Because these materials are neat olefin polymers,
they offer several advantages over typical styrene block copolymer-based,
Chemical Composition of PSAs 169

tackified PSAs. These include the absence of oils which can bleed into the
substrate, fewer skin sensitivity problems, a good thermal stability (<10%
viscosity decrease after 100 hr at 172 C in a forced-draft oven), low color
and odor, the ability to be blended with most olefin compatible adhesive
ingredients for common applications, and finally a lower density than con-
ventional rubber-based hot-melt PSAs leading to a decreasing usage of raw
materials. These raw materials are mainly amorphous terpolymers of alpha
and omega olefins. A range of properties can be obtained using this new
amorphous polyolefin (APO) technology. Examples include a Tg between
10 to 40 C, a probe tack of 0–800 g, and a static shear ranging from 1 to
30 h. The 180 peel strength (PSTC-1), 90 quick stick (PSTC-5), and shear
resistance (hr) amount to 2.5, 0.8, and 8, respectively.
New, mainly amorphous ethylene-propylene-butene-(l)-terpolymers
can be made tacky quite easily and they exhibit a low speed of crystal-
lization. There are propene-rich and butane-rich copolymers. Propylene/
hexene copolymers were also tested for hot-melt PSA applications [15].
There are pressure-sensitive and semi-pressure-sensitive products. The semi-
pressure-sensitive products display limited pressure-sensitivity in time. They
lose their pressure-sensitivity due to crystallization. With hot-melt adhesive
formulations with a long ‘‘open time’’ one takes advantage of the slow
crystallization. During cooling these products exhibit properties like those
of a hot-melt PSA for a long period; afterwards, however, they are no longer
tacky. They appear quite suitable for application by spraying. Hot-melt
PSAs with an excellent price/performance ratio are based on a highly
viscous, amorphous poly(alphaolefin) (APAO), rubber, resin, and a combi-
nation of plasticizers. The latter consist of a mineral oil and polyisobutylene
(see Section 1.2).
High molecular weight elastomers based on acrylates have been
produced by emulsion polymerization. Such compounds do not contain
carbon-carbon double bonds as polydienes do. Their backbone includes a
saturated alkyl and alkoxyacrylate (95–99%) and 1–5% curing monomers,
having a chlorine or hydroxyl functionality in the side chain. The main
polymer chain is based on ethyl hexyl or butyl acrylate. These polymers can
be used for HMPSAs (reactive hot-melts), solvent-based PSAs, pressure-
sensitive sealant tapes, and sealant caulks. They are used in tackified or
plasticized formulation [26] (see Section 1.2 also).
Silicones are used as raw materials for PSAs as well as coated on
release liners. In some cases silicone-grafted acrylic copolymers were used in
order to achieve good removability. Silicone-based PSAs were developed in
1953 [27]. Silicone PSAs are based on the combination of a ‘‘gum’’ compo-
nent (highly flexible, linear siloxane rubber) and a (extensively branched)
resin [28]. The gum component is a polysiloxane gum, such as methyl phenyl
170 Chapter 5

polysiloxane; the resin is also an organopolysiloxane. Specific resins are


reacted with linear, high viscosity organopolysiloxane fluids [29]. For
instance, for an HM silicone PSA, the resin/silicone ratio was adjusted from
54/45 to 65/53 [30].
Industrially, polydimethylsiloxane and polydimethyldiphenylsiloxane
are the main polymers for the production of silicone PSAs. For elevated
thermal stability phenylsilicones are suggested, and for better adhesion
characteristics and improved filler loading capability methylsilicones are
suggested. Such polymers can be crosslinked with benzoyl peroxide. First
low solids (60%) silicone adhesives were manufactured. New silicone
PSAs are based on a silylvinyl-silylhydride reaction (hydrosilylation). Such
addition-based crosslinking is catalyzed by platinum. Such formulations
possess 80–90% solids.
Silicone polymers have inherently flexible siloxane backbones and a
low Tg. Polydimethylsiloxane exhibits a Tg of about 120 C. In comparison
acrylics show a glass transition temperature of about 65 C to 40 C. Both
contribute to excellent performance at low temperatures; they also offer
good weatherability [29]. The performance properties of silicone PSAs can
be designed by changing the resin/gum ratio [30]. A comparison was made
between a typical dimethyl-based silicone and an acrylic-based organic PSA
[31]; the silicone PSAs display good tack, peel, and shear at temperatures as
low as 20 C.
Problems with silicone PSAs may arise concerning their shelf life and
the availability of adequate release liners. While silicone PSAs have existed
for over 15 years, there was no adequate release technology available for
them; their use was limited to industrial tapes. Several products on the
market utilize dimethyl silicone in combination with diphenyl silicone PSAs.
While diphenyl silicone PSAs have some advantages in specific applications,
dimethyl silicone PSAs represent the bulk of silicone PSAs because of
product cost. New release liner technology based on fluorosilicones was
developed, providing a release surface for several types of silicone PSAs.
A low energy surface does not suffice to offer release; this condition must
be satisfied, but other conditions that favor a trend towards a low release
energy are a contact angle greater than 30 and a negative spreading
coefficient. The surface tension of the adhesive must be higher than that of
the release coating. Fluorosilicones have critical wetting surface tensions
of 15–18 mN/m or lower, but in general do not perform as release surfaces
for PSAs because they do not display sufficient flexibility and ability
to reorient their dimethyl groups like polydimethyl silicones. Crosslinked
silicones may also be used as a release layer for silicone PSAs [32]. The
organopolysilicones may be produced directly by the hydrolysis of organic
substituted halosilicones and depolymerization, or by hydrogenation of
Chemical Composition of PSAs 171

higher molecular weight organopolysilicones and heat treatment. Special


silicone groups can be built into common polymers too. Such formulations
are discussed in detail in [33].

Segregated Elastomers for PSAs. As discussed in detail in [9] and [34]


the properties of polymers can be significantly improved by build-up of a
multiphase, composite structure. Elasticity requires a reversibly deformable
network. The existence of a network does not allow viscous flow. This
dilemma can be avoided by synthesizing and formulating viscoelastic
compounds that have partially segregated structures and behave like elastic,
crosslinked materials or viscous polymers depending on the temperature and
their formulation. In such heterogeneous systems one polymer exists above
the Tg, while the other exists below its Tg. The one component is glassy, the
other is rubbery. The main segregated polymers used as raw materials
for PSAs include: styrene block copolymers, and acrylic, polyolefin, silicone,
polyurethane, and polyester block copolymers.
Thermoplastic elastomers have been known since the 1960s [35].
In 1960 Shell developed Kraton in the USA, and in 1974 Cariflex; later,
ethylene butylene copolymers (SEBS) were developed. The first styrenic
thermoplastic elastomers were introduced to the market in 1964 [36]. The
styrene block copolymers used today usually have a molecular weight in the
range of 60,000–200,000 and a styrene content of 15–45% [37]. Because of
their hard segments such polymers are always harder than natural rubber.
Since their introduction styrenic block copolymers have found their way
into numerous and varied application fields in the adhesives industry. Their
molecular structure allows these thermoplastic rubbers to be formulated
and coated in solvent-free hot-melt systems which are widely accepted for a
variety of flexible assembly adhesives and as well as PSAs for labels and
tapes. Many variations can be made in the structure of these block
copolymers which has led to a wide range of grades that differ in physical
properties and are well suited for more specific applications. The most
commonly used block copolymers for hot-melt PSAs still are the linear
SIS grades with low polystyrene content, although interest is growing in
the use of low viscosity SBS grades in combination with newly developed
tackifying resins. Styrene-isoprene block copolymers are more expensive than
natural rubber. The chemistry of styrene block copolymers is discussed in
detail in [26].
Generally block copolymers used for PSAs contain three blocks A-B-A
[37] where A is an amorphous polymer with the Tg above room temperature
(thermoplastic), and B is an amorphous polymer with Tg much lower
than room temperature (in rubber, for example, A is polystyrene and B
polydiene). Because of the incompatibility of polystyrene endblocks
172 Chapter 5

and elastomeric midblocks, there are two microdomains in the polymer;


a discontinuous phase and an agglomeration of polystyrene endblocks is
formed through van der Waals forces. The first thermoplastic elastomers
(TPE) were of the linear triblock form, SBS or SIS [38]. The key to their
function is what was called a ‘‘spaghetti and meatballs’’ morphology. At
an appropriate balance of endblock polystyrene molecular weight vs. those
of the elastomer midblock, a two-domain structure is formed which
consists of polystyrene islands in a rubbery ocean. Polystyrene is not
miscible in polydienes and the polystyrene blocks at the end of each polymer
molecule associate with each other, forming an ‘‘associative crosslink’’ at
room temperature which behaves like a covalent vulcanized bond. If the
polystyrene content is lower than 50%, polystyrene domains are formed
through the incompatibility of the polystyrene with the polybutadiene
sequence.
In general, thermoplastic elastomers are block copolymers or
polyblends with separate phases [36]. In these copolymers with an A-B-A
structure, the A sequences contain 200–500 units and the B sequence
500–700 units [39]. The basic polymer parameters of SBCs include the
nature of the diene and vinylaromatic monomer (e.g., butadiene, isoprene,
styrene, -methyl-styrene, etc.), molecular structure (e.g., linear or radial),
molecular weight, sequence distribution (e.g., diblock, triblock, etc.),
and styrene content. The main midblock sequences are built up from
polybutadiene (in SBS), polyisoprene (in SIS), ethene-buthene (in SEBS),
ethene-propene (SEPS), and functionalized (maleinized or silanized) diolefin
derivatives, with polar groups exhibiting better adhesion to polar substrates
[40]. The most commonly used rubber is of the SIS type, such as Kraton D
1107 [38]. The SIS types are widely used in hot-melt PSAs, but are too soft
for non-pressure-sensitive hot-melt PSAs [41]. Because of the higher stiffness
of the polybutadiene blocks such polymers are recommended mostly for
non-pressure-sensitive applications.
The molecular structure of SIS polymers allows them to be dissolved,
formulated, and applied in solvent, or to be formulated and coated in a
hot-melt system at temperatures higher than the Tg of the styrene endblocks.
Styrene block copolymers can be used as high solids (60–70%)
solutions, or as hot-melt. Following evaporation of the solvent or cooling
down of the hot-melt, the physical crosslink structure rebuilds to provide
the desired strength and performance characteristics. As thermoplastic
elastomers have a less stable elastic domain and a second Tg (identical with
the melting point) they act like thermoplastics above this temperature.
Unfortunately, the physical bonding is less stable [42]. Moreover, the nature
of an SIS polymer prevents its use in PSAs that must withstand solvent or
high-temperature exposure at or above Tg of the polystyrene endblock.
Chemical Composition of PSAs 173

Recently new SIS polymers were introduced which can be crosslinked


through electronbeam (EB) curing. An example of these new radiation-
sensitive polymers (Kraton D 1320 X rubber) shows good crosslinking with
an electron beam dose of 5–7 Mrad [37,38]; it is a multiarmed SIS block
copolymer with only 10% styrene content. Further development of the
styrene block copolymers concerns branching, the sequence distribution,
and the soft monomer unit.
As discussed in [26] side chain segregation depends on the flexibility
and length of the branches. The compatibility of the polystyrene endblocks
and of the elastomeric midblocks depends on their molecular weight and
their volume fractions in the mixture. Phase separation in the radial block
copolymer is higher. Therefore physical crosslinking is higher. ABA and AB
block copolymers of the same type of composition in another geometry
(star-shaped or radial copolymers) have also been developed [43]. According
to the structure and placement of the side chains, branched, radial teleblock,
and multiarm star block copolymers are known [44]. Such polymer
structures are described by the terms ‘‘branched,’’ ‘‘radial,’’ and ‘‘star.’’
Radial and teleblock copolymers with isoprene branches were
prepared, but butadiene-based armed polymers have been proposed also
[45]. The number and molecular weight of the arms influences the solid state
morphology of such polymers. Structures with more than eight arms are
bicontinuous and ordered. According to [44] a formulation with a tackified
star block copolymer gives a higher shear resistance. The mechanical
properties of a radial block copolymer are better. A desirable melt viscosity
and shear resistance can be obtained by selecting only star block copolymers
with six or more arms. Star-shaped styrene block copolymers have an
(apparently) higher molecular weight. Therefore they are more reactive to
radiation curing. Because of their star structure, partial crosslinking does
not modify (essentially) the flexibility of the macromolecules.
Polystyrene-polybutadiene-polystyrene polymers are completely
amorphous and therefore show limited compatibility with waxes; the
polybutadiene midblock is unsaturated and therefore exhibits limited
stability during hot-melt processing. The saturated midblock thermoplastic
elastomers are commercial ABA-type block copolymers in which the
polyisoprene or polybutadiene is hydrogenated. The strength and stiffness of
the block copolymer depend on the chemical nature of the segments and
their mutual interaction. Block copolymers having the same styrene content,
about the same block molecular weight, but different types of midblock
display different mechanical properties. The SEBS polymer has the highest
strength and modulus. The commercial SEBS polymers have a much better
stability during hot-melt processing because their midblock is a saturated
polyolefin rubber. The EB rubber midblock of the commercial SEBS
174 Chapter 5

polymers is essentially a random copolymer of ethylene and 1-butene. In the


commercial SEBS polymers the 1-butene content is controlled at a level high
enough to make the EB copolymer midblock almost totally amorphous,
making the copolymer thoroughly rubbery and soluble in hydrocarbon
solvents. In SEBS the saturated polyolefin confers excellent resistance to
light, ozone, and heat to the polymer. Styrene block copolymers are usually
synthesized via a three-step process. A new process produces styrene block
copolymers via a two-step process using a difunctional initiator instead of a
monofunctional one [36].
In this new technology the midblock (e.g., the polyisoprene or
polybutadiene) is built first and the polymer chain grows in two directions at
the same time, with two reactive sites, one at each end. In the second step
two polystyrene blocks of the SIS (or SBS) are built at the same time. As a
result a pure 100% triblock polymer is formed. Uncrosslinked and unentan-
gled polymer chains do not contribute to the strength of the elastomer.
Therefore pure triblock SIS copolymers have also been manufactured. Such
pure triblock copolymers have a high-Tg endblock [46]. Diblock units work
as low cohesion diluting agents and decrease the cohesion. SAFT decreases
with the diblock content. Increasing the diblock content improves peel and
tack. A range of copolymers with up to 42% diblocks have been synthesized.
In [47] blends with up to 54% SI were studied. Differences due to the change
of the SI content occur in the frequency range of 5  103 – 5  101 rad/sec,
where an increase in the SI content leads to a decrease of the elastic modulus
G0 and to an increase in the loss modulus G00 . In this range the blends are
more dissipative if there is more SI. Tensile tests show that decreasing the SI
content or increasing the crosshead velocity leads to a higher stress for all
extensions and to a more pronounced strain hardening. Other substituted
styrene derivatives (e.g., p-t butylstyrene, n-propylstyrene sulfonate,
sulfonated SEBS block copolymer, etc.) were investigated also [48–50].
Another regulating possibility of the copolymer performance is given
by the nature of the hard monomer. Generally the hard segment size
influences the melt transition temperature of thermoplastic elastomers. For
instance -methylstyrene has a higher Tg and therefore provides better
thermal properties than styrene [51].
In the early 1980s HMPSAs based on acrylic block copolymers were
introduced. Phase-separated acrylic block copolymers display a thermally
reversible crosslinking. Unfortunately such experimental copolymers
possess a high viscosity, low tack, and shear. The hard blocks of acrylic
copolymers can contain styrene derivatives also. Polyurethane-acrylate hot-
melt adhesives have been prepared too. As discussed in detail in [26] many
attempts were carried out to prepare segregated acrylic block copolymers;
however, their industrial applicability is limited. A better approach is given
Chemical Composition of PSAs 175

by the use of acrylic oligomers and prepolymers able to be macromerized by


polymerization or crosslinking (see later).
Special, segmented EVAc copolymers have been developed as raw
materials for HMPSAs. They have a higher vinyl acetate content, and can be
considered elastomers. Such polymers possess alternative segments with low
and high vinyl acetate content.
As mentioned earlier, synthetic hetero block copolymers were
developed also. Such nonhydrocarbon-based TPEs contain heteroatoms
(e.g., silicones, polyurethanes, etc.). The mixing of siloxane resins with a
silicone elastomer in 1953 led to the discovery of silicone PSAs [27]. As
described in [52] the high solids silicone adhesives are multicomponent
systems which contain at least four components: a base elastomer, a tackifier
resin, and at least two crosslinkers. Such silicones have a block structure in
comparison with the random build-up of peroxide-cured silicones.
Polyurethane adhesives are used for special applications. Two-
component and single-component polyurethanes were developed. The
adhesive polymers are composed of end-functionalized soft segments
(polyether or polyester macroglycols) connected via a urethane linkage
and a spacer to an isocyanate group. The spacer is typically an aromatic
unit arising from 4,40 -methylene bis(phenylisocyanate) (MDI) or toluene
diisocyanate (TDI) used during synthesis. Two-component-based reactive
polyurethanes result from the combination of polyol-polyisocyanate or
-isocyanate terminated prepolymers cured with polyamines or polyols. The
polyether or polyester is flexible.
Single part polyurethane adhesives are polymers of low molecular
weight, whose chains are terminated with isocyanate groups. These NCO
groups are free to react with moisture, either present already on the web or
added as water fog at the nip station. Polyurethane adhesives with 100%
solids, solvent-based ones, and dispersions were developed. Emulsifier free
and (cationic, anionic, and nonionic) emulsifier containing polyurethane
dispersions can be manufactured.
Thermoplastic urethanes are synthesized from diisocyanates, short
chain diols, and polymeric diols. Generally the isocyanate can be regarded
as a stiff component of the copolymer, while the polyol may be regarded
as a relatively flexible component. The stiff segments tend to form laterally
ordered groups through intermolecular bonding. Flexible groups are dis-
ordered and randomly oriented. They are necessary to provide toughness
and flexibility. If the flexible backbone is sufficiently long, a certain amount
of coiling occurs, giving rise to rubbery domains. Polyesterdiols and poly-
etherdiols can be used as soft segments for thermoplastic urethanes.
Polyetherdiols have various chemical bases, structure, and molecular
weight. They include mainly C3 and C4 ethers [53].
176 Chapter 5

The Tg and morphology of such compounds are influenced by the


length and concentration of the polyethers. A foamed adhesive for tapes was
manufactured from a mixture containing an isocyanate terminated mono-
mer or oligomer, and a polymer comprising a backbone of polymers such
as polybutadiene, polyesters, and polyether which must contain at least
two active hydrogens capable of reacting with the isocyanate terminated
monomer and expandable filler. Such foam can be manufactured in situ [54].
Special polyurethanes contain vinyl unsaturation as the reactive site.
Thermoplastic elastomers may contain oxygen as a heteroatom in block
copolymers based on esters. Such a copolymer can have hard segments of
polybutylene terephthalate and flexible segments of polyalkylene oxide
terephthalate. It displays insufficient thermal resistance and inadequate pres-
sure sensitivity. Better results were obtained with PSA formulations based
on radiation curable macromers (see later). Special TPEs were discussed in
detail in [55].
As discussed above elastomers are used in two-component PSA formu-
lations. As described in [56,57] in some special cases certain base elastomers
are used per se, without tackification. Natural rubber latex, CSBR, or
butene derivatives can be employed for special products (mostly with low
adhesivity) without tackification.

1.2 Viscoelastomers
The special structure of the diene monomers (e.g., butadiene, isoprene, etc.)
leads to macromolecular compounds displaying rubber-like properties. In
order to achieve viscoelastic properties the natural, synthetic diene-based
elastomers must be combined with tackifiers. In order to achieve fluidity at
the processing temperature and a good adhesion/cohesion balance at room
temperature, synthetic block copolymers (with a crosslinked structure) were
synthesized. Both tackifying or internal crosslinking are used in order to
achieve balanced PSA properties. The mostly nonpolar backbone of the
hydrocarbon-based elastomers requires formulating in order to achieve a
suitable adhesion/cohesion balance. The use of polar monomers allows the
synthesis of single-component, unformulated elastomers with an adequate
viscoelastic balance. The most common monomers used for this purpose
are acrylic and vinyl monomers. It is evident that acrylics have the
most sophisticated chemical basis and therefore may satisfy the extreme
requirements concerning the different physical state (solvent-based,
dispersed, or hot-melt PSAs) and end-use properties (permanent/removable,
paper/film). On the other hand, the new ethylene multipolymers will
win new fields of application due to their dual (technical/economical)
advantages.
Chemical Composition of PSAs 177

Acrylic Viscoelastomers
It is generally agreed that PSAs based on acrylics are the most important
adhesives for the production of label stock. In the past several years new
classes of raw materials for PSAs (e.g., CSBR, ethylene vinyl acetate
copolymers) were developed. There was a considerable interest recently in
a comparative evaluation of acrylics and other raw materials in order to
forecast development trends. The development of acrylic polymers allowed
the synthesis of pure elastomers that are not crosslinked and may present a
viscoelastic behavior. Later, other comonomers, usually vinyl acetate and
its derivatives, were copolymerized with acrylics mainly for economical
reasons. The search for a low-cost, low-Tg monomer forced the development
of ethylene vinyl acetate elastomers and, related to these compounds, a new
renaissance of the maleinate copolymers. Actually there is a broad range
of uncrosslinked elastomers and viscoelastomers that are suitable as base
materials for PSAs.

Pure Acrylics. Acrylics have been used in PSA applications since 1928
[58]. In a manner different from natural and synthetic rubbers, hydrocarbon-
based acrylates (AC) are synthesized from esters of polar, organic acids,
mainly acrylic and methacrylic acid [59]. The comonomers are the acids
or other nonacrylic monomers. The choice of the main monomer determines
the Tg of the polymer and its characteristics, especially the tack. Introduction
of comonomers allows the (re)adjustment of the adhesion/cohesion balance.
Comonomers with reactive functional groups allow crosslinking and thus
the improvement of the shear resistance. An advantage of the acrylics is
their synthesis-based pressure-sensitive character (i.e., they possess pressure-
sensitive properties without additional tackifiers and they display stable
storage and aging properties). The acrylics are inherently tacky due to
their high entanglement molecular weight (Me), without the need to add
tackifier [60]. Moreover, acrylics are not sensitive towards oxidation.
Their adhesion, cohesion, and tack do not change after long aging periods
at high temperatures. Another advantage of acrylics is their lack of sensitivity
towards UV light; acrylics are colorless and they do not change their
performance characteristics during aging; they also display high transpar-
ency. Acrylics are resistant towards plasticizer migration; they do not contain
oligomers and therefore they do not give rise to migration. Acrylics are polar,
so they display a good adhesion towards polar substrates. Thanks to an
adequate adhesion/cohesion balance, acrylics usually display good wet-out
properties and die-cuttability.
Because of their wide chemical basis and well-known chemistry it is
possible to polymerize acrylics in solution, dispersion, or in the bulk state;
178 Chapter 5

there are solvent-based, water-borne, hot-melt, and radiation-curable acrylic


PSAs. The relatively low molecular weight and linear, gel-free structure
of solvent-based acrylics allows the use of a broad range of solvents, while
the chemically inert nature of these solvents makes the use of different
crosslinking agents possible. On the other hand the polarity, water insolu-
bility, and high reactivity of the main acrylic monomers allow the synthesis
of mechanically stable, protective, colloid-free, ‘‘grit-free’’, aqueous disper-
sions with a high solids content. At the present time hot-melt acrylic PSAs
are at the developmental stage; this is also the case for radiation-cured,
solventless adhesives.
Polyacrylates are used mostly as solutions or dispersions. Solid state
polyacrylate rubbers are also known [61]. They are produced by emulsion
polymerization technology to form a latex, which is subsequently coagu-
lated, washed, and dried in sheet form. Polyacrylic rubbers possess excellent
physical characteristics. They are resistant to a wide range of chemicals and
oils, have good low and high temperature resistance, and a high degree of
resistance to weather, atmospheric oxidation, and UV due to the saturated
backbone. Polyacrylates do not contain double bonds (as in diene rubber).
They are basically saturated copolymers containing 95–99% backbone
monomers consisting of alkyl, or alkyl and alkoxy acrylates such as ethyl,
butyl, methoxy ethyl, and ethoxy ethyl, and 1–5% cure site monomers with
chlorine or hydroxy functionality for vulcanization purposes. The main
applications of polyacrylate rubbers are solvent-based and hot-melt PSAs
(100% solids) pressure-sensitives. New solid and liquid polyacrylates with
many functional groups (e.g., hydroxy, carboxy, epoxy, and isocyanate
groups) for possible crosslinking were also developed. Physical character-
istics of polyacrylic elastomers include resistance to temperature (40 to
200 C), petroleum, synthetic oils, aliphatic hydrocarbons, oxidation, ozone,
and UV light.
Even though it is not the aim of this book to discuss the synthesis and
monomer basis of PSAs, a short overview of the acrylic components follows.
Long-chain polyacrylates do not exhibit sufficient cohesive strength; their
cohesion may be improved by copolymerizing with n-butyl acrylate, acrylic
acid, and other polar comonomers such as glycidylmethacrylate, N-vinylpyrr-
olidone, methacrylamide, acrylonitrile, etc. [62]. The patent literature gene-
rally identifies alkyl acrylates and methacrylates of 4–17 carbon atoms as
suitable monomers for PSAs [63]. The most commonly used monomers are
2-ethyl-hexyl acrylate, butyl acrylate, ethyl acrylate, and acrylic acid.
Monomeric acrylates or methacrylate esters of a nontertiary alcohol (having
1–14 carbon atoms) with the average number of carbon atoms between
4 and 12 may be used [64]. Additional monomers (e.g., itaconic acid,
methacrylic acid, acrylamide, vinyl acetate, etc.) can be included in minor
Chemical Composition of PSAs 179

amounts (e.g., a 74:20:6 iso-octyl acrylate:n-butyl acrylate:acrylic acid


monomer ratio or a model composition having 97.5/2.5 wt% of EHA or
BuAc and AA [65]). Similarly other acrylates derived from alcohols with
4–14 carbon atoms may be included.
As discussed earlier in Chap. 2, definite trends in physical properties of
the polymers can be observed, when higher molecular weight acrylic esters
are used. As their ratio to low molecular weight monomers is increased, the
following changes in properties occur, namely, the tackiness increases, the
hardness decreases, the tensile strength decreases, the elongation increases,
and the water absorbtion decreases.
The term hard monomer refers to a monomer which when homopoly-
merized yields a polymer with a Tg above 25 C, preferably above 0 C.
Among such monomers are methyl acrylate, alkyl methacrylates (e.g.,
methyl methacrylate, ethyl methacrylate, and butyl methacrylate), styrenic
monomers (e.g., styrene and methyl styrene), and unsaturated carboxylic
acids (e.g., acrylic acid, methacrylic acid, itaconic acid, and fumaric acid).
The term soft monomer refers to a monomer which when homopolymerized
yields a homopolymer with a low Tg (i.e., less than 25 C). Examples
are the alkyl acrylates (e.g., butyl acrylate, propyl acrylate, 2-ethyl hexyl
acrylate, iso-octyl acrylate, and isodecyl acrylate). Presently EHA and/or
butyl acrylate are preferred [66]. Soft monomers, like the diester of fumaric
acid (dibutyl fumarate) with 2–8 atoms in the ester group or alkyl acrylate
with 2–10 carbon atoms (e.g., EHA), should be incorporated at 50–85% in
PSAs [67]. Hard monomers like alkyl methacrylate (e.g., MMA) with 2–6
carbon atoms and unsaturated carboxylic acid (e.g., acrylic acid) with 2–8
carbon atoms should be incorporated at 0–10% in PSAs. As an ideal for-
mulation for solvent-based and water-based PSAs, the following composition
was given [68]: 70–90% soft monomers, 10–30% hard monomers, and 3–6%
functional monomers.
Kuegler [69] discussed copolymerization of acrylic esters with a mini-
mum amount of fumarate diesters. Copolymers containing octyl acrylate,
ethyl acrylate, vinyl acetate, and maleic anhydride were patented by the
National Starch and Chemical Corporation some years ago. These PSAs
have a Tg range of 45 to 65 C [70]. Emulsion acrylic PSAs from acrylic-
based fumaric ester interpolymers may be synthesized when peel and tack
values can be modified without affecting the cohesive strength [67].
It should be noted, that some acrylic comonomers play a special role
in PSA synthesis. Such monomers assure enhanced chemical reactivity or a
physical reinforcing effect. Such monomers are: acrylic acid, acrylnitryl,
and certain polyfunctional acrylates. As discussed earlier, acrylic acid is
incorporated in PSA compositions in order to improve the polarity and/or
to allow crosslinking reactions. The role of the carboxyl (COOH) sticker
180 Chapter 5

groups in the peel adhesion strength of elastomers is well known. Adhesion


went through a maximum at some optimal concentration of such groups
(ca. 3 mol%) [71]. For polymers with acrylic acid a much higher adhesion
hysteresis is obtained as compared to copolymers without AA [72]. This
might be partly due to the hydrogen bonding formation and surface
rearrangements during contact when AA is present. Acrylic acid content
dramatically increases the adhesion and adhesion hysteresis of acrylic
copolymers. The addition of AA also decreases the critical crack propaga-
tion speed above which a much stronger rate dependence of the adhesion
energy is observed. Pocius et al. [73] have studied the release behavior
of acrylic tapes from release coating based upon long alkyl side-chain
polycarbamates. Pressure-sensitive adhesive-like networks with 10% AA
show different release behavior from those without acrylic acid. The work of
adhesion and the interfacial energy differ for PSAs with or without acrylic
acid. As described in [73] the use of PSAs containing AA on vinylcarbamate
release coating may lead to blocking.
Acrylnitryl is incorporated in PSA copolymers in order to improve
their cohesion (through the special physical interactions exerted by the nitryl
group). As discussed in detail in [74–85] such an effect can be observed
in various acrylic or vinylaromatic, heterocyclic or diene copolymers of AN
and acrylates. In some cases a special sequence distribution can be achieved
also [80,81].
Polyfunctional acrylates are used generally as crosslinking monomers.
For instance 1,6-hexane diol diacrylate (HDDA) was proposed for acrylic
model compounds [73]. Such monomers are used in the off-line or in-line
synthesis of PSA raw materials (see Section 1.3).
The polymerization technology influences the properties of the acrylic
copolymers. Generally acrylates can be polymerized in solution or emulsion.
Solution or water-based emulsion polymerized products differ concerning
their macromolecular characteristics (molecular weight, molecular weight
distribution, and gel content). Research has shown that emulsion polymeri-
zation of acrylates carried out to complete conversion produces significant
amounts of microgels inside the particles due to chain transfer to the
polymer via hydrogen abstraction [86]. According to [65], for gel-free,
low-Tg acrylics (based on EHA or BuA) there was no significant difference
between emulsion vs. solvent-borne PSA film adhesive performances. On
the other hand emulsion and suspension polymerization give different PSA
raw materials as a function of the particle size [65,87]. Solution polymer-
ization carried out using batch or continuous feed, with constant or variable
comonomer ratio led to different products [88]. For instance, a critical
fraction of EHA was determined for which there is a transition between
adhesive and cohesive failure. The critical fraction depends on the
Chemical Composition of PSAs 181

polymerization process. Depending on the polymerization yield the


comonomer content varies also. Another parameter of influence is if
tackification was carried out during polymerization or during formulation
[89]. The phase structure of PSAs mixed in the monomer state and solution
polymerized was different than that expected from the phase diagrams of
solution blended systems. The lower molecular weight of acrylates and some
reactions of acrylates with rosin caused the first system to be more
compatible.
Natural rubber was used almost exclusively as the base elastomer
for PSAs, in the early stages of PSA manufacturing [90]. It remains difficult
to synthesize high molecular rubber-like acrylics with a low softening point
range, although polyacrylate elastomers have been known for about 20
years. There are several patents including such recipes, but their practical
use is limited. For instance acrylic hot-melt PSAs with an improved creep
resistance at ambient temperatures and desirable melt viscosity at elevated
application temperatures are prepared by copolymerizing acrylic and
methacrylic acid and alkyl esters with 10–40 wt% of an acrylate- or
methacrylate-terminated vinyl aromatic monomer-based macromolecular
compound [91].
In recent years acrylic-based adhesives which possess a balance of
high tack, high peel, and high shear properties were prepared by coating a
blend of monomers (such as iso-octyl acrylate and acrylic acid) on a web,
maintaining an inert atmosphere, and polymerizing the blend in situ. The
commercial preparation of such products, however, requires a substantial
investment in unconventional manufacturing equipment [92]. Currently low-
voltage electron beam accelerators with high output are used [93]. Acrylics
play a determinant role in this technology too. As stated in [94] most EB
curable PSAs are acrylic functionalized, they are based on (1) postcrosslink-
ing of SIS-HMPSA; (2) acrylate functionalized polyester prepolymers; (3)
acrylate functionalized liquid rubber; or (4) acrylate prepolymers. Ultra-
violet light-induced on-line curing of special prepolymers offers a less
expensive technology for acrylic hot-melts. Such acrylic hot-melts are
prepared by solution polymerization of the monomers along with a small
amount of UV-reactive comonomer [95,96]. Under the action of the UV
light the photoreactive groups crosslink the acrylic polymer backbone
molecules by a chemical grafting (see Section 1.3 also).

Acryl-Vinyl Copolymers. The main competitors to acrylic monomers


are the acrylic copolymers; they are macromolecular compounds with a rela-
tively high (more than 10% wt/wt) content of a second, main comonomer
such as vinyl acetate (AC/VAc) or styrene (AC/S).
182 Chapter 5

Vinyl-Based Viscoelastomers
Among the nonacrylic-based copolymers, one needs to mention the vinyl
acetate and ethylene copolymers and the vinyl ethers (VE). Ethylene
copolymers are elastomers with a second, main comonomer such as vinyl
acetate (EVAc), vinyl acetate and acrylics (EVAc AC), or maleic acid
derivatives. In order to produce vinyl acetate copolymers maleic acid
derivatives are usually copolymerized with vinyl acetate, in order to lower
the Tg of the poly(vinyl acetate/maleinate). Water-borne adhesives based on
polyvinyl acetate were discussed in detail [97].

Vinyl Acetate Copolymers. Polyvinyl acetate homopolymers have


been used as adhesive dispersions since 1940 [98]. Later they were
plasticized using alkyl maleinates and acrylates as comonomers. Vinyl
acetate copolymers display excellent versatility for carrier materials and
adhesive or nonadhesive coatings. Unfortunately, such copolymers are not
suitable for high performance pressure-sensitive use. The first copolymers
with ethylene were synthesized in 1960. From the range of vinyl acetate
copolymers usable for pressure-sensitive adhesives and pressure-sensitive
products, ethylene-vinyl acetate (binary or multicomponent) copolymers are
the most important products. These are used as bulk materials (for hot-
melts) or as dispersions. Functionalized ethylene-vinyl acetate copolymers
have been synthesized also. Special products include maleic acid as
comonomer. Ethylene copolymers are described in the next subsection.
Vinyl acetate-vinyl pyrrolidone copolymers are used mainly in water
dispersible or soluble formulations as base elastomers [57].

Polyvinyl Ethers. Polyvinyl-, ethyl-, and isobutyl ethers were used


some years ago as a main component of PSAs [99]. They are soluble
in a broad range of solvents (e.g., ethanol, acetone, butyl acetate, toluene,
and ethylacetate); methyl ether also is soluble in water. This polymer has
a good adhesion on polyethylene, is not sensitive towards atmospheric
humidity, and therefore has been used on a large scale for roll stock
materials. Because of their resistance to plasticizers, polyvinyl ethers (PVEs)
were proposed as a tackifier for acrylic-based formulations for PVC [100].
The water-soluble copolymer of methyl vinyl ethers and maleic anhydride is
used as a thickener [101]. The most common polyvinyl ethers are based on
vinyl methyl ethers, vinyl ethyl ethers, and vinyl isobutyl ethers [102]. The
polymers are supplied either solvent free, as solutions, or as dispersions; the
vinyl ether polymers may also be coated in the molten state [102].
Water vapor transmission of PVEs has the same value as for human
skin [100]; therefore PVEs are recommended for medical tapes. Because of
Chemical Composition of PSAs 183

their water solubility PVEs are used for water-removable labels and as well
as on humid surfaces [99]. Because of their compatibility with plasticizers
PVEs are migration resistant and are recommended for PSAs coated on soft
PVC. Polyvinyl ethers may also be used as tackifiers for acrylic dispersions,
and polyurethane- and electron beam-cured PSAs; thus, polyvinyl ethers
are used as tackifiers in crosslinked PUR-based PSAs and for electron
beam-cured acrylic PSAs [103–105]. Polyvinyl ethers are mainly used in
water-based PSAs in order to impart a hydrophilic character or as a tackifier
[102]. Unfortunately they are storage sensitive and have to be stabilized
using protective agents. Polyvinyl ethers also are used as adhesives applied
on PE [99,105].

Ethylene-Based Viscoelastomers
In order to use ethylene polymers as raw materials for PSAs, their
plastomeric character must be changed into an elastomeric one, and they
must display a good compatibility with viscous components to achieve the
unique viscoelastic behavior of PSAs. Through copolymerization branched
structures are introduced into the ethylene linear backbone, thus it is made
less compact. A level of 20% vinyl acetate makes it is PE practically
crystalline, a level of 40% produces a fully amorphous material. Above 30%
vinyl acetate, EVAc copolymers behave like elastomers [106]. Recognizing
the potential advantages of EVAc for use in PSAs, a lot of research and
development work was carried out in order to come up with practical,
commercial systems. Ethylene-vinyl acetate (EVAc) copolymers have been
available for more than 25 years; their wide use in non-pressure-sensitive
bonding applications is based to a large extent on improvements over other
adhesives in those areas where traditional PSAs were deficient. In addition
EVAc feed stocks were relatively low in cost and readily available [107].
According to [108] the ethylene content of an ordinary EVAc latex is
about 15%; such a dispersion is manufactured by emulsion polymerization
at medium pressure (4 MPa). The most important methods used for
producing ethylene-vinyl acetate latex with high ethylene content and long
sequence ethylene-ethylene structure include: emulsion polymerization at
high pressure (20 MPa); dissolution of an ordinary EVAc polymer in an
organic solvent which is added to water containing emulsifier at high stirring
rate, and then evaporation of the solvent; direct emulsification of an
ethylene-vinyl acetate copolymer at high speed and high pressure; and
dissolution of an EVAc copolymer in a monomer mixture of VAc
and acrylate.
In ethylene-vinyl acetate copolymers ethylene is the softening
component; its softening effect is stronger than that of dibutyl maleinate.
184 Chapter 5

For an equivalent softness of EVAc copolymer films a level of 43% dibutyl


maleinate, 45% butyl acrylate, 73% ethyl acrylate, or only 25% ethylene is
necessary. As little as 18% ethylene content decreases the Tg of polyvinyl
acetate down to 0 C. For the same effect 20% vinyl maleate, 30–32%
dibutyl maleinate, or 33–34% butyl acrylate are required [109]. The
development of soft, tacky PSAs, starting from the hard vinyl acetate
homopolymer is illustrated in Fig. 5.1.
Dupont introduced the first commercial products under the Elvax
trademark in the early 1960s for use in hot-melt PSAs. These proved to
exhibit excellent properties compared to the polyethylene-based hot-melt
PSAs and the aqueous adhesives such as dextrine and polyvinyl acetates.
Water-borne EVAc for laminating adhesives were patented and introduced
by Air Products in 1973. They quickly became the standard for
many applications, particularly as replacements for externally plasticized
polyvinyl acetate emulsions.
Polymerization of EVAc-containing polymers with intrinsic pressure-
sensitive properties is a more basic approach. Ethylene-vinyl acetate
chemistry requires specialized, high pressure polymerization equipment
and most of the work was carried out at companies having prior expertise.

Figure 5.1 Transition from hard, non-tacky to very soft, tacky PSA.
Chemical Composition of PSAs 185

Wacker Chemie GmbH was awarded the first U.S. patent [110]. Wacker’s
claims covered a copolymer blend of the following composition:

Monomer % by weight

Ethylene 10.0–30.0
Acrylic ester 29.0–69.0
Vinyl acetate 20.0–55.0
Methacrylamide 0.2–8.0
Others 0–12.0

The Tg range is specified at 20 to 60 C. The National Starch


and Chemical Corporation is the other current holder of U.S. patents
for EVAc-based interpolymer PSAs. A different approach was taken at
National Starch; drawing on extensive EVAc experience in the packaging
field, it was decided to maximize the ethylene content in the copolymer.
A high EVAc content provides optimum specific adhesion and plasticizer
tolerance when the adhesives were used on flexible PVC. While the original
National Starch EVAc materials retained their peel value when aged on
plasticized PVC face stock material, the cohesive strength was low. Develop-
ment work resulted in a new U.S. patent [56], covering the following
composition:

Monomer % by weight

Vinyl ester of alkenoic acid 30–70


Ethylene 10–30
Di-2-ethyl hexyl maleate
or di-n-octyl maleate
or the corresponding fumarates 20–40
Monocarboxylic acid 1–10

Wacker Chemie GmbH continued developmental work and high


solids vinyl acetate ethylene dispersions (with 60% solids) were synthesized
[111]. The tackifying influence of maleic acid and maleic acid derivatives on
vinyl acetate adhesives was recognized since the early development of vinyl
acetate copolymers [112].
The first trials of incorporating polar monomers into ethylene (or
ethylene vinyl acetate) copolymers were carried out with acrylates.
Unfortunately, faster reacting acrylates (relative to ethylene or vinyl
acetate) have a tendency towards block copolymerization with vinyl acetate.
186 Chapter 5

The disparity in the reactivity ratios between ethylene and acrylates with
respect to vinyl acetate also limits ethylene incorporation into a terpolymer
system. Maleates (or fumarates), on the other hand, do not readily
homopolymerize, leading to alternating copolymers with vinyl acetate; this,
coupled with a much more favorable reactivity towards vinyl acetate and
ethylene, enables the production of terpolymers with an increased ethylene
content. The two preferred comonomers among EVAc terpolymers for
optimal pressure-sensitive performance are dioctyl maleate and 2-ethyl
hexyl acrylate. The higher molecular weight of dioctyl maleate allows a
lower mole fraction, permitting a higher ethylene content. Thus polymers
commercialized by National Starch contain 20% or more ethylene by
weight. Acrylic and ethylene copolymers may be incorporated into aqueous
ethyl hexyl acrylate-based PSA systems [113].
Water-borne EVAc multipolymers, resulting from high pressure
polymerization of ethylene and vinyl acetate with additional monomers,
constitute a new class of PSAs. Their properties combine the most desired
features of acrylic copolymers and tackified rubbers: high quick stick,
resistance to oxidation and discoloration, clarity, adhesion to low energy
surfaces, and compatibility with plasticized PVC material [114]. A
comprehensive treatment by Benedek [22] covered attempts to utilize
EVAc dispersions as basis for formulating PSAs. Two decades ago Benedek
compared tackified EVAc- and SBR-based PSAs with acrylates and
concluded that, with the exception of shear strength, the adhesive properties
of EVAc and SBR PSA dispersions are inferior to those of acrylate
dispersions. This statement is still valid.
Acrylics exhibit lower peel adhesion and tack than ethylene-vinyl
acetate dioctyl maleate copolymers. Tackified SBR, although exhibiting the
tack of these polymers, generally displays lower initial peel strength [115].
Generally the monomers are polymerized in an aqueous medium under
pressure not exceeding 100 MPa. The quantity of ethylene entering into the
polymer is influenced by the pressure, agitation, and viscosity of the
polymerization medium. Thus in order to increase the ethylene content of
the copolymer, higher pressures are to be employed. A pressure of at least
10 MPa is common [115].
As mentioned earlier, ethylene-vinyl acetate copolymers were used
since the beginning of hot-melt adhesive technology as raw materials for
hot-melt adhesives. The positive influence of the vinyl acetate comonomer
units on the sealability of the ethylene-based plastics is well known for those
familiar with the bonding technology of polymers. Ethylene-vinyl acetate as
a semicrystalline polymer imparts internal strength, film-forming charac-
teristics, and flexibility to hot-melt systems [116]. Unfortunately EVAc
copolymers do not possess the agressive tack required for hot-melt PSAs.
Chemical Composition of PSAs 187

In the past decade some development was achieved through the synthesis
of segmental EVAc copolymers, containing sequences with high vinyl
acetate content and sequences with low vinyl acetate content [117]. These
hot-melt PSAs possess the well-known UV and oxidation resistance of the
ethylene-vinyl acetate copolymers.
As discussed in detail in [118] ethylene-vinyl acetate copolymers have
been suggested as detackifiers, and antiblocking and release agents also.

1.3 Viscous Components


Rubbers are elastic materials, exhibiting a pronounced self-recovering
character. As PSAs are viscoelastic formulations, a change of the elastic
character is possible by decreasing the plateau modulus.
A decrease of the plateau modulus is produced by the use of low
modulus and/or low-Tg formulating additives. The latter are viscous mate-
rials, providing a pronounced viscoelastic character to the rubber/additive
blend. Their use as a reducing agent of the modulus is well known from PVC
chemistry where plasticizers are used to transform hard and brittle materials
into soft, elastic ones. Later on they were used to soften the polyvinyl
acetate-based adhesives.

Plasticizers
The addition of a plasticizer has an effect similar to that of a tackifying
resin. Plasticizers improve the ability of PSAs to flow under bond formation
conditions, but reduce the higher frequency modulus, thereby improving
tack and peel. A detailed discussion of the plasticizers used in PSAs
formulations is given in Chap. 8.

Tackifiers
Pressure-sensitive adhesives are usually blends of rubbers with low
molecular weight resins. The resin is described as a tackifier if added
to the rubber, as the tough dry rubber is converted into a product with
PSA properties. As stated in [119] tackifier resins are compounds having
a low or medium molecular weight, used to change and control the rheology
of base elastomers and viscoelastomers in pressure-sensitive formulations.
According to the best known theory of tackification the resin is a solvent
of the polymer; in tackified blends multiphase structure exists. However,
there are reactive resins also which can chemically modify the base
elastomers or viscoelastomers [120]. The molecular weight and MWD,
chemical nature, polarity, and color of the resins play an important role
in their use. Different natural or synthetic resins, e.g., rosin derivatives,
188 Chapter 5

hydrocarbon resins, coumarone indene, polyterpene, terpene-phenol, and


phenolic and ketone resins are used as tackifiers. Tackifier resins were
described in a detailed manner by Rich [121]. Reactive resins and hybrid
resins are employed too. Reactive resins are functionalized common resins
or special synthetic products with reactive groups, hybrid resins are mixtures
of chemically different tackifiers. Colophonium derivatives, acid resins,
resinates, and resin esters, and their chemistry, properties, applications, and
suppliers were discussed by Jordan [122,123]. The properties of hydrocarbon
resins are covered by Jordan [124,125] and the use of resins in the rubber
industry was reviewed by Fries [126] who describes the structure, properties,
and use of different resins (hydrocarbon resins, rosins, and phenol-
formaldehyde resins) in a comparative manner. A short review of tackifiers
is given in [127].
Among natural resins the rosin derivatives are the most widely used
ones. Rosin is a thermoplastic acid resin obtained from pine trees. The rosin
group consists of rosin, modified rosin, and derivatives. The most widely
used rosin acids include abietic acid, neoabietic acid, primaric acid
(neoprimaric acid), dehydroabietic acid, dihydroabietic acid, and tetra-
hydroabietic acid [126,128]. The storage stability and aging stability of the
resins may be improved by hydrogenation, disproportionation, dimeriza-
tion, and esterification [128]. There are different methods to extract the
rosin. Gum rosin is obtained as oleoresin from the living tree. Wood rosin
originates from aged pine, whereas tall oil rosin is obtained from tall oil
which is a byproduct of the paper industry. Different kinds of resins (e.g.,
disproportionated tall oil rosins, polymerized tall oil rosins, cured rosins,
hydrogenated wood rosins, and polymerized wood rosins) were suggested
[129]. The dimerized rosin acid content (35–60%) may yield harder or softer
colophonium resins [130]. The range of resins used in formulating is
illustrated in Table 5.1. Pentaerythritol esters of hydrogenated rosins
with a ring-and-ball softening point of 104 C and an acid number of 12
are also used.
Hydrocarbon resins are by far the preferred tackifiers in PSA
formulations, representing about 80% of the market. Hydrocarbon-based
resins are low molecular weight polymers based on petroleum, pure
monomers, natural terpenes, or coal derivatives [128]. The most important
raw materials are aromatic C9–Cn fractions, C4–C5 diolefins, styrene, -
methyl styrene, vinyl toluene, dicyclopentadiene, alpha and beta pinene, d-
limonene, isoprene, and piperylene. Generally, hydrocarbon resins are
classified as follows: aliphatic, aromatic, alkyl aromatic, and hydrogenated
hydrocarbon resins. Low color, inert, low molecular weight, thermoplastic
hydrocarbon resins produced from the terpene monomer alpha pinene
with a ring-and-ball softening point of 112–119 C were suggested also.
Chemical Composition of PSAs 189

Table 5.1 The Range of Resins used for Formulation Purposes

Softening point ( C)

Resin type Trade name Drop R&B

Gum rosin — — 78
Wood rosin — 81 73
Tall oil rosin — — 80
Polymerized rosin Poly-Pale 102 95
Hydrogenated rosin Staybelite 75 68
Pentaerythritol wood rosin Pentalyn A 111 —
Glycerine-hydrogenated rosin Staybelite Ester 10, — —
Dermulsene DEG 83 —
Pentaerythritol-hydrogenated rosin Pentalyn H 104 —
Glycerine-highly stabilized rosin Foral 85 82 —
Pentaerythritol-highly stabilized rosin Foral 105 104 —
Hydroabietyl phthalate Cellolyn 21 63 —
Olefin Hydrolin — 100–135
Cycloaliphatic hydrogenated olefin Permalyn — 85–135
Aliphatic petroleum hydrocarbon Piccopale/Piccotac — 70–122
Modified aromatic hydrocarbon Hercotac AD — 85–115
Dicyclopentadiene Piccodiene — 73–140
Mixed olefin Statac, Super Statac, — 75–105
Wingtack — —
Alpha pinene Piccolite — 10–135
Beta pinene Croturez, Piccolites — 10–135
Terpene Zonarez, Nirez 100 — 10–145
Alpha-methyl styrene-vinyl toluene Piccotex — 75–120
Alpha-methyl styrene Kristalex — 25–120
Styrene Piccolastic — 5–190
Terpene phenolic Piccofyn — 100–135
Dermulsene DT 75 — —
Coumarone-indene Cumar — 100

In some cases special chemically reactive synthetic resins such as phenol-


formaldehyde or melamine condensation products can be used. A detailed
discussion of resins will be given in Chap. 8.

1.4 Components for In-line Synthesis


The manufacture of the adhesive can be carried out off-line or in-line. Off-
line manufacturing includes mixing of the adhesive components followed by
190 Chapter 5

coating of the ready to use PSA. In-line manufacture of the PSA consists of
simultaneous coating and curing or postpolymerization, i.e., it is an in situ
manufacture of the adhesive. In this case a special ‘‘ready to coat ’’ mixture
of polymerizable or crosslinkable monomers, oligomers, or polymers (e.g.,
radiation cured hot-melts) is first applied onto a carrier. Such a reaction
mixture is transformed after coating into a ready to use adhesive. The
postcoating synthesis of the PSA must be carried out by the converter. In
some cases the in-line transformation of the coated layer concerns only
the modification of an adhesive, in order to improve its performance (e.g.,
crosslinking of solvent-based acrylates used for protective films, in order to
achieve removability). Therefore the raw materials used for in-line synthesis
include the whole range of components used for the manufacture of pressure-
sensitive adhesives (e.g., monomers, oligomers) and for their formulation
(e.g., elastomers, viscoelastomers, viscous components, and additives) and
other special components (e.g., initiators, crosslinking agents, etc.). Such
a formulation recipe depends on the chemical/physical technology used
for the adhesive manufacture. As discussed in [131] the full or partial
postmanufacture of the PSA is the result of the chemical development
induced by the trend toward solvent-free fabrication. Such formulations
with 100% solids include hot-melts and radiation-cured reaction mixtures.
Technological reasons and insufficient progress in acrylic-based hot-melts
forced the development of oligomer- and macromer-based curable
formulations. The main procedures for in-line synthesis include crosslinking
and polymerization. They can be combined also. The polymeric raw
materials investigated for postpolymerization are PSAs, PSAs with
unbalanced adhesive properties, or nonadhesive oligomers. Monomers
(as main components or as additives) can be used also. It should be noted
that in-line synthesis (modification) of the adhesive supposes functionality.

Curable Raw Materials


Monomers, oligomers, and prepolymers can be polymerized, and constitute
together with polymers the range of curable raw materials.
As discussed earlier tack, peel, and shear resistance of PSAs depend
on the cohesive strength. However, enhancing the cohesive strength, greatly
decreases the chain mobility needed to form a bond and hence the measured
tack decreases. On the other hand, natural rubber can be processed to a
relatively low molecular weight but, upon straining, it crystallizes, and hence
resists separation. Induced strengthening mechanisms are favorable if a high
tack is desired. Thus for high tack, the elastomers used as PSA components
should have a low cohesion at low strain rates to facilitate bond formation,
but a high cohesive strength at high strains to resist bond breakage.
Chemical Composition of PSAs 191

The classical way to obtain a higher cohesive strength is to use


crosslinkable base materials. Thus a viscoelastomer can be transformed into
an elastomer. For instance viscoelastic polyisobutylene was transformed
into an elastomer by crosslinking. This approach was initiated by the
formulation of the old rubber/resin-based PSAs, where natural rubber with
an inherent gel content was used as the base elastomer. Later on chemically
crosslinked or crosslinkable synthetic rubbers containing reactive functional
groups (e.g., CSBR) were prepared. One other possibility was given by the
synthesis of sequenced, physically crosslinked block copolymers (i.e., SIS
and SBS). Acrylics are actually the most suitable PSAs raw materials on the
market, but also the most expensive ones. In this context ease of tackifica-
tion of the HC rubber-based materials should be emphasized. First, chemical
crosslinking based on built-in reactivity was practiced. Later radiation
curing was developed. Next a short description of the most important
representatives of the chemically crosslinked viscoelastomers and elastomers
is presented. The advantages of crosslinking comprise increases of the cohesive
strength, the temperature, the chemical resistance, and the anchorage.

Chemical Crosslinking. The crosslinking of acrylic-based copolymers


(used for adhesives) has been discussed in many industrial papers [132].
Other crosslinked compounds (e.g., polyurethanes) also may be used as base
material for PSAs [133]. The ability of acrylic monomers to polymerize
with each other leads to many applications. These polymers can crosslink
with epoxy resins, amines, reactive resins, as well as alkyl urea derivatives
[134]. Although the suitable polymers can contain functional, crosslinking
monomers, such as N-methylol acrylamide, such monomers may release
formaldehyde upon curing or cause the loss of tack and adhesion. Thus the
preferred polymers contain less than about 1% of N-methylol acrylamide
monomer units. According to [65] in an acrylic emulsion 0.5–1.0 wt% IBMA
(isobutoxy methylacrylamide) was copolymerized to provide interlinking of
the microgels in the film upon heating.
In some polymer systems the unsaturated carboxylic acid is the
effective ingredient for modifying polymer properties (e.g., cohesive
strength); these modifications can be controlled through crosslinking with
a suitable agent such as chromium acetate. Multifunctional comonomers
containing ethylene units, such as dialkyl maleate, trialkyl cyanurate,
tetraethylene glycol dimethacrylate, etc., can be crosslinked by eliminating
the unsaturation. The crosslinking of polyvinyl acetate dispersions using
metal salts was discussed by Homanner [135]. Pressure-sensitive adhesive-
like networks were prepared using the following starting compositions: 90%
EHA and 10% 1,6-hexanediol diacrylate (HDDA); 80% EHA, 10% AA,
and 10% HDDA [73]. Such polymers possess a crosslink density similar to
192 Chapter 5

commercial acrylic PSA. The mechanism of crosslinking for solvent-based


acrylics was studied by Milker and Czech [136]. They used formulations
with hydroxyethyl acrylate and acrylic acid as reactive monomers, and
titanates and acetyl acetonates (Al, Co, Ti, Zn, Zr) as crosslinking agents.
Solvent-based acrylates containing 5% wt of hydroxyethylmethacrylate
(HEMA) were crosslinked with PUR [137]. As an example of crosslinking
for a solvent-based acrylic PSA the following procedure is described:
1. An organic solution of a PSA, a hindered phenol antioxidant, and
a tackifying rosin ester are blended.
2. An organic solution of N,N0 -bis-l,2-propenylisoftalimide cross-
linker is added to the blend.
3. A thin layer of the adhesive solution is coated onto a sheet
backing.
4. The coated sheet is heated to remove the solvent and to cure the
polymer and crosslinker [138].
In another example an acrylic copolymer with N-methylol acrylamide as
the reactive comonomer was cured using ethyl acetoacetate diisopropoxy-
aluminum [139].
Crosslinking may be carried out using metal complexes, isocyanates,
amino resins, polyamide, polyamine, epichlorhydrine resins, and other
compounds [136]. Metal acetyl acetonates and orthotitanic acid esters are
used together with alcohols as stabilizing agents. Crosslinking can only
be obtained provided that the polymer chain is equipped with acid groups.
The hydroxyl group alone does not contribute to crosslinking; however, in
combination with COOH groups, an unexpected synergism reflected in a
higher shearing strength follows. The application of isocyanates remains
limited in practice, since the pot life of PSAs being crosslinked with
polyisocyanates is rather short. The special feature of amino resins used as
a crosslinker is the low crosslinking speed at room temperature.
A new class of polyaziridine derivatives giving a low crosslinking
density has been developed. According to their chemical composition and to
the nature of built-in reactive sites, quite different crosslinking agents may
be used. For general use polyaziridines were proposed as crosslinking
agents [140]. An organic solvent solution of N,N0 -bis-l,2-propenyl-
isoftalimide may be used as a crosslinker. A chrome stearic acid complex
(0.1 wt% of the dry adhesive) also was suggested [141]. Crosslinking
via aluminum acetate/acrylic acid reactions was suggested for ethylene/
maleinate copolymers [115]. Curing of silicone-based PSAs is done using
benzoyl peroxide.
The additive trialkyl cyanurate is recommended for the electron beam
crosslinking of polymers. This component allows a controlled adjustment of
Chemical Composition of PSAs 193

crosslinking density and improves mechanical characteristics. Macromole-


cular compounds like polystyryl ethyl methacrylate also may be used.
The chemical crosslinking of rubber derivatives was developed also.
Water-soluble polyamide-epichlorohydrine-type materials are effective as
crosslinking agents for CSBR latexes [142]. Although many SBR latexes
are self-crosslinking, special crosslinking agents can be used. When modifying
by crosslinking it is common to add a solution of some suitable zinc
compounds. A careful balance must, however, be considered as the addition
results in decreased tack. It should be noted that addition of zinc oxide to
carboxylated neoprene latex dramatically reduces the pressure-sensitive
character of the adhesive [143]. A crosslinkable PSA composition was
formulated on the basis of a butadiene acrylated rubber; it is crosslinked
with aluminum propoxide or sec-butoxide [144].

Radiation Curing. Generally polymerization by radiation is similar to


the conventional process of thermal (free radical) polymerization initiated
by peroxides or azo compounds. Different kinds of high energy beams (IR,
UV, or electron beams) may be used as a source of energy. The crosslinking
of UV light-curable PSA formulations can be initiated by laser beam, too
[145]. Similarities between UV- and EB-curing exist also, but each has its
special features. In the UV curable formulation free radicals are generated
from the photoinitiator, in the EB curable formulation the free radicals are
generated on the reactive functional sites upon exposure to an EB source.
The electron beam is a much higher energy source than UV light and
promotes a higher level of crosslinking, therefore UV light-curable
formulations cannot be used unchanged for EB curing. The EB-induced
polymerization supposes the reaction of a vinyl unsaturation with electrons
of a suitable energy level. Because of the lower energy of UV radiation
generally a photoinitiator is needed, but initiatorless UV light-curable
systems were developed also [146]. Such technologies cover various domains
of the converting. The use of electron beam-curing in the paint industry was
discussed in [147,148]; the development of silicone release manufacturing via
electron beam-curing was described by Pilar [149]. Electron beam-curing
may also be used for lacquers and laminating adhesives [150–152].
Ultraviolet light-induced curing of laquers and printing inks was tested
industrially since 1973 [153,154]. Water-based electron beam-cured coating
systems were discussed by Loutz [155]. As discussed in detail in [156]
pressure-sensitive adhesives crosslinked by ultraviolet light or electron
beams have been available for 15 years or more, but their application has
not been a total success. Radiation-cured tapes were patented in 1954 [157].
Such tapes were manufactured first with a solvent-based pressure-sensitive
adhesive composition, dried, and cured with an electron beam [158].
194 Chapter 5

Radiation curing of PVC carrier films for tapes was introduced in the 1980s
[159]. Proposals and guidelines for the formulation of beam-curable PSAs
were given in [160]. The first rubber-based adhesives were beam-cured. In
Europe a plant at Tesa-Werk Offenburg (Beiersdorf) for the curing of tapes
was started [161] where classical adhesives are being used. The methods and
possibilities of electron beam-curing for continuous webs were also discussed
by Holl [147]. A comparative study of the costs for electron beam and
water-based coating is given by Pagendarm [162]. Actually silicone release
curing by radiation and curing of printed coatings is a common technology.
As prognosticated by [163], in 2000 about 70–80% of label printing
machines used UV light-induced curing. Hybrid systems for combined UV
light/electron beam or hot air/electron beam were introduced for offset
and flexo printing [164]. Progress in UV curable acrylic hot-melt precursors
promises a larger use of this technology for PSAs.
Generally, formulated adhesives are cured after their coating, i.e., the
curability of the main elastomer or viscoelastomers is strongly influenced by
the performances of the other formulating components. Ewing and Erickson
[36] studied the response of various resins and stabilizers to electron beam
radiation. Their results show that proper selection of the formulating
ingredients can minimize the required radiation dosage for the adhesive,
while maintaining the desired balance of tack, shear, and solvent resistance. A
detailed discussion of the formulation used in radiation curable PSAs is given
in Chap. 8. Next, a short description of the main components is given.

Polymerizable Precursors
As discussed earlier, crosslinking allows the (build-up or) postmodification
of the elastic network necessary to regulate the adhesive-cohesive balance
of the PSA. It supposes the existence of crosslinkable polymeric structures.
Thus segregated constructions are made. Another possibility to create such
composites is given by polymerization. This procedure uses monomers to
synthesize polymers (viscoelastomers) or polymeric and polymerizable (or
curable) compounds (macromers) which are built into an elastical network.
For this process the converter can choose to use either a classic
polymerization technology (i.e., thermal, free radical initiated polymeriza-
tion, polyaddition, or polycondensation) or a radiation-induced reaction. It
is evident, that in some cases this choice is limited by the chemical nature of
the (pre-) polymer. It is to be emphasized that in certain cases (i.e., acrylic
hot-melt development), postpolymerization of an oligomer is a necessity
because of the lack of suitable macromolecular compounds. According to
[165] the main possibilities to formulate radiation curable formulations
include the syrup approach (oligomer/monomer mixture), the reactive
Chemical Composition of PSAs 195

oligomer approach, and crosslinking of pressure-sensitive adhesives.


For monomer- and oligomer-based radiation curable formulations poly-
merization and postcrosslinking may occur simultaneously.
Monomers for In-line Synthesis. Generally, special monomers are
used together with polymeric compounds for in-line synthesis of the
adhesive. In principle the liquid monomers are applied to the carrier, which
is an endless belt, and the polymerization is completed thermally or in an
oven by UV light. Electron beam-cured formulations are composed of
oligomers, prepolymers (50–100%), multifunctional monomers, and addi-
tives. The most important oligomers used are polyester acrylics, epoxy
acrylates, urethane acrylates, and polyether acrylates. Hexanediol diacry-
lates, trimethyl allylpropane triacrylates, tripropylene glycoltriacrylates, and
pentaerythritol tetra-acrylate are used as monomers (called reactive diluting
agents). The term reactive diluent generally refers to an unsaturated ethylene
monomer which is miscible with the principal oligomers, which reduces the
viscosity of the composition, and which reacts with the oligomer to form
a copolymer. In some cases the addition of the reactive diluent not only
reduces the viscosity of the uncured diluent composition but also increase
the elongation of the cured coating [166]. Ethyl acrylate, butyl acrylate,
glycidylmethyl acrylate, and acrylic acid may be used as monomers for
prepolymers [167]. In some cases vinyl lactone is built-in as a sensitive
comonomer [168]. The reactive diluents used to reduce the viscosity usually
have hydroxyl groups as active sites. An example is hydroxylated
caprolactam acrylate [169]. Copolymerizable photoinitiators have been
studied for the curing of functionalized acrylates also. Such monomers may
contain vinyl unsaturation, diol, or epoxide groups [170,171]. Benzocyclo-
butenone acrylamide monomer has been incorporated into polymers also.
Under UV radiation benzocyclobutenone (BCPO) readily reacts with
itself and also with an alcohol to produce an ester. Thus UV light-induced
crosslinking occurs without an initiator. Solvent-based adhesives with
polymers containing 2-hydroxyethylacrylate and BCPO were prepared by
solution polymerization and they were mixed 1/1, giving upon curing
enhanced adhesion. Postinduced covalent crosslinking is superior to ionic
reactions and proceeds without the need of any initiator [172].
Oligomers for In-Line Synthesis. Oligomers were developed for in
situ manufacture of pressure-sensitive adhesives. Such oligomers (50–10%
of the recipe) can possess various chemical compositions (hydrocarbons-
or heteroatom-based) and different functionalities. Both operations—the
preparation of the prepolymer and its postpolymerization—can be effected
by classical (polymerization, polyaddition, or polycondensation) or
radiation-induced polymerization. Possible formulations for UV-cured
196 Chapter 5

PSAs include acrylic oligomers; unfortunately their shear resistance is too


low [173]. Liquid, saturated copolyesters with one terminal acrylic double
bond per 3000–6000 molecular weight units are used in formulations for
UV- and electron beam-curable PSAs [174]. Acrylated polyester oligomers
(with a molecular weight of 3000–8000) were tested as hot-melt PSAs [155].
According to Harder [175] a HMPSA on acrylic basis has been
prepared by solution polymerization. Low molecular weight, low viscosity
polymers have been synthesized which can be coated by 90–140 C and
postcrosslinked, using EB or UV curing. For instance, Acronal DS 3429 is a
UV-curable acrylate copolymer for HMPSAs [176]. It is a highly viscous
clear liquid, which can be processed at 100–140 C. It possesses a viscosity of
about 15,000 mPa.s at 120 C. The photoinitiator is built into the polymer;
no C ¼ C double bonds exist in the raw material. Therefore this product is
not suitable for EB curing.
As discussed earlier (Section 1.1), hydrogenated styrene block copoly-
mers have excellent light and heat stability but display high viscosities and
cannot be crosslinked. Therefore new, low viscosity, curable raw materials
were synthesized. Ultraviolet light-curable reactive oligomers were manu-
factured. They are low molecular weight products having special functional
groups which participate in UV-light crosslinking (specially by cationic UV-
curing). A poly(ethylene-butylene) rubber bearing a primary hydroxyl
functionality on one end and epoxidized polyisoprene functionalities on the
other end, with a Tg of 53 C was developed (Kraton EKP-200) by Shell. The
grade L-1203 is a linear poly(ethylene-butylene) rubber bearing a terminal
aliphatic primary hydroxyl group on the same backbone and possesses a
Tg of 63 C. Another macromer possesses a terminal hydroxyl functionality
on both ends. Generally a blend of such products having different func-
tionalities is used. Other special oligomers are described in detail in [33].

Polymers for In-Line Synthesis. The most important requirements for


the polymers used as the electron beam-curing material for PSAs are a
low crosslinking density and a Tg below 20 C [116]. Ultraviolet light-cured
hybrid systems were discussed by Barisonek and Froelic [178]. Storage-
stable latently curable acrylic formulations based on triethylene glycol dimeth-
acrylate and polymeric hydroperoxides were patented [178]. Generally the
initiator system for UV-cured PSAs may be free radical or cationic [179].
Crosslinking via polymer bonded benzophenone groups has been studied
also. There is a correlation between the activity and structure of polymeric
photoinitiators containing side-chain benzophenone or chromophores.
Such built-in photosensitizers have been tried for styrene-butadiene
copolymers also.
Chemical Composition of PSAs 197

Photoinitiators. By photoinitiated polymerization the light energy


produces reactive species (radicals, cations, or anions) which can initiate
thermal chain reactions or short life intermediates (e.g., radicals, carbenes,
or nitrenes) which can produce crosslinking reactions. Because of the lower
energy level of UV radiation a photoinitiator is needed. Benzophenone,
thioxanthone, benzoin ether, benzylketals, -acyloxime esters, and dialkoxy-
acetophenone are the most used photoinitiators. Different photoinitiators
act through different mechanisms. For instance, benzoin derivatives, benzyl-
and acetophenone-ketals, aliphatic azoderivatives, peroxides, aromatic
disulfides, perhalogenides, -benzoinoxime ester and benzoinphosphine
oxides act via intermolecular bond association. Aromatic ketones, especially
benzophenone, thioxanthones, quinones, and heterocyclical compounds act
through internal hydrogen abstraction. They need accelerators (like amines
or alcohols) which easily release hydrogen atoms. Ketones with amines,
allylthiocarbamides, etc., act through photoinitiated electron transfer.

1.5 Components for Special Pressure-Sensitive


Formulations
As discussed earlier (Chap. 3, Section 1.2) in the last decade research in the
field of bioadhesives allowed the development of a special class of products
which display pressure-sensitivity although their components (alone) are not
elastomers or viscoelastomers [180,181]. Such systems can be prepared by
formulation only. They display pressure-sensitivity due to the mutual effect
of plasticizing and crosslinking of the formulation components in the
presence of water. This new class of pressure-sensitive products includes
hydrogels. Hydrogel (containing 4–12% water) based miscible blends of
high molecular weight PVP with a short-chain PEG reveal a pressure-
sensitive character of adhesion within a narrow range of PVP-PEG
compositions [182]. In this case adhesion is a result of hydrogen bonding
of both polyethyleneglycol (PEG) terminal hydroxy groups to the carbonyls
in polyvinylprrolidone repeating units. The PVP-PEG hydrogels behave
like typical rubbers and couple the properties of PSAs and bioadhesives.
According to [182] bioadhesion can be defined as a pressure-sensitive
character of adhesion toward highly hydrated, soft substrates, comprising
mucin (the glycoprotein predominant in the mucous layer). In contrast
to bioadhesives conventional PSAs are mainly hydrophobic elastomers
providing much stronger adhesion to dry substrates. However, they lose
their adhesion as the substrate or adhesive is exposed to bulk water. The
distinctive feature of bioadhesives is a plasticizing effect of water. As stated
in [180] the Vinogradov equation [183] (derived originally for elastomers),
198 Chapter 5

can be applied to describe the debonding mechanics of PVP-PEG adhesive


hydrogels. In the PVP-PEG mixtures pressure-sensitivity arises within a
narrow range of compositions and is also affected by blend hydration. Later
it was demonstrated, that there is a general method to tailor PSAs by mixing
nontacky hydrophilic polymers with short-chain plasticizers, which bear
complementary functional groups at their ends. Although chemical
composition and structure of PVP-PEG hydrogel have nothing to do with
those for conventional PSAs, nevertheless the PVP-PEG hydrogen-bonded
network obeys also Dahlquist’s criterion of tack. There is a general
applicability of the PVP-PEG model to other hydrophilic polymers and
plasticizers. Adhesive hydrogels can be obtained by PEG blending with
such long-chain complementary hydrophilic polymers as other poly(N-vinyl
amides), particularly poly(N-vinyl caprolactam), their copolymers, poly-
acrylic acid etc. The Kovacs equation (Chap. 3, Section 1.2) predicts that a
high adhesion has to appear in the polyethyleneglycol blends with polyvinyl-
caprolactam and other complementary long-chain partners with the same
composition range as it occurs for PVP-PEG blends (35 wt% PVP and
5 wt% PEG). This prediction is also confirmed experimentally [184,185].
PEG is not a unique short-chain telechelic plasticizer and cohesive strength
enhancer. Tacky blends can be obtained by mixing PVP with other hydroxyl
or carboxyl terminated short-chain plasticizers, ethylene glycol and its
polymers, with molecular weight of 200–600 g/mol, low MW propylene
glycol, and alkane diols (propane diol up to hexane diol). The highest peel
strength has been found for PVP with glycerol which possesses the
maximum density of hydroxyl groups per molecule. Strictly speaking
carbonic diacids are not plasticizers, but noncovalent crosslinking agents
acting as cohesive interaction enhancers. Those can be applied in
combination with plasticizers which reduces Tg. Water is an appropriate
plasticizer to impart adhesion to PVP blends with carbonic acids. The key
result is that the free volume and cohesive strength of the polymer need to
be of appropriate magnitude, and in a strictly specified ratio to each other
for adhesion to appear. A similar conclusion was previously derived from
the finding that the maximum in the debonding force corresponds to the
minimum value of the CpTg product. This is demonstrated for PVP-PEG
blends over the range of compositions and for other (common) PSA base
polymers (e.g., PIB, natural rubber, etc.) [186].

1.6 Other Components


The most important single components in a PSA are the polymers
(elastomers and tackifiers). Other micromolecular components such as
Chemical Composition of PSAs 199

plasticizers, antioxidants, fillers, crosslinking agents, initiators, detackifiers,


solubilizers, primers etc., are also included. Depending on the application
technology carriers (solvents or water) may be required. Depending on the
coating technology wetting agents are used; converting and end-use
properties of the label may require special additives. Such compounds can
be divided into chemical and technological additives. A detailed discussion
of these components is given in Chap. 8.

1.7 Release Coatings


The release coating does not constitute a component of the pressure-
sensitive adhesive but it is a part of the pressure-sensitive product, and thus
it influences the performance of the pressure-sensitive adhesive. Therefore a
short description of the release coatings will be given. The energy dissipation
during debonding is of two kinds: a surface dissipation in the vicinity of the
interface due to the separation and a second type of dissipation due to
the bulk deformation of the viscoelastic adhesive and of the release coating.
It is expected that the surface dissipation is mainly controlled by the nature
of the interface, i.e., the composition of the release coating, while the bulk
dissipation would be mainly controlled by the viscoelastic response of the
adhesive. The release coating is very thin relative to the adhesive layer
and, as it is more elastic than the adhesive, the energy dissipation during the
debonding mainly takes place in the adhesive. As stated by Gordon et al.
[187–189] until recent studies the release liner was simply viewed as a low
surface energy platform from which to release organic PSAs. While
empirical correlation appeared to exist between silicone bulk rheological
properties and absolute release magnitudes, it was not yet clear as to
whether this was truly a bulk or interfacial rheological effect. It is evident,
that the viscoelastic properties of the release influence the magnitude of
the release force also. The release mechanism involves a low polarity surface
or low surface tension and incompatibility of the release surface and
adhesive surface. Generally the adhesive raw materials are high polymers
bearing nonpolar or/and voluminous side chains or functional groups. The
adhesive effect of low polarity is well demonstrated by silicones and
fatty acid derivatives. The release surface will normally be provided by
a layer of a silicone polymer or by a material providing similar release
properties coated on the liner substrate (see Chap. 3). Non-silicone release
coatings were developed too. The various release coating technologies
available on the market and their strengths and weaknesses were discussed
by Craig [15].
200 Chapter 5

Silicone-Based Release Coatings


Silicones are the best known dehesive macromolecular compounds. They
can be imbedded into plastomers and elastomers too. As stated in [190] the
most used release coatings with general applicability are based on silicone
polymers. Such products were developed many years ago for greaseproof
and release papers and antiblocking coatings. The use of silicone as a release
coating is based on its low surface energy (22–24 mN/m), where the surface
energy of most adhesives averages 30–50 mN/m. This difference prevents
the adhesive from wetting the silicone surface. The adhesion energy of the
acrylic adhesive on silicone release coating is 10 times smaller than that
on steel [191]. The basic release liner is comprised of paper with a very thin
silicone coating layer, adhering sufficiently to the adhesive to hold the
laminate together, but enabling peeling off of the release paper from the face
material [192]. Silicone release resins are linear polymers or prepolymers in
liquid form, with or without solvents. They are coated onto paper, film, or
other backing materials and then cured typically by heat and/or surface
catalytic action to form solid, nontacky, crosslinked polymers in situ. The
release resins are formed from halosilicones and consist predominantly of
repeated units of the structure:
R
j
O Si 
j
R
where R is a hydrogen or a hydrocarbon radical, usually a lower alkyl
or aryl (typically methyl) group, O is oxygen, and Si is silicon. The degree of
polymerization is such as to produce a liquid linear prepolymer material
with no significant crosslinking. It is believed that the currently known
silicone linear release prepolymers consist of at least 95% repeating units of
this structure with reactive end groups, but that small quantities of other
modifying units may be present, if so desired [192]. For silicone-based PSAs,
crosslinked silicone release liners are suggested [193]. Siloxane grafted vinyl
copolymers may also be used as release layers [194]. Silicone coatings can
be divided into two main groups, namely, thermally cured and radiation-
cured ones.
Thermal curing occurs through condensation or addition reaction.
Thermally cured silicone release systems consist of a reactive polymer, a
crosslinker, and a catalyst. The radiation curing process is based on the
chemical effect of radiation. Moisture curing silicone release coating
technology is only at the developmental stage. More than 20 years of
research were devoted to the concept of silicone addition curing systems
Chemical Composition of PSAs 201

(using a platinum-based catalyst) for paper release applications [195].


However, it was not until the 1980s that this technique really began to
develop after better control over the problem of inhibition had been achieved.
Although more expensive than the condensation system, additive
curing offers several advantages [196]. It allows greater control over release
characteristics. There is no postcure; once the coating is properly cured,
there is little tendency to block. Differential release can be achieved using
different polymers and additives, and, as there is no postreaction, the
release force remains stable with time. Actually 1–2 m thick silicone
release is coated with a speed of >5 m/sec and displays acceptable cure in
<4 sec [187].

Catalyst System. With a tin catalyst system the curing is performed


via a condensation reaction. The coating materials are supplied as two
components which are crosslinked in the coating process; component A is
a silicone polymer and component B contains the tin catalyst. The curing
reaction acts as a postcure; on double-sided coatings, migration may occur
which increases blocking [197].
With the platinum catalyst system the curing occurs via an addition
reaction. The coating materials are supplied as two components which
are crosslinked in the coating process; component A is a vinyl functional
siloxane and acts as the catalyst, and component B contains a silane hydride
functional polymer. For addition cure emulsions a platinum or rhodium
component dispersed in a methyl vinyl siloxane polymer acts as a catalyst.
The oxidation state of the metal, the nature of the binder, and the content of
platinum will determine the level of reactivity of the catalyst and its shelf life
stability. The base polymer consists of one or several dimethylpolysiloxanes
which contain reactive vinyl groups. The portions of these groups, their
number, environment, and accessibility will help to determine the reactivity of
the system and the type of material obtained after cure, which constitutes the
skeleton of the system [15]. For these systems a crosslinker composed of one
or several methyl hydrogenopolysiloxanes may be used. In order to inhibit the
reaction at room temperature stabilizers (e.g., acetylenic alcohol, acetylenic
ketone dialkyl maleates, and azodicarboxylates) may be used [15].

Controlled Release. The main goal of release formulation is to


increase release so as to be able to efficiently ‘‘dial a release.’’ In principle the
silicone release coating is a polydimethylsiloxane (PDMS)-based silicone
network. Its adhesive or nonadhesive performances can be modulated either
by PMDS chemical modification or by adding additive molecules. As
discussed in [187] adhesion theory might lead one to expect that interfacial
absorbtion, molecular interdiffusion, and entanglements are mechanisms
202 Chapter 5

envisioned as a mean of controlling the release force. The industrial solution


to release control is the inclusion of a methylsilicate in a manner analogous
to tackifier resins in the organic adhesive they are in contact with. A tri- or
tetrafunctional resin is used as a control release agent (CRA). The properties
of the silicone release coating are tuned by introducing a siloxane-based
resin to create a differential, which is the ratio of the base polymer release to
a higher release due to high release additives (HRAs) in PDMS. The
additive molecules are called MQ molecules. Adhesion increases with
increasing amount of resin present in the elastomer. How does the resin act?
The role of the methylsilicates functioning as HRAs to modify the
viscoelastic properties of PDMS was highlighted by Gordon et al. in
[198]. It has been shown that the methylsilicate effectively reduces the
segmental mobility of the flexible PDMS backbone. According to [199]
the network density increases (Mc decreases) as the MQ amount increases
in the network. The network is more and more dense. The influence of
the molecular weight is more complex. The network density is higher for
the medium molecular weight, whatever the additive amount, and the Mc
is higher for the higher polymer length. For higher molecular weight of the
base polymer the additive content does not seem to have a great influence,
whereas for a molecular weight equal to 11,000 g/mol the adhesive and
attractive forces vary dramatically when the additive content is equal to
20%. For additive molecule amounts higher than 30% in the polymer blend,
the presence at the interface of MQ molecules induces polarity increase and
stress transfer at the interface. The consequence is an increasing of both the
speed exponent (n) of the viscoelastic function , and of the intrinsic
adhesive strength Go in the Gent and Schultz equations [200] relating the
viscoelastic response to the interfacial response:

Gp ¼ Go  ðT,vÞ ð5:1Þ

Gp ¼ Go Rnp ð5:2Þ

where Gp is the adhesive strength and Rp is the peel speed. Increasing the
molecular weight between crosslinks of the linear PDMS chain in the
network increases the release force. Likewise, increasing the concentration
of the methylsilicate HRA increases the release force. However, the largest
impact of the HRA on the release force is typically at low to intermediate
peel velocities (0.005 m/s to 0.17 m/s) while increasing the molecular weight
between crosslinks mainly influences release at high peel rates [201,202].
Chaudhury and Whitesides [203] proposed a release mechanism for cured
PDMS-based release coatings in which interfacial slippage minimizes the
Chemical Composition of PSAs 203

bulk shear deformation experienced by organic adhesives. Commercially


cured PDMS release coatings exhibit interfacial slippage. Traditional
HRAs used in release coatings ‘‘freeze out’’ interfacial slippage resulting
in increased adhesive deformation. The HRA reduces the segmental
mobility of the PDMS chain within the cured network leading to increased
shear stress and an increase in the practical peel force. At lower MQ content
the PDMS viscous surface layer contributes to the very low adhesion
behavior due to the existence of a surface weak boundary layer. Concerning
the molecular weight effect, it appears that, depending on the network mesh
size, MQ additive molecules are either trapped (low Mc) or not (high Mc)
in the network. Such segregation is important for the modulation. MQs
participate in the crosslinking reactions too. Crosslinking reactivity between
the curing agent and PDMS or MQ molecules depends on the initial PDMS
chain length. As a consequence, for intermediate molecular mass (11 k) a
surface weak boundary layer of free PDMS chain can exist. According to
[191] the proportion of resin in the silicone release coating controls the
propagation velocity of the cracks and therefore the maximum deformation
of the adhesive before debonding.

Radiation-Cured Systems. Ultraviolet- and electron beam-cured


coatings are one-component solventless systems. The process is based on
the chemical effects of radiation; in the UV system it is necessary to use some
heating energy [204]. Ultraviolet-cured silicones have been proposed for
narrow webs; for broad webs EB-curing has been recommended [205].
Although beam-curing (UV and electron beam) silicones are available
on the market this technology has not succeeded yet (apart from a few
exceptions). The main advantages are the low space requirements and
minimum substrate strain compared with thermal systems. The advantages/
disadvantages of beam-curing will be discussed in Section 1.4 of Chap. 5.
Silicone coating technologies may be divided into three major
categories, according to their application form [206]: solvent-based (addition
and condensation products) silicones, emulsions, and solventless (thermal or
electron beam-cured) silicones [204]. In 1986 about 15% of silicone coatings
were carried out with solvent-free silicones [205], while solventless silicones
covered 40% of the whole consumption.
In the United States two electron beam-cure silicone release coating
machines were installed in 1983 [147]. In Europe the first release coating
machine using electron beam-curing was installed in 1984 [207]. Radiation-
curable release coatings may be prepared with a low adhesive material
of limited compatibility in the liquid, so that a thin layer migrates to the
surface of the film [208]. A controlled release effect may be achieved using
a different polydimethylsiloxane level [209]. Acrylic or epoxy derivatives of
204 Chapter 5

silicone may be used as active components. Generally polyfunctional acrylic


or mercapto groups are used as reactive sites on the dimethylpolysiloxane
[210].

Non-Silicone-Based Release Coatings


As discussed in [190] for some products there is no need for the high level
of adhesive performance afforded by silicones. For adhesive coatings with
very low tack and peel the use of silicones would give laminates having
insufficient bond strength. For these applications non-silicone release layers
are suggested. Various compounds such as CMC, alginates, and shellac
have been tested [211]. Natural rubber latex [212], fatty acid derivatives,
melamine, Werner complexes, complexes of basic chromium chloride with
fatty acids, metal complexes of fluorinated derivatives, starch and water-
soluble fluorinated compounds, vinyl esters having long pendant groups
containing tetrafluoroethylene and hexafluoropropylene oligomers, etc.,
were proposed also [213–215]. The usual non-silicone release coatings
for PSAs are based on polyvinyl carbamates, acrylic ester copolymers,
polyamide resins, and octadecyl vinyl ether copolymers. As an example,
acrylonitrile-stearyl acrylate copolymers, chlorinated polyolefins, and
ethylene-vinyl acetate copolymers may be used as solutions [216]. The
aqueous release coatings (which are not silicone-based) presently marketed
are based on either acrylic ester copolymers or vinyl acetate copolymers
and are available in crosslinkable and self-crosslinking types. The self-
crosslinking polymers have found application where heat resistance is
required [217].
From the solvent-based, non-silicone release coatings carbamates have
a broad application spectrum. As discussed in [218] polyvinylcarbamate is
the main release agent used for protective films. It can be applied alone or in
mixtures with chlorinated polyolefins. Kinning [219] has studied the release
behavior of acrylic tapes from release coating based upon long alkyl side-
chain polycarbamates [polyvinyl-N-octadecyl carbamate (PVNODC) and
polyvinyl-N-decylcarbamate (PVNDC)]. The surface energy of carbamate is
the same or lower than that of a silicone [203]. The lower surface energy of
such release coatings is probably due to self-assembly of the side chains
into a liquid crystalline order, thus presenting only methyl groups to the air.
As described by Kinning [219] the use of PSAs containing acrylic acid on
polyvinylcarbamate release coating may lead to blocking. For these
products the interfacial energy is almost the same, but the work of adhesion
(with acrylic adhesives without acrylic acid) differs (it is 40 mJ/m2 for
PVNDC and 30.7 mJ/m2 for PVNODC). Polyvinyl-N-decylcarbamate is
much more likely to provide blocking compared to polyvinyl-N-octadecyl
Chemical Composition of PSAs 205

carbamate. This was ascribed to a melting and loss of structure in the release
coating as the temperature was increase [73].

2 FACTORS INFLUENCING THE CHEMICAL


COMPOSITION

The most important means to adjust the chemical composition of a PSA


are the synthesis of the base materials and the blending or formulation of
the components. However, other factors may influence the chemical compo-
sition of the PSA, such as the physical state of the synthesized elastomer,
the coating technology, and the solid state components of the laminate.
As pointed out earlier a different end-use requires a different rheology;
thus the end-use properties also determine the chemical composition of the
adhesive.

2.1 Synthesis
The synthesis of the raw materials for PSAs provides a straightforward way
to modify the properties of PSAs through the chemical composition. Such
synthesis includes polymerization and polymer analogous reactions.

Polymerization
The polymerization process which leads to the raw materials for pressure-
sensitive adhesives or to the pressure-sensitive adhesive itself is a function
of the polymerization system employed and of the polymerization
conditions. The polymerization system is determined principally by the
monomers and initiators used, which impose the mechanisms of
the polymerization, but other additives may affect the characteristics of
the macromolecular compounds synthesized. Generally the choice of the
monomers and of the polymerization technology ensures an adequate
change of the adhesive properties. For special products and (in-line)
manufacture technology postpolymerization based on polymer analogous
reactions was developed. Therefore, as discussed in detail in [156,220]
formulation has to cover off-line synthesis and in situ synthesis too. In
principle, polymer synthesis can lead to raw materials for PSA (e.g.,
elastomers and viscoelastomers) or to PSAs. The synthesis of polymeric
pressure-sensitive raw materials is generally based on pure monomers
and is carried out as off-line synthesis. The synthesis of pressure-sensitive
adhe-sives based on monomers, oligomers, and polymers may occur in-line
206 Chapter 5

or off-line. The off-line manufacture covers synthesis of reactive base


materials and it may include the partial use of such reactivity.

The Choice of the Monomers. Previously the importance of the Tg


and modulus for PSAs was discussed (see Chap. 3). The first step in the
design of an adhesive system is the selection of the monomers to build-up
the polymer backbone. The comonomers, molecular weight, and degree of
crosslinking or plasticizing of the polymer strongly influence its Tg. Ethylene
is the least expensive monomer usable to decrease the Tg. According to
their functionality the comonomers used for ethylene copolymers can be
classified as follows: softening, hardening, polar, and reactive comonomers
(see Section 1.2 too). The requirement for highly branched soft monomers
(e.g., ethyl hexyl acrylate or butyl acrylate) as the main polymer backbone
was pointed out earlier. On the other hand it may be desirable for various
reasons to incorporate any of several modifying comonomers as part of the
terminally unsaturated vinyl monomeric portions of the adhesive system.
For example, acrylonitrile imparts hardness and solvent resistance, t-butyl
styrene improves tack, methyl methacrylate makes the adhesive harder,
octyl vinyl ether softens the adhesive, etc.
The synthesis of the main polymer may influence the chemical
composition of the adhesive through a built-in reactivity which can be
activated at a later time. Functional groups are usually incorporated into
the polymer for chemical reactivity reasons (e.g., crosslinking). This type of
chemical reaction is desired to minimize the adhesive film’s thermoplastic
response and maximize its tensile properties. As an example, as the polymer
contains carboxyl groups, the bond strength and other properties can be
modified and improved by adding multivalent metal ions in a suitable form
to promote ionic crosslinking. As discussed in [220] the following perfor-
mance characteristics of PSAs are improved by carboxy groups: adhesion,
cohesion, crosslinking, and dispersion stability. Carboxy functionality is
preferred for acrylates and for styrene-butadiene latices. Unfortunately
the carboxy groups can cause side reactions too. For instance, the carboxy
groups from the PSA can react with Si-H groups [221] or with the carbamate
release (as discussed earlier). The acrylic acid content increases adhesion
and adhesion hysteresis of acrylic copolymers. The addition of acrylic acid
also decreases a critical crack propagation speed above which a much
stronger rate dependence of the adhesion energy is observed (see Chap. 3).
Hydroxy functionality may be used for curing also. For instance, Takemura
et al. [137] used a solvent-based acrylate with 5 wt% HEMA and
PUR crosslinker. The main built-in crosslinkable monomers with
such functionality are listed in [220] as follows: acrylic acid, itaconic and
methacrylic acid (1–4 wt%), methylol acrylamide, and hydroxypropyl
Chemical Composition of PSAs 207

acrylate. Stabilizing hydrophilic comonomers e.g., ammonium acrylate,


sodium acrylate, styrene sulfonate, sodium sulfonate-hydroxypropyl
acrylate, and sodium-2-sulfoethyl-methacrylate were suggested also. Hydro-
philic comonomers e.g., partially neutralized methacrylic acid, hydroxyethyl
acrylate, and hydroxypropyl acrylate were employed for water-soluble
compositions. As described in [220] special monomers may work as
antioxidants, surfactants, biocides etc. and may improve removability,
flame resistance etc. For example, latexes were made using allyl ammonium
linear-alkylbenzene sulfonate as a surface active comonomer, to improve the
wettability and water resistance of the PSA. The contact angle of the latex
film produced with the non-migrating surfactant not only had a higher
original contact angle (125 vs. 80 ) but the contact angle did not drop even
after several minutes [222].

The Choice of the Polymerization Procedure. The polymerization


procedure influences the composition, molecular weight, macrostructure,
and microstructure of the polymers. The main polymerization methods
include free radical, ionic, and ionic-coordinative initiation, and chain
growth mechanisms. Their versatility for different monomers is quite
different. Ionic and ionic-coordinative polymerization lead principally to
polymers with higher structural order. Technological synthesis conditions
(e.g., gas, liquid, or solid state, solvated or dispersed systems) can influence
the polymerization reaction also. Ionic or ionic-coordinative polymerization
allows the regulation of the macro- and microstructure of the polymers.
Monomer insertion, sequence distribution, chain branching, and sterical
structure of the polymers can be finely regulated. The control of the
monomer insertion allows the build-up of side-chain unsaturated cross-
linkable polymers; sequence distribution and chain branching provide
segregated polymer structures. Free radical polymerization is the main
procedure for the synthesis of raw materials for PSAs. It is carried out in
solution, or in water-based dispersions. Solution polymerization is carried
out to produce linear polymers which may be subsequently crosslinked;
emulsion or suspension polymerization are carried out to synthesize linear
or crosslinked polymers. Ionic and ionic-coordinative polymerizations are
carried out in solution. Living (solution) polymerization with anionic
catalysts provides a greater control of molecular structures and composition
of synthetic rubber than radical emulsion or solution. With adequate
catalysts the control of the MW and MWD, the sequence distribution of the
monomers, the isomeric structure of the diene units, and either a linear or
branched (e.g., star) molecular structure was achieved. It should be noted
that the nature of the monomers may have a determinant role for the same
polymerization procedure (e.g., solution) and catalyst system. For instance,
208 Chapter 5

soluble metal-organic catalysts may lead to different tacticity (e.g., atactic,


syndiotactic) and sequence distribution (e.g., random or block) as a function
of the chemical nature of the monomers used (acrylics, vinylaromatics,
heterocyclics, and dienes) [74–85].
The choice of the comonomers in combination with special
polymerization conditions influences the final chemical composition. This
general statement is illustrated by the synthesis of special ethylene-maleinate
multipolymers [130] or special styrenic block copolymers (Dexco triblock
copolymers). As discussed in [223] in the MMA-EHA copolymer, the EHA
content increases vs. conversion differently as a function of the polymeriza-
tion method.
Molecular weight adjustment during synthesis offers another way to
influence the chemical basis and adhesive properties. Here solvent-based
or emulsion-based polymerization of the same monomers provides another
possibility to influence the molecular weight. Emulsion kinetics differ
from those of mass, solution, and suspension polymerization which show
an inverse relationship between the rate of polymerization and the
number average degree of polymerization. For instance, as discussed in [89]
solution polymerized acrylates have a MW of 200,000 to 300,000 compared to
emulsion polymerized acrylates with a molecular weight of 150,000–450,000;
however, mixtures of acrylate and tackifier polymerized together as the same
constitution have a molecular weight of 60,000–100,000 due to transfer
through the tackifier. The PSA mixed in a monomer state and solution
polymerized was no longer a simple blend of polymer and tackifier. The
polymeric conversion of the mixture was also lower.
According to [65] the main reason for the much lower shear resistance
for emulsion-based PSAs is due to the discrete microgel morphology of
such polymers in contrast to solvent-borne acrylics formed as a continuous
network morphology. Interlinking of the microgels by covalent bonds in the
film is needed. This can be achieved by the aid of a functional monomer,
e.g., IBMA.
Aqueous dispersions and solutions of the useful polymers can be
manufactured by procedures known in the art to be suitable for the
preparation of unsaturated olefin carboxylic acid ester polymers, such as
acrylic ester polymers. For instance, aqueous polymer dispersions can be
obtained by gradually adding each monomer simultaneously to an aqueous
reaction medium at rates proportional to the respective percentage of each
monomer in the finished polymer, then initiating and continuing the
polymerization with a suitable polymerization catalyst. Promoters are free
radical initiators and redox systems such as hydrogen peroxide, potassium
or ammonium peroxydisulfate, dibenzoyl peroxide, lauroyl peroxide,
ditertiary butyl peroxide, 2,20 -azobisisobutyro-nitrile, etc., either alone or
Chemical Composition of PSAs 209

together with one or more reducing components such as sodium bisulfite,


sodium metabisulfite, glucose, ascorbic acid, etc.
The acrylate monomer may include an alkyl group with 1 to about 10
carbon atoms. Suitable alkyl acrylate monomers are methyl acrylate; ethyl
acrylate; N-propyl; isopropyl acrylate; butyl acrylate; 2-ethyl hexyl acrylate;
and heptyl, octyl, nonyl, and decyl acrylate. The preferred alkyl acrylate
comonomers are butyl acrylate and 2-ethyl hexyl acrylate. Acrylic acid used as
a comonomer improves the cohesion [224]. Acrylic esters were copolymerized
with small portions of monomers such as acrylic acid, methacrylic acid,
itaconic acid, acrylamide, methacrylamide, acrylic esters, vinyl esters,
-alkoxy-alkyl unsaturated carboxylic acid amides, half esters, half amides,
amide esters, amides and imides of maleic anhydride, and the alkylaminoalk-
ylene monoesters of maleic, itaconic, or citraconic acids. If the average
number of carbon atoms in the longest alkyl chain exceeds 12, the adhesive
tends to become waxy and lacks sufficient adhesion. The useful PSA polymers
contain a sufficient amount of one or more of the described functional
monomers to increase the cohesive strength of the adhesive, relative to an
otherwise identical PSA in the absence of such functional monomers.
Detectable enhancement of cohesive strength is found in many polymers at
functional monomer concentrations as low as 0.05 wt%.
Working with a butyl acrylate/vinyl acetate/methacrylic acid model
terpolymer system, Brooks [225] used computer aided regression analysis
to identify key factors influencing the properties and PSA performance of
acrylic emulsion polymers. Experiments were designed to study the influence
of the acid comonomer level, the chain transfer agent, surfactants, and
buffer concentration on peel, shear, and tack values. Molecular weight,
polydispersity, and gel content were also measured. The strongest influences
were found with acid comonomer level and buffer concentration. The choice
of the monomers is discussed in detail in [220].

The Choice of the Additives. Additives influence the chemistry and


the physical interaction of the formulation components. They affect the base
polymer (e.g., initiators, crosslinker, filler) or the dispersed polymeric system
(e.g., surfactant, filler). The nature of the initiators influences the polymer
structure in free radical polymerization too. For instance, peroxides cause
more hydrogen abstraction resulting in more branching. Azo compounds
yield more linear polymers. Emulsion polymerization leads to a polymer
with a chemical composition which is strongly influenced by the compo-
nents of the dispersing medium. Its stability as a disperse system is affected
by such components too. Generally the stability of the emulsion depends
on the following parameters: electrical charges, elastic recovery, and gel
structure. Such parameters are a function of the monomers used and depend
210 Chapter 5

on the technological additives employed during manufacture of the


dispersed systems. The role of the additives will be described in detail
in Chap. 8.

Polymer Analogous Reactions


The polymer analogous reactions can be carried out on macromolecular raw
materials or on the finished product components in-line or off-line. Such reac-
tions lead to new polymers having special characteristics. They are affected
by built-in functionality and by the modification method used. Such reac-
tions include the functionalization of polymers and the use of the existing
functional groups for chemical reactions. The main synthesis procedures
based on polymer analogous reactions are grafting and crosslinking.

Functionalization. Functionalization allows the build-up of reactive


groups on polymer backbones. Industrially functionalization is carried out
during the synthesis of macromolecular products (mainly elastomers or
plastomers) or as a postreactor modification of plastomers used as carrier
webs for various pressure-sensitive products. It will be described in Chap. 8.

Grafting. Grafting is a general method for functionalization and


production of segregated structures. It was developed as a synthesis and
postreactor modification method. Grafting of reactive polar groups onto
elastomers or plastomers can be carried out off-line as a synthesis method
for polymers and/or it can be employed during the manufacture of pressure-
sensitive products, mainly as a surface treating method [226]. Although
grafting achieved a special importance for pressure-sensitive products
due to its use for the functionalization of the plastic carrier materials
(graft polymerization, plasma and chemical corona treatment [227]), most
developments in its use for PSAs have an experimental character [120]. For
instance, PSAs having a vinyl polymer backbone grafted with polysiloxane
moieties exhibit repositionability [228].

Crosslinking. The importance of crosslinking was discussed pre-


viously in Section 1.4. A detailed description will be given in Chap. 8.

2.2 Formulation
Some adhesives typically are used without or with quite low levels of
tackifying resins, but tackifying resins can contribute significantly to tack
and peel. In fact, formulation does not only mean tackification. Formula-
tion is the general notion for a recipe containing all the components
necessary to impart to an adhesive the combination of the final (desired)
Chemical Composition of PSAs 211

adhesive, coating, converting, and end-use properties. As described in [229]


‘‘in the industrial practice formulation is the blending of different
components of a recipe in order to achieve adequate technological and
end use performance characteristics.’’ In principle formulation is carried out
to achieve the end-use properties or to fulfill the requirements of a given
manufacturing technology. The end-use properties include the adhesive and
converting performances; the manufacture technology covers the manu-
facture of the adhesive and of the PSP too. Details of formulation are
described in Chap. 8. An exhaustive discussion of formulation is given
in [230].

2.3 Physical State of PSAs


As discussed in [229] formulation influences the adhesive coating technology
and vice versa. Generally the chemical compositions of solvent-based,
water-based, and hot-melt PSAs are quite different. As discussed earlier
for solvent-based and water-based PSAs, their formulating freedom is a
given, that is, their composition does not have any limiting factors other
than the Tg and modulus to be achieved. On the other hand hot-melt
PSAs have to be fluid at the coating temperature, that is, their melting point
has to be lower than their thermal resistance and, at the same time, they
must display a low viscosity at the (high) coating temperature, but a
high viscosity at the end-use temperature. This means that within the design
of the chemical composition of hot-melt PSAs the processability of
the adhesive should be considered. For this purpose conventional elastomers
are not adequate. In a similar manner the high coating temperature
requires aging and temperature-resistant tackifiers. In the case of water-
based PSAs, the adhesive should possess water solubility or dispersability
(i.e., the base elastomer should be designed so as to incorporate hydrophilic
monomers).

Raw Materials for Hot-Melt PSAs


Twenty years ago only solvent rubber-based PSAs existed. Natural, high
molecular weight rubber was used as the base elastomer. In order to achieve
better solubility and/or special adhesive properties, masticated or unmas-
ticated rubber is preferred. Even a highly masticated natural rubber (with
a low gel content) remains too viscous in the molten state in order to be
coated per se, and has too low a temperature resistance/cohesion at the end-
use temperature. [It should be mentioned that for special medical products
212 Chapter 5

‘‘dryblending’’ (warm or in the molten state) and coating of the blend by


roll pressing on a porous web were carried out [231].] For hot-melt PSAs
low melt viscosity rubbers are needed (actually viscosities of 7.5 Pa.s are
common); for these polymers an acceptable cohesion level may be achieved
only by physical crosslinking. Therefore thermoplastic elastomers are used
[232]. In general hot-melt PSAs are based on a thermoplastic elastomer,
an end compatible resin, a midblock compatible resin, oil, and antioxidant
[233]. The most commonly used block copolymers for hot-melt PSAs are
linear SIS block copolymers with a polystyrene content usually ranging
between 10 and 25%. They possess a relatively low modulus and are easier
to tackify than SBS types. SIS block copolymers display better tack, compat-
ibility, and aging resistance; they have a lower viscosity in hot-melt PSAs
and remain soft after aging. Current developments in the field of thermo-
plastic elastomers have broadened the available raw material base for hot-
melt PSAs (Section 1.1). The main raw materials used in hot-melt adhesives
(e.g., polyolefins and EVAc copolymers, polyesters, and polyamides) are
described in [234]. Dupont’s ethylene-vinyl acetate copolymers were
developed for use as a base resin in hot-melt PSAs. They are easily blended
and have an excellent thermal oxidation stability. Unfortunately they do not
provide much tack and peel. For hot-melt PSAs an unstatistical EVAc
copolymer with a higher VAc content, higher molecular weight, and a
sequential structure was evaluated as a raw material [234].
Although solvent and aqueous emulsion acrylic PSAs were in use for
some time, it is only recently that hot-melt acrylics were developed. The wide
variety of acrylic monomers available and their reactivity provide endless
opportunities to tailor polymers to suit specific applications. Reactive
hot-melt PSAs can be developed using solid and liquid functionalized
polyacrylic elastomers, in combination with special hardeners or catalysts
containing isocyanate for heat curing or UV curing with appropriate agents
such as -chlorobenzophenone or others. Such acrylates can be processed at
120 C and have a viscosity of 8 Pa.s [235].
Polyesters are also used as raw materials for hot-melt PSAs. Many
epoxy and carboxyl components can be used, hence the raw material base of
polyester-based hot-melt PSAs is broader than that of block copolymers.
The saturated copolyesters with 1–2 radiation curable double bonds have an
application temperature of about 100 C and a viscosity of 2–10 Pa.s. Through
synthesis or by postfunctionalization other reactive groups (e.g., carboxyl-,
alkyl-, isocyanate-, acryl-, or alkoxy-silyl groups) may be built into the
polymer. Copolyesters are not dangerous to human health (some have FDA
approval) and are insensitive to plasticizer migration [236]. The longer the
chain length, the better the cohesion and adhesion, but the higher the melt
viscosity.
Chemical Composition of PSAs 213

Raw Materials for Solvent-Based PSAs


Solvent-based acrylic adhesives represent about 40% of the total solvent-
based adhesive label stock production. Such systems are appreciated for
their high performance level and quality, but remain the most expensive.
They are increasingly being replaced by water-based adhesives but will likely
remain preferred for demanding applications (e.g., outdoor usage, high
humidity conditions). In the future, they may also be replaced by new hot-
melt or reactive solventless systems. Rubber-based solutions still account
for 55% of the solvent adhesives, but there is little research or development
going on in this area.
Through the choice of the chemical base for solvent-based PSAs, the
designer may use a broad range of chemically quite different raw materials.
Theoretically all the base components suitable for molten or water-dispersed
PSAs are soluble in organic solvents (i.e., they may be used for solvent-
based PSAs as well). End-use and commercial considerations may play a
determining role. Solution polymers generally have a narrower molecular
weight distribution (MWD) than emulsion polymers; therefore the
versatility of solution polymers allows a great deal of latitude [42]. Batch
polymerization can theoretically yield a MWD approaching 1.0; in practice
levels of 1.4–1.8 are more common. Continuous polymerization results in a
MWD of 1.8–2.0 in most cases. The MWD along with the molecular weight
exerts a very strong influence on processability.

Raw Materials for Water-Based PSAs


Among the various coating technologies water-based adhesives have grown
from a small usage in the 1970s to the status of preferred technology,
representing 63% of the whole label stock production in 1988. Actually they
cover more than 80%. Solvent- and hot-melt-based adhesives are estimated
at respectively 10% and 10% of the total production. Taking into account
the growing importance of the water-based PSAs technology a broad range
of raw materials was developed. It should be mentioned that unlike solvent-
based adhesives—where a rubber-based technology was consolidated and
where only a late development in polymer synthesis (acrylics) made the
manufacture possible of nonrubber-based PSAs—it was not possible to
develop an aqueous technology based on natural or synthetic rubber latex
in the field of water-based PSAs. The emulsions typically contain about
40–70% polymer when manufactured, whereas preferred latexes typically
have a content of about 40–60 wt% polymer solids. The correlation between
solids content and viscosity includes the particle size and particle size
distribution (PSD). A bimodal or multimodal PSD increases the solids
content. The dispersed polymer particles can be of any size suitable for the
214 Chapter 5

intended use, although a particle size of at least about 120 nm is presently


preferred. Most often the latexes have particles with a diameter in the range
of about 120–1000 nm. Decreasing the particle size (at very low particle size)
the viscosity increases exponentially. Introducing polydispersity in an acrylic
dispersion might result in a lower viscosity at the same volume fraction than
monodisperse spheres. Dispersions with spherical particles display lower
viscosities than other particle forms.
Acrylics as raw materials for water-based PSAs did not have any
competitors for a long period of time. Only lately has the development of
vinyl acetate copolymers led to a new, price attractive class of water-based
dispersions as raw materials for PSAs. Some years ago water-based and
hydrocarbon-based (rubbery) dispersions were synthesized on the basis of
(carboxylated) styrene-butadiene copolymers. Natural rubber, CSBR, and
acrylics make up the main base polymers for water-based PSAs.
Adhesives based on acrylic dispersions represent over 90% of the
European production of water-based label stock adhesives. The success of
water-based adhesives, particularly acrylics, is due to several factors. First
water-based adhesives eliminate the use of solvents and their associated
environmental concerns. Furthermore water-based adhesives display good
coating characteristics with different types of coating equipment and offer
good convertability to the converters. Blending with tackifier emulsions
allows the development of products with specific adhesive performance
characteristics, although adhesion on polyolefin substrates offers potential
for improvement.

Raw Materials for Radiation-Cured PSAs


As discussed earlier (Section 1.4) the main possibilities to formulate
radiation curable formulations include the syrup approach (monomer/
oligomer mixture), the reactive oligomer approach, and crosslinking of
pressure-sensitive adhesives. Electron beam-cured adhesives are usually
acrylate-based systems, formulated from oligomers with reactive diluents in
addition to the additives required to achieve the functional properties [237].
Generally prepolymers and reactive diluting agents are used [238], but the
danger of phase separation (especially at concentrations lower than 10%
of the diluting agent) remains [237]. The purpose of the reactive diluent
monomer in the UV-curable adhesive is to provide a solubility medium for
the resin system [239]. Such monomers used include 2-ethyl hexyl acrylate,
hydroxyethyl methacrylate, cyclopentadiene acrylate, and tetrahydrofur-
furyl acrylate. Diluent monomers may also effect the PSA characteristics.
A major obstacle in the application of radiation-curing technology for
adhesives was the development of film-forming properties without adversely
Chemical Composition of PSAs 215

affecting the tack. Therefore multifunctional monomers were used in limited


quantities in the adhesives, because they minimize the segmental motion in
the polymer and substantially reduce adhesive properties. On the other hand
too high a radiation dose may lead to damage of the face stock and/or loss
of tack, whereas too low a dose produces uncrosslinked monomers. There
are, however, multifunctional monomers with physical and functional
characteristics that provide positive contributions to the adhesive properties.
Such a monomer is tetraethylene glycol diacrylate which displays enhanced
polarity and molecular flexibility. When incorporated into the UV-curable
PSA formulations as an adhesion modifier, a permanent type adhesive is
obtained exhibiting 2.5–3.0 lb of peel adhesion with more than 2 h of shear
strength [239]. Substitution of tetraethylene glycol diacrylate with other
multifunctional acrylates (e.g., pentaerythritol triacrylate, hexanediol
diacrylate, trimethylol propane) imparts a nonpermanent characteristic.
As discussed earlier (Section 1.4) a range of 100%-solid hot-melt
PSAs on acrylic basis was developed using macromers [155,240]. Polyester
oligomers were suggested too. Low molecular weight polybutadienes
(MW ¼ 3000–5000) are able to make rigid resins more flexible and impart
rubber-type properties. In addition to reactive side group vinyl unsaturation
they possess a terminal functionality which can provide additional reactivity
and/or property enhancement; they can contain amine, vinyl, carboxyl, and
hydroxyl terminal functionalities, which can be incorporated in UV-curable
adhesives at 15 wt%.
A variety of photoinitiator systems were tested for radiation-curing
PSAs, including: benzophenone, isobutyl benzoin ether, diethoxyaceto-
phenone, and benzoin ethyl ether [239].
Using the same formulation but a different radiation method (UV or
electron beam) the adhesive characteristics are quite different [239]. The
adhesive values for electron beam-curing are overall quite low compared
to data acquired for the same formulations cured via the UV process. This
is probably due to the fact that the electron beam is a much higher
energy source than UV, thus promoting a higher level of crosslinking within
the system.
The first radiation curable styrene block copolymer (Kraton D-1320X)
was processed like a common thermoplastic elastomer. This is a star-shaped
SIS block copolymer. Classical hot-melt formulations do not respond
similarly when subjected to irradiation. To obtain optimum crosslinking
with a minimum of radiation, the selection of suitable resins and plasticizers
is necessary; for example, PSAs based on the newly developed SIS copoly-
mers crosslink much faster with saturated resins than with unsaturated
resins and plasticizers. Polybutadiene alone needs extremely high doses;
blends with acrylic esters do not display acceptable PSA properties. Only
216 Chapter 5

built-in hydroxy and epoxy groups and double bonds (e.g., ethyl hexyl
acrylates) give good results [238]. The best adhesive properties are achieved
with a dose of 30–40 kJ/kg; unsaturated polyesters and monomers with
alkylamino groups give good results for a radiation dose of 80 kJ/kg [238].

2.4 End-Use
High speed automatic label application requires moderate to high initial
tack in order to prevent label delamination in the production cycle.
Moderate to high shear properties are necessary on squeezable containers
to allow the label to remain in contact with the bottle during deformation
[241]. Good wet-out characteristics are extremely important for proper
application of clear label stock materials. See-through labels (no-label look)
also require water-like clarity of the adhesive system; therefore solvent-
based acrylics usually are used. Some applications using opaque label
materials require rubber-based systems. These few examples illustrate how
very different end-use requirements may influence the design of the chemical
composition. Next the features of composition design for the most
important PSA categories are discussed.

Raw Materials for Reel and Sheet Label Stock


Roll labels (usually on paper) are estimated at about 77% of the total
label volume; sheet labels and other products represent 20% and 3%,
respectively, of the overall production [242]. Quite different application
(labeling) technologies are used for small labels and decals, for machine
(gun) labeling, and hand applied labeling. In most cases the adhesive tack
and release force (peel force from the release liner) need to be closely
monitored. Reel labels require an aggressive tack, while sheet labels need
more cohesion and less tack. Therefore the base elastomer for reel/sheet
labels will be selected in a quite different manner.
Cohesive, low tack, high molecular weight polymers of short (side)
chain acrylics, and ethylene vinyl acetate copolymers are adequate for such
sheet applications. Soft, very tacky polymers of ethyl hexyl acrylate or butyl
acrylate (i.e., long side-chain acrylics) are preferred for reel labels. These
may be used as solvent-based or water-based adhesives. Competitive
formulations for reel labels may also be designed on a hot-melt PSA basis.
Water-based formulations are high tack, tackified acrylics, or CSBR-based,
tackified PSAs. The use of special ethylene vinyl acetate-maleinate multi-
polymers also was suggested [130]. The first-generation ethylene vinyl acetate-
dioctyl maleate copolymers offered good tack and peel strength, broader
specific adhesion than acrylates, and excellent PVC plasticizer resistance.
Chemical Composition of PSAs 217

Cohesive strength remains limited however, and the product is used mainly
in roll label applications at moderate to high service temperatures.

Raw Materials for Film Labels


Clarity considerations together with a high UV resistance limit the choice of
raw materials for film labels. Face stock/adhesive interactions also act as
limiting factors. Therefore only a few materials appear suitable as PSAs on
film face stock materials; the choice is restricted to the range of acrylics
(usually solvent-based ones).

Raw Materials for Permanent and Removable Labels


As discussed in Chap. 2, removability implies a low peel level and no
adhesion build-up. These criteria require low tack, high cohesion adhesives,
and good plasticizer compatibility. Because of the relatively low cohesion of
hot-melt PSAs and of the high migration tendency of water-based PSAs
(and their limited crosslinking) solvent-based PSAs were used as removable
PSAs. The latest developments in water-based acrylics allow the formulation
of removable PSAs on a dispersion basis. However, care should be taken
concerning the choice of the base elastomer. As an example it was observed
that ethylene vinyl acetate copolymers used for film coatings interact with
the soft PVC face stock material. Therefore aged peel values for removable
formulations are higher than desired. A detailed description of the chemical
basis of PSAs used for removable labels is given in Chap. 8.
As far as the pressure-sensitive adhesive component is concerned, the
properties which contribute to peelability are principally limited tack, a low
build-up factor, and a relatively soft adhesive. These properties can be
achieved in a conventional way by omitting or only using a low level of
tackifier, including tack deadeners such as waxes, fillers, or special agents,
and by including plasticizers. The chemistry of the polymer also has an effect
but, apart from choosing among polymers which are commercially
available, the chemistry is controlled primarily by the manufacturer of the
polymer rather than the adhesive formulator. Small amounts of a highly
polar monomer, such as acrylic acid, cause adhesion build-up. On the other
hand, small portions of acrylonitrile and methacrylonitrile improve
removability. An amount of about 30% natural rubber latex added to a
vinyl acetate-maleinate dispersion yields a removable PSA, without leaving
any adhesive residues upon debonding [243].

Raw Materials for Other Requirements


As discussed in detail in [230] there are numerous pressure-sensitive products
tailor made for special applications. In the past decade, to the main product
218 Chapter 5

classes of labels and tapes, were added protective films, forms, and special
products. In each such product range there are special requirements for the
environmental stability and regulation of the environmental stability [244].
These performance characteristics include temperature resistance and water
resistance as the main criteria. Both will be discussed in detail in Chap. 8.
Next a short description of the problem is given.
Temperature Resistance. The standard temperature resistance of
PSA formulation is given by the rheological characteristics of the
components. The stability in time of such performance characteristics is
affected by the chemical resistance of the raw materials. For instance butyl
systems should perform well at low temperatures, while acrylics are superior
in a room temperature environment [245]. Low temperature peel adhesion
for silicones increases at 75 C; for organic PSAs the peel value is zero
below 25 C [31]. Hence silicone PSAs are recommended for low
temperature conditions.
Water Resistance. Whereas wash-off labels were required on the
market some years ago [246], conventional PSAs are hydrophobic in nature
[141]. As such, they adhere poorly to wet or damp surfaces and generally
cannot be effectively removed with water, even using automated cleaning
procedures which employ detergents, warm water, or caustics.
Benedek [247] discussed the problem of the water resistance of PSAs
in detail. The water resistance of water-based adhesives depends on the
chemical composition of the base polymer and on the composition of the
emulsion. Generally the water resistance Rw is the sum of the synthesis given
water resistance Rws and of a formulation given water resistance Rwf [248]:

Rw ¼ Rws þ Rwf ð5:3Þ

Thus the postregulation of the water resistance can be achieved by


modification of the base polymer or (for water-based adhesives) of the
dispersed system. The influence of the monomers and of the surface active
agents used during the polymerization was described. As pointed out,
formulating depends on the adhesion-water resistance balance. Generally,
for common pressure-sensitive products water resistance is required. A
special performance is the water resistance of PSAs based on water-based
dispersions. Such products possess a heterogeneous, composite structure,
including migrating and water-soluble components (e.g., surface active
agents, monomers and polymers). Water-soluble surfactants remain in
the dried film and give water-sensitive films. By migrating to the film-air
interface surfactants cause the polymer film to be moisture sensitive
or hydrophilic; by migrating to the film-substrate interface they can cause
Chemical Composition of PSAs 219

a reduction in adhesion. Prevention of this migration can be achieved using


copolymerizable or volatile surfactants. By neutralizing a traditional anionic
surfactant with a reactive moiety, for example allylamine, one can render the
surfactant a reactive surfactant [249]. Allyl-acrylic latexes were produced
[250]. Ammonium salts of styrene-maleic anhydride lead to copolymers with
improved water resistance [105]. In some cases enhanced film hydropho-
bicity has also been found to enhance water whitening resistance [251,252].
For example Feng and Winnick [252] noted that polybutylmethacrylate
latex exhibited a high degree of water whitening whereas polybutylmeth-
acrylate-co-methacrylic acid latex copolymer did not. Miller and Barnes
[253] have found a strong effect of water-soluble low MW species on both
water absorbtion and water whitening. Water resistance can be improved by
the choice of the monomers too. As stated by Yang [254] a benefit from
vinyl neoester comonomer is the improved water resistance for copolymers
containing vinyl acetate.
Water-soluble (wash-off) formulations are limited by a narrow raw
material basis and low adhesion levels. The practical requirement for wash-
off labels were identified as the need to have a temperature-dependent water
solubility and wet tack (i.e., adhesion on condensed water coated surfaces).
A relatively simple formulation exhibiting water solubility can be designed
by changing the ratio of water soluble/insoluble components in a common
formulation (pseudo-solubility). Time- and storage-independent wash-off
formulations generally need more than 20 wt% special water-soluble
components.
Adhesives based on polyvinyl methyl ether were among the earliest
available water-soluble PSAs. While they showed fair adhesive properties
and are used for limited applications, several deficiencies restrict their suita-
bility, mainly that they did not absorb water rapidly enough to adhere to wet
surfaces. Adhesives based on polyvinyl methyl ether, however, were selected
for wash-off labels [141]. Partial esters such as polyvinyl methyl ether/maleic
anhydride with nonionic surfactants of the nonyl phenol ethylene oxide
adduct type have also been used. Polyvinyl ethers do not adhere to wet
surfaces and maleic anhydride copolymers are very humidity-sensitive.
Other water-soluble polymers such as polyvinyl pyrrolidone (PVP) with
water-soluble plasticizers, copolymers of acrylic acid and alkyl acrylates,
and others also were used [255]. A need exists for an adhesive which has
good adhesion to both wet and dry surfaces, whether cold or warm, whether
polar or nonpolar, with a sufficient permanent character to be retained as
tamper-proof when combined with label papers, but yet easily removed
when desired, with warm or cold water, or with detergents and alcohols used
in commercial or domestic cleaning operations. Such a formulation may be
given by acrylic acid-based polymers with a broad MWD [141].
220 Chapter 5

Water-soluble hot-melt PSAs are made from polyethylene oxide


(MW ¼ 50,000) and polyalkylene oxide (MW ¼ 190–15,000) [256]. Other
formulations contain a water-soluble ester of acrylic acid and polyethylene
or polypropylene glycol [257]. Water-soluble PSA is made up of 2-ethyl
hexyl acrylate or other tackifying monomers (optionally methyl methacry-
late or other diluent monomers) and 7–30 wt% hydrophilic monomers
selected from methacrylic acid (and salts thereof), hydroxy alkyl acrylates,
and hydroxy alkyl methacrylates [258]. Other water-activatable or soluble
hot-melt PSAs are based on vinyl pyrrolidine/vinyl acetate copolymers [259].
New generation acrylic water-based PSAs exhibit enough tack and
peel after immersion in hot water [260].

Adhesion on Porous Substrates. In the area of health care tapes and


disposable products, acrylate polymers are innocuous, in general, when
exposed to human skin. Many health and other tape (label) applications
exist where it is desirable to have a pressure-sensitive film coated onto a
porous web. Acrylic hot-melt PSAs offer a relatively easy route to coating
PSAs onto film or nonwoven webs, with a minimum of adhesive penetration
into the web [261]. Acrylic hot-melt PSAs have better initial tack after
aging than classical, rubber-based ones [262]. They have a viscosity of
26,000–30,000 mPa.s at 180 C and display good peel adhesion values on
different substrates [263]. A detailed comparison of the PSA properties
based on different raw materials will be given in Section 2 of Chap. 6.

2.5 Coating Method


The use of direct or transfer coating in the manufacture of pressure-sensitive
products depends on technical and economical considerations. The nature
of the product, its build-up, the nature of its components, and the technical
equipment affect the choice of a coating technology. The coating method
(direct or transfer coating) can influence the chemical composition of the
adhesive. One can assume that direct coating on porous surfaces allows a
rapid penetration of the adhesive in the face stock material.
Evidently, low molecular compounds migrate preferentially (see
Section 2.7 of Chap. 8). The viscosity of the coated adhesive and its
components influences both the wet-out and the bleedthrough. For instance
a special water-based dispersion for paper masking tapes, designed to be
direct coated on crepe paper with a coating weight of 38 g/m2, has 60%
solids content and a high viscosity (5000 mPa.s) [264]. For easy-to-wet,
directly coated face stock materials low viscosity dispersions may be used if
bleedthrough is not an important issue. For difficult surfaces (release liners)
or/and porous materials a higher application viscosity appears to be
Chemical Composition of PSAs 221

required. The viscosity of the PSAs to be coated may be controlled when


formulating (e.g., by choice of the solvents used, solids content, thickness,
pH, etc.). The viscosity of the bulk material (rheology) is composition and
molecular weight dependent.
Chemical Basis for Directly Coated PSAs. Each class of raw
materials (e.g., natural or synthetic rubbers, acrylics, ethylene vinyl acetate
copolymers, etc.) with or without tackifier may be used for direct coating of
different face stock materials. In each physical state (molten, dissolved, or
dispersed) materials may be used for direct coating onto paper face stock
materials. Restrictions must be taken into account concerning the porosity/
viscosity balance (usually for water-based PSAs) and the thermal sensitivity
of special face papers. As with film coatings, this acts as a limiting factor
when using hot-melt PSAs. Almost all rheological behavior (i.e., a high
fluidity or a shear-induced viscosity decrease) may be tolerated for direct
coating of paper or films.
In order to avoid penetration (i.e., bleedthrough of hot-melt PSAs) the
application viscosity should amount to 17,000–20,000 mPa.s at 180 C [262].
Hot-melt PSAs are also suggested for coating onto nonwoven substrates
[265]. Raw materials also differ as a function of the coating machine. As
an example aqueous acrylic tape adhesives are designed as a function of
the coater geometry [266]. The required viscosities are: for rotogravure,
75–100 mPa.s; for Meyer Rod, 300–800 mPa.s; and for knife-over-roll,
4000–8000 mPa.s. On the other hand foam generation is higher for Meyer
Rod than for rotogravure, and even higher for knife-over-roll. A detailed
discussion of the formulation differences as a function of the coating method
is given in Section 2.3 of Chap. 9.
Chemical Basis for Indirectly (Transfer) Coated PSAs. Indirect
(transfer) coating implies good wetting at low surface tension, or high
viscosity. Both are intrinsic properties of hot-melt PSAs or some solvent-
based PSAs. Difficulties appear mostly with the use of water-based PSAs,
where special wetting agents and thickeners must be used.

2.6 Solid State Components of the Laminate


Quite different plasticizer levels are required for the formulation of
adhesives used for films (50%) or other materials (2–6%) because of the
different flexibility and porosity of these substrates [267]. As illustrated
by this example the solid state components of the laminate influence the
choice of the adhesive components by their different affinity towards these
components and by changing the adhesive composition through chemical
interaction with the PSAs.
222 Chapter 5

Pressure-sensitive adhesives for film coating must meet some special,


face stock-related requirements. Pressure-sensitive adhesives for PVC
coating should meet the general requirements for film coating in addition
to some special PVC-related requirements, as the requirements for soft PVC
are quite different from those for hard PVC; similarly, filled (opaque)
PVC possesses different properties from clear PVC. Sensitive film face
stock materials need a nonaggressive adhesive formulation; for example,
soft PVC usually requires tackifier-free adhesive formulations. Stabilizers
present in the PVC (e.g., Pb, Sn) produce oxidative degradation of the
tackifier resin which is associated with the loss of the tack. Polyolefin
carriers impose improved wettability and anchorage, and aging properties.
For such materials dimensional stability is affected thermally.
Porous, primerless papers as face stock material need nonmigrating
adhesive formulations. Generally, migration of the low molecular weight
components from the adhesive (e.g., plasticizer, tackifier, surface active
agent, water, etc.) into the paper, or from the film (plasticizer) or paper
(water, special chemicals like components of thermal papers) into the PSAs,
may alter the chemical composition of the coated adhesive. These selected
examples illustrate how the chemical composition may influence the solid
state components of the laminate and vice versa. Not only the adhesive
properties, but also the coating characteristics of the PSAs may be influ-
enced by the solid state laminate components. As an example, carboxylated
neoprene latex is preferred to other latexes because it provides better specific
adhesion to aluminum foil. Tables 5.2 and 5.3 summarize the quite different
requirements imposed on PSAs as a function of the face stock material.

2.7 Product Build-Up


Pressure-sensitivity is the result of special chemical and macromolecular
build-up, but macroscopical product build-up may contribute to pressure-
sensitivity too. By transfer coating, the surface quality of the silicone layers
affects the smoothness of the PSA layer. The high gloss surface finish of the
release coating (with OPP) liner transfers to the adhesive increasing its tack
level by about 30% [268]. On the other hand transfer coating on such release
needs special formulation. Numerous pressure-sensitive products require
multiple (adhesive and nonadhesive) coated layers. In in-line coating
the coated web may become a complex structure, for instance: release-
corona-primer-PSA-corona-primer-PSA. As discussed in [269] the coating
technology for in-line manufacture is needed for special labels, tapes, or
forms. In such cases it is preferred to have the same vehicle for the different
coating layers (or to use 100% solids) and the adhesives should exhibit
Chemical Composition of PSAs 223

Table 5.2 PSA for Labels: Main Requirements as a


Function of the Face Stock Material

Face stock

Required Performance Characteristics Paper Film

Adhesive properties þþ þþ
Converting properties þþ þþ
End-use properties
Water resistance –– þþ
Shrinkage resistance –– þþ
Aging resistance þ– ––
Migration resistance –– þþ
Clarity þþ ––
þ þ very important, þ – important, – – not necessary

Table 5.3 Performance Characteristics of Different PSAs


as a Function of the Face Stock Material

Characteristics as a Function
of the Face Stock Material CSBR EVAc AC

Film Clarity þþ þþ þþ
Aging resistance –– þþ þþ
Water whitening resistance þ– –– þ–
Shrinkage resistance þþ þ– þþ
Paper Migration resistance –– þ– þ–
þ þ very important, þ – important, – – not necessary

coating versatility for direct and transfer coating too. A special formulation
is required for double-side coated and carrier less tapes [270].

2.8 Environment Related Considerations


Environmental considerations influence the chemical composition (formu-
lation) of the PSA also [269]. The most important problems concern
recycling during manufacture of PSAs and recyclability of pressure-sensitive
adhesive coated products, but advances in so-called ‘‘bio-based’’ pressure-
sensitive adhesives have been made too (see Chaps. 9 and 10).
224 Chapter 5

REFERENCES

1. H.A. Krecenski, J.F. Johnson and S.C. Temin, J. Macromol. Sci. Rev. in
Macromol Chem. and Physics, C26, 143 (1986).
2. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 2.
3. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000.
4. G.R. Hamed and C.H. Hsieh, J. Polymer Physics, 21, 1415 (1983).
5. Adhes. Age, (12) 36 (1986).
6. P. Green, Labels and Labeling, (11/12) 38 (1985).
7. J. Kendall, F. Foley and S.G. Chu, Adhes. Age, (9) 26 (1986).
8. J.C. Fitch and A.M. Snow Jr., Adhes. Age (10) 23 (1977).
9. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.
10. Papier und Kunststofjverarb., (9) 48 (1990).
11. Coating, (18) 240 (1972).
12. J.C. Chen and G.R. Hamed, Rubber. Chem. Technol., (2) 319 (1987).
13. G.R. Hamed and C.H. Hsieh, Rubber Chem. Technol., 59, 883 (1986).
14. Coating, (12) 75 (1969).
15. C. Craig, Adhes. Age, (7) 31 (1988).
16. Adhes. Age, (12) 35 (1987).
17. A. Zawilinski, Adhes. Age, (9) 29 (1984).
18. Labels and Labeling, (11/12) 38 (1985).
19. R. Midgley, Adhes. Age, (9) 17 (1986).
20. Du Pont, Bulletin, DS 83/7 DGC/JP.
21. Adhäsion, (11) 30 (1983).
22. I. Benedek, Adhäsion, (12) 17 (1987).
23. A. Sustici and B. Below, Adhes. Age, (11) 17 (1991).
24. P. McConell (Eastman Kodak, USA), US Pat. 4,072,812/07.02.78; in C.N.
Clubb and B.W. Foster, Adhes. Age, (11) 18 (1988).
25. G. Trotter (Eastman Kodak, USA), US Pat. 4,264,756/28.04.81; in C.N. Clubb
and B.W. Foster, Adhes. Age, (11)18 (1988).
26. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.2.1.
27. A. Steedman, European Adhes. Seal., (9) 6 (1997).
28. PCT/US/85/02424/WO 87.
29. R. Goodwin Jr., US Pat. 2,857,736.
30. K.L. Ullmann and R.P. Sweet, ‘‘Silicone PSAs and Rheological Testing,’’ Proc.
of the 22nd Annual Meeting of the Adhesion Society, Panama City Beach, FL,
February 21–24, 1999, p. 410.
31. L.A. Sobieski and T.J. Tangney, Adhes. Age, (12) 23 (1988).
32. J. Pennace, Screen Printing, (7) 65 (1988).
33. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 4.1.3.
34. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 4.1.2.
35. C.K. Otto, Allg. Papier Rundschau, (16) 438 (1986).
Chemical Composition of PSAs 225

36. E.G. Ewing and J.C. Erickson, Tappi J., (6) 158 (1988).
37. G. Holden, E.T. Bishop and N.R. Legge, J. Polymer Sci., (26) 37 (1969).
38. K.E. Johnsen and M. Dehnke, Tappi, Hot Melt Symposium 1987; in K. E.
Johnsen and Q. Luvinh, DEXCO. Tri-block Copolymers for Adhesives Industry,
European Tape and Label Conference, Exxon, Brussels, 1989, Part 19.
39. R. Koch, Kautschuk, Gummi, Kunststoffe, (9) 804 (1986).
40. N. De Keyser, ‘‘Styrenic Block Copolymers: the Quality Choice for Solventless
Adhesives,’’ in Thermoplastic Rubbers, Shell Technical Manual, 8.19, 02.1992,
Shell, UK.
41. D. St. Clair, Adhes. Age, (11) 23 (1988).
42. Kautschuk, Gummi, Kunststoffe, (1) 30 (1987).
43. J.A. Miller and E. von Jakusch (Minnesota Mining and Manuf. Co., St.Paul,
MN), EP0306232B1/07.04.1993.
44. F.F. Lau and S.F. Silver (Minnesota Mining and Manuf. Co., St. Paul, MN),
EP 0130087B/02.01.1985.
45. U.S. Patent 4163077; in F.F. Lau and S.F. Silver (Minnesota Mining and
Manuf. Co., St. Paul, MN), EP 0130087B/02.01.1985.
46. Adhes. Age, (11) 32 (1985).
47. A. Roos and C. Creton, ‘‘Adhesion of PSA Based on Styrenic Block
Copolymers,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 371.
48. J.C. Chen and L.J. Fetters, Polym. Eng. Sci., 27 (17) 1300 (1987).
49. M.L. Williams, R.F. Landel and J.D. Ferry, J. Amer. Chem. Soc., 77, 3701
(1955).
50. X. Lu, ‘‘Pressure Sensitive Adhesive Based on Block Copolymer Ionomer,’’
Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA,
p. 161.
51. K.E. Johnsen, ‘‘New Developments in Thermoplastic Elastomers,’’ Tappi
Hot-Melt Symposium 85, June, 16–19, Hilton Head, SC, 1985.
52. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 3.2.3.
53. G.A. Luscheykiy, F.M. Medvedeva, L.I. Voytesonok, M.K. Polevaya and
L.D. Pin, Plast. Massy, (19) 16 (1985).
54. D. Petit (Norton Performance Plast.,Wayne, NJ, USA) EP717091/15.12.95; in
Coating, (4)134 (1997).
55. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 285.
56. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 3.2.5.
57. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 5.
58. R. Jordan, Adhäsion, (1/2) 17 (1987).
59. J. Andres, ‘‘Haftklebstoffe in der Papier und Kunststoffverarbeitung,’’ 4. PTS
Klebstoff Seminar, PTS Vortragsband, Munich, Germany, 03/84, p. 36.
60. A. Zosel, Int. J. Adhesion and Adhesive, 18, 265 (1998).
61. M. Virin, in TECH 12, Advances in Pressure Sensitive Tape Technology,
Technical Seminar Proceedings, Itasca, IL, May, 1989.
62. P.K. Dahl, R. Murphy and G.N. Babu, Org. Coatings. Appl. Polymer Sci.
Proceedings, (48) 131 (1983).
226 Chapter 5

63. E.W. Ulrich, US Pat., 2,884,126 (1959).


64. A. Zosel, Adhäsion, (3) 17 (1966).
65. S.D. Tobing and A. Klein, ‘‘Synthesis and Structure Property Studies in
Acrylic Pressure-Sensitive Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001, Williamsburg, VA, p. 131.
66. Y. Sasaki, D.L. Holguin and R. Van Ham (Avery Int. Co., Pasadena, CA,
USA), EP 0252717 A2/13.01.1988.
67. C.M. Chum, M.C. Ling and R.R. Vargas (Avery Int., Co., Pasadena, CA,
USA), EP 1,225,792/18.08.1987.
68. R. Hinterwaldner, Coating, (4) 138 (1992).
69. H. Kuegler, US Pat., 2,544,692/25.01.1985.
70. Nat. Starch Chem. Co., USA, US Pat. 1,645,063; in Coating, (7) 184
(1974).
71. R.P. Wool and I. Lee, ‘‘Polymer-Solid interfaces; Role of the Sticker and
Receptor Groups,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28,
2001, Williamsburg, VA, p. 297.
72. L. Li, M. Tirrell, A.V. Pocius and G.L. Korba, ‘‘Adhesion Studies on
Acrylic PSAs: Temperature and Composition Effect,’’ Proceedings of the
23rd Annual Meeting of the Adhesion Society, Myrtle Beach, SC, Febr., 20–23,
2000, p. 31.
73. L. Li, C. Macosko, G.L. Corba, A. Pocius and M. Tirrell, ‘‘Interfacial Energy
and Adhesion Between Acrylic PSA and Release Coatings,’’ Proc. 24th Annual
Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 270.
74. Cr.I. Simionescu, N. Asandei, I. Benedek and C. Ungurenasu, European
Polymer J., (5) 449 (1969).
75. Cr.I. Simionescu, N. Asandei, I. Benedek and C. Ungurenasu, European
Polymer J., (6) 925 (1970).
76. I. Benedek, N. Asandei and Cr.I. Simionescu, European Polymer J., (7) 995
(1971).
77. Cr.I. Simionescu, I. Benedek and N. Asandei, European Polymer J., (7) 1549
(1971).
78. Cr.I. Simionescu, I. Benedek and N. Asandei, European Polymer J., (7) 1561
(1971).
79. Cr.I. Simionescu, I. Benedek, S. Ioan and N. Asandei, ‘‘Copolymerization
of Polar Monomers Using the Cp2TiCl2-AlEt3 Catalyst,’’ Proc. IUPAC Int.
Symp. on Macromolecules, Jul., 2–8, 1972, Helsinki.
80. Cr.I. Simionescu, I. Benedek, S. Ioan, M. Galin and N. Asandei, Rev. Roum.
Chim., 18, 325 (1972).
81. Cr.I. Simionescu, S. Ioan, I. Benedek and N. Asandei, Rev. Roum. Chim.,
18, 325 (1972).
82. I. Benedek et al., Rom. Pat., 56852/31.03.1970.
83. I. Benedek et al., Rom. Pat., 54213/1969.
84. I. Benedek et al., Rom. Pat., 59125/14.10.1971.
85. I. Benedek et al., Rom. Pat., 59674/07.01.1975.
86. P.A. Lovell and T.H. Shah, Polym. Commun, (2) 32 (1998).
Chemical Composition of PSAs 227

87. M. Kresti et al., WO 97/46634; in S.D. Tobing and A. Klein, ‘‘Synthesis and
Structure Property Studies in Acrylic Pressure-Sensitive Adhesives,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 131.
88. A. Aymonier, E. Papon, J.J. Villenave and P. Tordjeman, ‘‘Direct Relation
Between Copolymerization Process and Tack Properties of Model PSA’s,’’
Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg,
VA, p. 280.
89. M. Kajiyama, ‘‘Phase structure of PSA Prepared from Solution and
Emulsion,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 283.
90. E. Lauretti, F. Mezzera, G. Antarelli and A.L. Spelta, Gummi, Asbest.,
Kunstst., (16) 296 (1985).
91. J.A. Schlademann (Atlantic Richfield, USA), U.S. Pat. 0183386 A2/26.10.1984.
92. J.P. Keally and R.E. Zenk (Minnesota Mining and Manuf. Co., USA),
Canad. Pat. 1224.678/10.07.1982 (US Pat. 399350).
93. Coating, (7) 238 (1989).
94. H. Braun and Th. Brugger, Coating, (9) 307 (1993).
95. ‘‘UV-Curable Acrylic Hot-Melts for PSA Application,’’ BASF Symposium,
2/15/96; in S.D. Tobing and A. Klein, ‘‘Synthesis and Structure Property
Studies in Acrylic Pressure-Sensitive Adhesives,’’ Proc. 24th Annual Meeting
of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 131.
96. K.H. Schumacher and T. Sanborn, ‘‘UV-Curable Acrylic Hot-Melt for
Pressure Sensitive Adhesives-Raising Hot-Melts to a New Level of
Performance,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 165.
97. R. Pfister, Coating, (6) 171 (1969).
98. F.M. Rosenblum, Adhes. Age, (6) 22 (1972).
99. Coating, (2) 45 (1969).
100. Die Herstellung von Haftklebstoffen, T 1.2.2; 17d, Nov. 1979, BASF, p. 1.
101. Coating, (4) 23 (1969).
102. H.W.J. Mueller, Adhäsion, (5) 208 (1981).
103. K. Hagenweiler and K. Scholz (BASF), DOS, 2.328.430/1973.
104. K.C. Stueben, Adhes. Age, (6) 72 (1977).
105. Coating, (8) 247 (1969).
106. Adhäsion, (3) 13 (1974).
107. R.P. Mudge (National Starch Chem. Co., USA), US Pat. 4,753,846.
108. X. Han, S. Xu and H. Li, ‘‘Synthesis of Special Structural Vinyl-Acetate
Ethylene-Butyl Acrylate Latex,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001, Williamsburg, VA, p. 337.
109. E. Merz, Double Liaison—Chimie des Peintures, (226) 263 (1974).
110. Wacker GmbH, German Pat. 4,322,516 (1982).
111. Adhäsion, (12) 28 (1983).
112. Adhäsion, (2) 3 (1977).
113. G.R. Frazer (Johnson, USA), EP 259.842/16.02.88; in CAS, 21 (3) (1988);
109:111731.
228 Chapter 5

114. S. Cartasegna, Kautschuk, Gummi, Kunststoffe, (12) 1188 (1986).


115. R.P. Mudge (National Starch Chem. Co., USA), EP 0,225,541/11.12.85.
116. Adhes. Age, (11) 28 (1981).
117. A. Koch and C.L. Gueris, Allg. Papier Rundschau, (16) 40 (1986).
118. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 298.
119. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 4.1.
120. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 3.5.
121. R.D. Rich (Loctite Co.), EP 250.090/23.12.87 in CAS, 17 (92) (1988);
109:398787.
122. R. Jordan, Coating, (8) 213 (1982).
123. R. Jordan, Coating, (10) 278 (1982).
124. R. Jordan, Coating, (12) 335 (1982).
125. Allg. Papier Rundschau, (16) 462 (1986).
126. H. Fries, Gummi, Asbest, Kunststoffe, (9) 454 (1985).
127. A. Deshpande, D. Hyer, M. Kemper, K. Krajca, G. Locko and M. Moran,
‘‘Tackifiers for PSAs,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28,
2001, Williamsburg, VA, p. 177.
128. L. Jacob, ‘‘New Developments of Tackifier Resins for SBS Block
Copolymers,’’ 11th Munich Adhesive and Finishing Seminar, 1986, Munich,
Germany, p. 87.
129. Coating, (9) 27 (1972).
130. P. Mudge, ‘‘Ethylene-Vinylacetate Based, Water-Based PSA,’’ in TECH 12,
Advances in Pressure-Sensitive Tape Technology, Technical Seminar
Proceedings, Itasca, IL, May 1989.
131. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.1.1.
132. P. Kroger and W. Schimmel, ‘‘Vernetzung von Copolymerisaten auf
Acrylatbasis,’’ 11th Munich Adhesive and Finishing Seminar, 1984, in
Coating, (2) 66 (1985).
133. N. Toshio, N. Ken, A. Kazumi and F. Kazumide (Nitto Electric Co., Japan)
Japanese Pat., C3 275 79/05.02.88; in CAS, 17, (4) (1988); 109:39109m.
134. M. Novak, American Ink Maker, (10) 44 (1983).
135. A. Homanner, Adhäsion, (12) 16 (1982).
136. R. Milker and Z. Czech, ‘‘Solvent Based Pressure-Sensitive Adhesives,’’ 16th
Munich Adhesive and Finishing Seminar, 1991, Munich, Germany, p. 134.
137. A. Takemura, J. Asahara, R. Sugaya, N. Hori and H. Ono, ‘‘The Surface
Properties and Performances of PSA,’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 375.
138. O. Yoshinori, A. Kimihisa, A. Minoru and H. Takenao (Toa Gohsei, Japan),
Japanese Pat., C3 89 345/20.04.88, in CAS, 21, (10) (1988), 109:112327.
139. J.A. Fries, ‘‘New Developments in PSA,’’ in TECH 12, Advances in Pressure
Sensitive Tape Technology, Technical Seminar Proceedings, Itasca, IL, May
1989, p. 27.
140. Adhes. Age, (12) 43 (1981).
141. C.T. Albright, EPA 0099087B1/23.12.1987.
Chemical Composition of PSAs 229

142. Hercules, Bulletin, OR-212C.


143. Adhes. Age, (10) 24 (1977).
144. Brit. Pat., 1,063,324; in Coating, (4) 114 (1969).
145. W. Blum, Z. Czech and F.R. Herrmann (Lohmann GmbH & Co.KG,
Neuwied, Germany), DE 4243270A1/21.12.1992.
146. H.P. Seng, Coating, (7) 245 (1992).
147. P. Holl, Verpackungs-Rundschau, (3) 214 (1986).
148. E. Funk, Paper, Film, Foil Conv., (8) 65 (1973).
149. G. Von Pilar, ‘‘Einige Anmerkungen zur Diskussion über die Anwendung
strahlen-vernetzender Systeme im Bereich Trennpapiere,’’ 7th Munich
Adhesive and Finishing Seminar, Munich, Germany, 1982.
150. H. Klein, Plastverarb., 37, (4) 58 (1986).
151. G. Strohner, Adhäsion, (1/2) 24 (1985).
152. G.W. Drechsler, Coating, (7) 235 (1993).
153. Der Polygraph, (13) 963 (1973).
154. Der Polygraph, (14) 1000 (1973).
155. J.M. Loutz, Coating, (9) 328 (1987).
156. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 3.1.2.
157. J.O. Hendricks (Minnesota Mining and Manuf. Co., St. Paul, MN) US
Pat., 2,956.904; in J.R. Seidel, 4. PTS Klebstoff Seminar, München, PTS
Vortragsband, Nr. 02/1985, p. 64.
158. H.W.J. Müller, 4. PTS Klebstoff Seminar, München, PTS Vortragsband,
Nr. 03/1984, p. 12.
159. P. Holl, ‘‘Methoden unde Möglichkeiten der Elektronenstrahlvernetzung
bei bahnförmigen Materialien,’’ 5. PTS Klebstoff Seminar, München, PTS
Vortragsband, Nr. 02/1985, p. 84.
160. H. Röltgen, Coating, (12) 311 (1985).
161. Coating, (11) 292 (1982).
162. E. Pagendarm, 10th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1985, p. 42.
163. E. Strahler, Coating, (5) 163 (1996).
164. Papier u. Kunststoff Verarb., (2) 53 (1996).
165. Y. Sasaaki, ‘‘Electron Beam Curing of Pressure-Sensitive Adhesives,’’
European Tape and Label Conference, 03.07.1993, Brussels, Belgium.
166. C.V. Chieh (Morton Thiokol, USA), EP 0,147,142.
167. F.T. Birk, Coating, (9) 278 (1985).
168. W.J. Traynor, L.C. Moore, R.M. Martin and J.D. Moon (Minnesota Mining
and Manuf. Co., St. Paul. MN, USA); US Pat., 4,762,982/23.02.88; in CAS,
17 (4) (1988); 109: 39101c.
169. R. Hinterwaldner, Adhäsion, (3) 14 (1985).
170. W. Beumer, M. Koehler and J. Ohngemach, Radcure’86, Conf. Proc. 10th,
4/43 (1986); in CAS Coatings, Inks Related Products, 22:3 (1988).
171. G. Li Bassi and F. Broggi (Fratelli Lambert S.p.A, Albizzate, Italy),
Polym. Paint Colour J., 178, 197 (1988); in CAS Coatings, Inks Related
Products, 18:2 (1988).
230 Chapter 5

172. K. Li, P. Mallya, P. Iyer, C. Kuang and W. Wang, ‘‘Synthesis of


Benzocyclobutenone Containing Polymers for Ultraviolet, Light Curable
Pressure-Sensitive Adhesive Applications’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 368.
173. A.J. Morris, J. Plast. Film Sheeting, (1) 50 (1988); in CAS, Adhesives, 25
(1988); 109:191649v.
174. H.F. Huber and H. Müller (Dynamit Nobel AG, Troisdorf, Germany); in
10th Radcure, 1986, Conference Proc., 12/1–12/12.
175. C. Harder, ‘‘Acrylic Hot-Melts—Recent Chemical and Technological
Developments for an Ecologically Beneficial Production of Adhesive Tapes;
State and Prospects,’’ European Tape and Label Conference, Brussels,
Belgium, Apr., 28–30, 1993.
176. G. Auchter, J. Barwich, G. Rehmer and H. Jäger, Adhäsion, 37 (1–2) 14
(1993).
177. R. Hinterwaldner, Adhäsion, (10) 24 (1985).
178. E.M. Barisonek and G. Froelic, Adhäsion, (11) 28 (1985).
179. R. Hinterwaldner, Adhäsion, (12) 15 (1985).
180. M.M. Feldstein, T.A. Borodulina, R.Sh. Vartapetian, S.V. Kotomin, V.G.
Kulichikhin, D. Geschke and A.E. Chalykh, ‘‘Correlations Between
Activation Energy for Debonding and that for Self-Diffusion in Pressure-
Sensitive Adhesive Hydrogels,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001, Williamsburg, VA, p. 137.
181. T.A. Borodulina, M.M. Feldstein, S.V. Kotomin, V.G. Kulichikin and
G.W. Cleary, ‘‘Viscoelasticity of PSA and Bioadhesive Hydrogels Under
Compressive Load,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28,
2001, Williamsburg, VA, p. 147.
182. A.A. Chalykh, A.E. Chalykh and M.M. Feldstein, Polym. Mater. Sci. Eng,
81, 456 (1999).
183. G.V. Vinogradov, A.I. Elkin and S.E. Sosin, Polymer, 19, 1458 (1978).
184. M.M. Feldstein et al., US Pat. Appl. 09/900.67; in T.A. Borodulina, M.M.
Feldstein, S.V. Kotomin, V.G. Kuilichikin and G.W. Cleary, ‘‘Viscoelasticity
of PSA and Bioadhesive Hydrogels Under Compressive Load,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 147.
185. M.M. Feldstein et al., PCT Appl. US 01/21417(2001); in T.A. Borodulina,
M.M. Feldstein, S.V. Kotomin, V.G. Kuilichikin and G.W. Cleary,
‘‘Viscoelasticity of PSA and Bioadhesive Hydrogels Under Compressive
Load,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 147.
186. M.M. Feldstein, A.E. Chalykh, A.A. Chalykh, G. Fleischer and R.A. Siegel,
Polym. Mater Sci., Eng., 81, 467 (1999).
187. G.V. Gordon, T.M. Leaym, M.J. Owen, M.S. Owen, S.V. Perz, J.L. Stasser,
J.S. Tonge, M.K. Chaudhury and K.A. Vorvolakos, ‘‘Resin–Polymer
Interaction in Silicone Release Coatings,’’ Proc. of the 23rd Annual
Meeting of the Adhesion Society, Myrtle Beach, South Carolina, Febr.,
20–23, 2000, p. 39.
Chemical Composition of PSAs 231

188. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
‘‘Rheology in the Release of Silicone Coatings,’’ in Book of Abstracts, 215th
ACS National Meeting, Dallas, 29 April (1998).
189. B.Z. Newby, M.J. Chaudhury and H.R. Brown, ‘‘Macroscopic Evidence of
Interfacial Slippage on Adhesion,’’ Science, 269, 1407 (1995).
190. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter. 4.3.1.
191. G. Josse, O. Poizat and C. Creton, ‘‘Probe Tack Tests of PSAs on Silicone
Release Coatings,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society,
Myrtle Beach, South Carolina, Febr., 20–23, 2000, p. 36.
192. A.E. Bly, Adhes. Age, (10) 29 (1972).
193. J. Pennace and G. Hersey, PCT/IPN/WO 87/03537/18.06.1987.
194. L.H. Clemens, S.S. Kanter and M. Mazurek (Minnesota Mining and Manuf.
Co., St. Paul, MN, USA), US Pat., 4,728,571/01.03.1988; in Adhes. Age, (7) 45
(1988).
195. A. Fau and A. Soldat, ‘‘Silicone Addition Cure Emulsions for Paper Release
Coating,’’ in TECH 12, Advance in Pressure-Sensitive Tape Technology,
Technical Seminar Proceedings, Itasca, IL, May 1989, p. 7.
196. A.W. Bamborough, 16th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1991, p. 96.
197. I. Kesola, ‘‘Production of Release Paper,’’ European Tape and Label
Conference, Exxon, Brussels, Belgium, 1989, p. 9.
198. T. Cosgrove, I. Weatherhead, M.J. Turner, R.G. Schmidt, G.V. Gordon and
J.P. Hannington, ‘‘NMR Spin-Spin Relaxation Studies of Reinforced
Polydimethylsiloxane Melts,’’ Polym. Prepr, 39, 545 (1998).
199. M. Brogly, L. Vrevin, G. Castelein and J. Schultz, ‘‘Silicone Release Coatings/
PSA Adhesive Strength Modulation. How PMDS based Network
Composition and Structure Influence Nanoscale Properties of the
Interface,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 134.
200. A.N. Gent and J. Schultz, J. Adhesion, 3, 281 (1972).
201. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
Adhes. Age, (11) 35 (1998).
202. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
‘‘Release Force Control: The Bottom Line. Synergism Between Adhesive and
Liner; Proceedings of the Pressure-Sensitive Tape Council, 1999, Technical
Seminar, Washington DC, May 5–7 (1999).
203. M.K. Chaudhury and G.M. Whitesides, Langmuir, 7, 1975 (1991).
204. R. Thomas, Allg. Papier-Rundschau, (16) 437 (1986).
205. Coating, (12) 244 (1984).
206. Coating, (12) 366 (1986).
207. H. Röltgen, Coating, (11) 400 (1986).
208. R. Hinterwaldner, Coating, (3) 73 (1985).
209. K. Brach (Design Cotg. Co., USA), US Pat. 4,288,479/08.09.1981, in Adhes.
Age, (12) 58 (1981).
232 Chapter 5

210. R. Hinterwaldner, Coating, (6) 158 (1983).


211. D.H. Teesdale, Coating, (6) 243 (1983).
212. J.D. Blizzars and D. Narula (Dow Corning Corp., Midland, MI), EP
0,183,377/29.10.1984.
213. L. Bothorel, Emballages, (278) 372 (1972).
214. L. Dengler, Coating, (6) 143 (1974).
215. K. Kriz and T.H. Plaisance (De Soto Inc.) US Pat. 0129433 15.12.1987; in
CAS, Coating, Inks Related Products, 10, 2 (1988).
216. G. Galli, L. Penzo and F. Pina (Manuli Autoadhesivi, Italy) US Pat.
4,725,4547 16.02.88.
217. R.D. Gafford and G.R. Faircloth, Adhes. Age, (12) 24 (1987).
218. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.1.2.
219. D.J. Kinning, J. Adhesion, 60, 249 (1997).
220. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 3.1.
221. R. Thomas, Allg. Papier Rundschau, (16) 437 (1986).
222. A.K. Schultz and N. Kofira, ‘‘The Use of Non-Migrating Surfactants in
Pressure-Sensitive Adhesive Applications,’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr., 25–28, 2001, Williamsburg, VA, p. 163.
223. M. Okubo, Makromol Chem., Makromol. Symp. (35/36) 307 (1990).
224. Adhäsion, (2) 15 (1973).
225. T.W. Brooks, Tappi J., (9) 29 (1984).
226. M.D. Green, F.J. Guild and R.D. Adams, ‘‘Analysis of Functional Surface
Modification of Pre-Treated Polypropylene, Using Multi-Modal XPS and
AFM,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28, 2001,
Williamsburg, VA, p. 254.
227. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.
228. M.H. Mazurek (Minnesota Mining and Manuf Co., St. Paul, MN, USA), EP
46,739,332/15.08.1991.
229. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 2.1.
230. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000.
231. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter
4.2.1.
232. D. De Jaeger, 11th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1986, p. 87.
233. Adhäsion, (4) 28 (1985).
234. R. Hinterwaldner, Coating, (4) 110 (1991).
235. Acronal DS3429, EDM/KE,4/94,1994, BASF, Ludwigshafen, Germany.
236. European Adhesives and Sealants, (9) 3 (1987).
237. R. Kardashian, Adhes. Age, (6) 38 (1986).
238. J.R. Seidel, 4th PTS Adhesive Seminar, Munich, Germany, 1984, p. 49.
239. W.C. Perkins, Radiation Curing, (8) 8 (1980).
240. Coating, (3) 68 (1985).
241. J.M. Casey, Tappi J., (6) 151 (1988).
Chemical Composition of PSAs 233

242. J. Lechat, The Pressure Sensitive Labelstock Market in Western Europe, Finat
World Congress, Monaco, 1989.
243. Vinnapas, Eigenschaften und Anwendung, Technical booklet, Wacker,
Munich, Germany, 1976, p. 11.
244. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 2.4.2.
245. M. Gerace, Adhes. Age, (8) 85 (1983).
246. Allg. Papier Rundschau, (5) 228 (1987).
247. I. Benedek, Adhäsion, (4) 25 (1987).
248. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 131.
249. A.K. Schultz, PCT Int. Appl. WO 9832726 (1998).
250. A. Schultz and A. Siddiqui, PCT Int. Appl. WO9832773 (1998); in A.K.
Schultz and N. Kofira, ‘‘The Use of Non-Migrating Surfactants in Pressure-
Sensitive Adhesive Applications,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr., 25–28, 2001, Williamsburg, VA, p. 163.
251. D.R. Bassett, Coating Technology, J. Coatings Technology, 73 (912) 43
(2001).
252. J. Feng and M.A. Winnick, Macromolecules, 30, 4324 (1997).
253. C.M. Miller and H.W. Barnes, ‘‘Factors Affecting Water Resistance of
Latex-Based PSAs,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25–28,
2001, Williamsburg, VA, p. 153.
254. H.W. Yang, ‘‘Effect of Polymer Structural Parameters on PSAs,’’ Proc. of the
22nd Annual Meeting of the Adh. Soc., Panama City Beach, FL, Febr., 21–24,
1999, p. 63.
255. Coating, (10) 309 (1969).
256. Coating, (10) 180 (1975).
257. Adhäsion, (10) 80 (1966).
258. A. Kenneth (Allied Colloids), EP 0,147,067/29.11.1983.
259. Adhes. Age, (12) 53 (1982).
260. Coating, (3) 360 (1985).
261. Tappi J., 67 (9) 104 (1983).
262. R. Hinterwaldner, Coating, (7) 176 (1980).
263. Coating, (11) 316 (1984).
264. Ucecryl M10X, Preliminary Technical Bulletin, UCB Chemicals, Drogenbos,
Belgium, Sept., 1992, 3103/27582.
265. Wolfgang Graebe, Papier u. Kunststoffverarb., (2) 49 (1985).
266. Adhes. Age, (11) 28 (1983).
267. Coating, (4) 11 (1974).
268. P. Greenway, International Forms, (3) 40 (1997).
269. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 2.2.1.
270. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 30.
6
Adhesive Performance
Characteristics

1 ADHESION-COHESION BALANCE

Pressure-sensitive adhesives possess adhesion, required for bonding and


debonding, and cohesion necessary against debonding. Adhesion is
characterized by tack and peel, whereas cohesion is described by shear
resistance (and partially by peel). The special balance of these properties, the
adhesion/cohesion balance, embodies the pressure-sensitive character of the
adhesive. For any PSA, tackified or not, both tack and peel adhesion can
be considered as the end result of two distinct processes, namely, a bonding
process and a bond breaking process. The efficiency of the bonding process
is related to the adhesive’s ability to exhibit viscous flow. In order to achieve
peel adhesion the bonding stage involves some dwell time. During this time
the adhesive must flow in the absence of any externally applied forces. The
more liquid-like the behavior of the polymer under these conditions, the
more pronounced the degree of bond formation. The debonding process
involves a more rapid deformation of the adhesive mass. The polymer’s
resistance to deformation at higher strain rates becomes very important; the
higher this resistance, the higher the force which must be applied to separate
the adhesive from the adherend (i.e., the peel resistance). Therefore, high
tack, high peel strength adhesives should exhibit good flow at low strain
rates, but good resistance to flow at higher strain rates.

1.1 Tack
In Chapter 2 a short definition of the tack was given and the
interdependence of tack/adhesive rheology was discussed; this chapter

235
236 Chapter 6

describes the external aspects of the tack. Tack of PSAs is not an exactly
defined, physical characteristic; it may be defined as a separation energy.
Tack is a function of T (i.e., Tack > 0 if T > Tg). The tack is inversely
proportional to the elasticity modulus [1]. On the other hand, the modulus
depends on the branching, as does the Tg.
Tack is the resistance offered by an adhesive film to detachment from
a substrate; it is the measure of the stretchiness of an adhesive, or the ability
to form an instant bond when brought into low pressure contact with a
substrate to which the adhesive is to adhere [2]. Tack, and the methods used
to measure it were discussed by Johnston [3]. The concept of tack remains
difficult to define. It was called tack, wet tack, quick stick, initial adhesion,
finger tack, thumb tack, quick grab, quick adhesion, and wettability. The
Pressure-Sensitive Tape Council prefers quick stick [4] and defines it as
‘‘that which allows a PSA to adhere to a surface under a very slight
pressure.’’ The American Society for Testing and Materials [5] defines tack
as the force required to separate an adherent and an adhesive at the interface
shortly after they were brought rapidly into contact under a light pressure
of short duration.

Measurement of the Tack


Tack is measured either by touch or quantitatively by means of a loop tack
tester. Loop tack, defined as the ‘‘quick stick’’ tack value, is the force
required to separate at a specific speed, a loop of material (adhesive surface
facing outwards) which was brought into contact with a specific area of a
standard surface (in the absence of significant pressure). Tack also may
be measured by the rolling ball or rolling cylinder method (for a detailed
description of the methods see Chap. 10). In the last decade the development
of contact mechanics favored the use of probe tack-like tack test methods.
Asymmetric adhesion tests using either flat or spherical punches played an
important role in understanding the fundamental mechanism of adhesion.
The geometry of a spherical punch pressing against a flat substrate has been
used in fundamental studies of the tackiness of crosslinked elastomers [6–9].
Flat punch tests have often been the choice of industrial research labs and
quality control and are the basis of commonly used ASTM standards.

Tack Measured as Coefficient of Friction. It is believed that the


resistance to the rolling motion of a ball on a PSA reflects the tackiness
of the adhesive, because the motion must be closely related to bonding and
debonding processes, which occur simultaneously at the contact surface.
This way of expressing tack is useful in some practical cases, but the physical
meaning of the measured values is not necessarily clear. It is more scientific
Adhesive Performance Characteristics 237

to express the tack in terms of a rolling coefficient of friction which depends


on the physical properties of the materials [10]. The rolling coefficient of
friction can be determined experimentally from the ‘‘pulling cylinder’’
method more easily than from the ‘‘rolling ball’’ method, and one can
theoretically calculate the rolling coefficient of friction by making some
assumptions concerning the deformation and failure mode of PSAs.
The velocity of the rolling ball changes at every moment and, at the
same time, the rolling coefficient of friction varies as a function of velocity.
Assuming that F is the static frictional force, N is the normal force, and I
are the angle of rotation and moment of inertia of a cylinder, respectively,
f is the rolling coefficient of friction of the PSA, P and Mg are the force
components in parallel with and perpendicular to the rolling cylinder,
respectively, and if the cylinder is pulled at a constant velocity, P is given by
the following equation:

P ¼ ðf =RÞMg ð6:1Þ

The rolling coefficient of friction of a viscoelastic material can be written as


the sum of two terms:

f ¼ fc þ fa ð6:2Þ

where fc represents the rolling friction caused by the compressive


deformation of the substrate material and fa that caused by adhesion or
extensional deformation of the substrate, which are shown in Fig. 6.1.
In the case of PSAs fa must be much larger than fc. If compressive
deformation of the adhesive is neglected, the strain of the adhesive " at is
expressed as [10]:

" ¼ ðR=hÞð1  cos Þ ð6:3Þ

and the rate of strain "0 is:

"0 ¼ ðR =hÞ sin ¼ ðv=hÞ sin ð6:4Þ

where h and v are the original thickness of the adhesive layer and velocity
of the cylinder, respectively. For a mechanical model, the stress  generated
by the elongation of the adhesive can be expressed as a function of ,
where f is the angle at the moment of failure. Assuming that the moment
caused by the extended part of the adhesive is equal to fa Mg (according to
the definition of the rolling coefficient of friction) fa may be expressed as a
238 Chapter 6

Figure 6.1 Rolling friction caused by compressive deformation of the substrate


(adhesive) and by its extension deformation during rolling of the ball.

function of the cylinder characteristics (i.e., width b, radius R, moment, and


stress):
Z
fa ¼ ðR2 b=MgÞ ð Þ cos sin d ð6:5Þ
0

Graphs of fa versus log v are different for different viscoelastic models and
failure criteria [10]. Mizumachi calculated the fa/v functions for a single
Voigt element, a single Maxwell element, two Maxwell elements, multiple
Maxwell elements, and multiple Maxwell elements in parallel (Fig. 6.2).
In some cases a dependence of fa on the modulus E, viscosity , velocity
v, adhesive thickness, and cylinder radius was found. It would be reasonable
to expect that a curve of f versus log v for a PSA generally exhibits a peak
value or values that become small when the velocity becomes extremely
high or extremely low [10]. For adhesive practice it is very important to
summarize the dependence of the tack on the adhesive rheology (modulus/
viscosity) and on the coating weight Cw:

Tack ¼ f ðE,,Cw Þ ð6:6Þ

The influence of the rolling ball or cylinder geometry on the measured


tack value may be derived from the general conditions of the adhesion,
Adhesive Performance Characteristics 239

Figure 6.2 Viscoelastic model elements for PSA. 1) A single Maxwell element;
2) two Maxwell elements in parallel connection; 3) a single Voigt element.

friction, and wear of elastomers [11]. Equilibrium conditions and the


kinetics of adherence between a single hard spherical asperity and a smooth
surface of a soft elastomer were studied using fracture mechanics concepts
(see Chap. 3, Section ‘Contact Mechanics’). According to Greenwood [12],
when an elastic cylinder rolls over a plane, the movement of the area of
contact can be regarded as a propagating crack at the trailing edge (fracture)
and a healing crack at the leading edge. If the cylinder is truly elastic and the
process of forming or destroying free surface reversible, then there will be no
contribution to the rolling friction from the surface energy. In practice, with
an imperfectly elastic roller, the friction depends both on the viscoelastic
properties of the roller and on the surface energy; surprisingly the two seem
to be not additive but multiplicative. It has long been believed that only the
fracture process at the trailing edge is significant, and that the energy
recovered from the destruction of free surface at the leading edge is
negligible. Conditions for the appearance of reattachment folds and
detachment waves were established as a function of several parameters;
namely, the normal applied load, the sliding or rolling speed, and the radius
of the curvature. The radius aH of the contact area is given by:

aH ¼ ðPR=KÞ=3 ð6:7Þ
240 Chapter 6

where P is the normal applied load, R is the radius of the ball, and

K ¼ ð4E=3Þð1  2 Þ ð6:8Þ

where E and  are the modulus and Poisson’s coefficient of the adhesive. The
contact area is given by two different fields, one of compression in front of
the ball and another of tension (smaller than that of compression) behind
the ball [13]. The friction force of the ball depends on the temperature and
on the velocity of the ball, and this dependence is similar to that of the loss
modulus on the frequency and temperature (i.e., it may be described by
the Williams–Landel–Ferry equation) [14]. The friction is given by the
propagation of the small debonding ‘‘waves’’ (Fig. 6.3), which are related to
the peeling off of the ball,

VR ¼ nF ð6:9Þ

where  is the velocity of the ball, R is the local resistance against the motion
of the ball, V is the relative speed of the ball, and F is the peel force. As can
be seen from Eq. (6.9), rolling ball measurements can be formulated as peel
measurements. (Since the peel edge can be regarded as the tip of an
advancing crack, the adhesive ahead of the advancing peel front is subjected
to very large hydrostatic tension due to the lateral constraint imposed by the
substrate and the cover sheet [15]. The fibrils during peeling are modeled as
elastic strings and the fibril break down at a critical length like by the rolling
of the ball. It is evident that peel adhesion will depend on the relaxation of
the adhesive material. The friction is the result of the competition between
the entrainment of the adhesive by the ball and the interfacial relaxation due
to the molecular motion. This interfacial relaxation depends on the rheology
of the PSA, but at the same time it depends on the thickness of the adhesive

Figure 6.3 Friction given by the propagation of small debonding ‘‘waves.’’


Adhesive Performance Characteristics 241

and its anchorage onto the face stock material. If this holds, the nature of
the face stock surface has a strong influence on the tack measured by the
rolling ball. This conclusion is very important because of the quite different
dependence of the tack measured as loop tack on the face stock material. In
this case the flexibility of the surface plays a determining role on the mea-
sured tack value. It should be mentioned that, on the basis of our knowledge
about peel adjustment with the aid of the face stock surface (primer), it is
quite normal to have the same dependence for tack measurements by rolling
ball, whereas shown earlier, peeling phenomena occur.
The examination of the friction force allows a correlation between tack
and shear resistance. The friction force Ff is the sum of two components [16]:

Ff ¼ Fadh þ Fdef ð6:10Þ

where Fadh is the adhesion component and Fdef the deformation component.
The latter is a function of the internal cohesion of the material, therefore
tack will be a function of the cohesion as well.
Tack Measured as Cohesion. Hamed and Hsieh [17] measured tack as
cohesion. In the debonding step tack is a function of the cohesion; the upper
limit of the tack of an elastomer is its cohesive strength. If during bond
formation, complete contact and interdiffusion occur, then the interface will
be modified, and the measured tack must be identical to the cohesive strength
of the elastomer [17]. Hamed and Hsieh proposed ‘‘relative tack’’ (tack
divided by cohesive strength) as a measure of the completion of the tack
bond. If the relative tack equals zero, then no tack bond was formed. Relative
tack is rate dependent; Hamed and Hsieh measured tack as the average peel
force per unit width. At low test speeds the ratio of tack to cohesive strength is
less than the same ratio at higher test rates [17]. At sufficiently low stress rates
the short inter-diffused chains have time to relax, relieving the applied stress.
On the other hand, the cohesive strength is controlled by the longer chains,
which cannot slip past one another very easily. At higher strain rates even the
short molecules may have little time for the molecular rearrangements needed
to relieve stress; hence bond strength will increase rapidly as there is
insufficient time for failure by chain slippage and chains may rupture.
Tensile tests of SIS block copolymers show that decreasing the SI
content or increasing the crosshead velocity leads to a higher stress for all
extensions and to a more pronounced strain hardening [18]. The strain
hardening observed in the fibrillation part of the probe test curves can be
correlated to tensile tests; in both cases the lower the SI content, the more
pronounced the strain hardening. Experimental results show, that the
beginning of the probe test (where the cavitation process takes place)
242 Chapter 6

quantitatively corresponds to the small strain shear properties of the


adhesives (i.e., tack depends on the cohesion) while the end of the test
(fibrillation, strain hardening) can be qualitatively—so far—related to the
large strain elongational processes.
Tack and Adhesive Fracture Energy. Adhesion and tack of polymers
are not fundamental material properties like the modulus or the viscosity;
they strongly depend on the test methods used and the measurement
conditions. Adhesion performance is characterized by the adhesive fracture
energy. Two stages characterize the tack test that is based on the
measurement of the strength vs. time: the contact phase and the separation
phase. The tack energy (G) increases with the aptitude of the adhesive to
deform by cavitation and stringing during the debonding process. For ‘‘non-
stringing adhesives’’ (low viscous dissipation in the bulk of the adhesive) G is
sensitive mainly to the surface energy and to the surface roughness of the
probe. According to [19] a constant contact load is applied during a given
contact time then the probe is pulled up at constant debonding rate; a
constant contact area is applied for very short contact time (less than 1 sec)
and the probe is pulled up. The force and the contact area are monitored vs.
time. The product of the area subtended by the debonding force vs. time by
the pulling rate gives the tack energy. Actual contact area increases with the
thickness and then applatizes. Tack energy is a square function of the wet
area. First Zosel [1] developed an instrument which measures the fracture
energy and allows the adjustment of the most important parameters during
the bonding and debonding processes, such as contact time, contact
pressure, rate of separation, and temperature. The (bonding/debonding)
force vs. time plot of the instrument can easily be transformed into a stress-
strain curve, giving the tensile stress as a function of the tensile strain during
bond formation and separation. An appropriate measure of adhesive bond
strength and tack is the energy of separation per unit area of surface (i.e., the
adhesive failure energy) [20–22]. This apparatus resembles the probe tack
tester and a very high correlation between both methods was found.
According to [23] the use of a Mechano Optical Tack tester allows the
determination of the tack strength (F ), tack energy (G), and the actual
contact area (A), for tests at controlled contact force (Fc), contact time (tc),
and release rate (r) [24–29]; G (J/m2) is given by the relation:
Z tf
G ¼ ðr=AÞ FðtÞ dt ð6:11Þ

Tack curves are load displacement curves generated during asymmetric


adhesion test. Several past investigations have shown that for elastomeric
Adhesive Performance Characteristics 243

adhesives the relation between the crack tip velocity and G is a unique
function for each combination of substrate and adhesive [30,31]. Empirically
one suggested form of this unique function is:

G ¼ Go ½1 þ  ðaT , . . .Þ ð6:12Þ

where  ¼ (/*)n, Go describes the interfacial contributions to the adhesion,


and  includes the quantities pertaining to bulk dissipative processes near
the crack tip. Theoretical models were proposed by Gay and Leibler [32]
and Creton and Leibler [33] to account for the effect of the surface energy,
surface roughness, and the adhesive Young’s modulus on the cavitation and
fibrillation process of stringing adhesives.
Peel Test Used for Estimating the Tack. A modified 90 peel test
(now known as PSTC Test Method 15 for Quick Stick) is used for tack
measurements [34,35].
Tack Measured as Plasticity. The measurement of the tack is not as
reproducible as the measurement of the Williams plasticity [36]. Practically
tack is measured as quick stick, loop tack, or rolling ball tack. None of these
methods is sufficient to characterize tack alone (see Chap. 10). Therefore the
tack index was defined as [37]:

A þ B þ 50ð17  CÞ
Tack Index ¼ ð6:13Þ
3

where A is the value measured in a 90 quick stick test, B is the value of a
loop tack test (g/cm), and C is the value of the rolling ball tack in cm (up to
a maximum of 15 cm). This index represents an attempt to take into account
the three measures simultaneously.
Tack Level. The main adhesive properties (tack, peel, and cohesion)
depend on the state of the polymer (pure or formulated), the coating
technology (direct/transfer), the coating weight, and the face stock. Of
course the most important parameter remains the chemical composition of
the polymers. Thus estimating the tack level supposes a comparison of PSAs
on the basis of different polymers. Acrylics are polymers of long-chain alkyl-
acrylates. There are only a few commercially available low-Tg acrylates
(e.g., ethyl hexyl- and butyl acrylate). High tack implies a high level of these
fluidizing comonomers. The obtained polymers are tacky, but they lack the
necessary cohesive strength. Hence one can conclude that common
(nonformulated) acrylic PSAs are not tacky enough and that special acrylic
PSAs are not cohesive enough. On the other hand, the tack of pure acrylic
244 Chapter 6

PSAs is lower than that of classical (tackified) rubber-based PSAs, but


higher than that of other untackified adhesives (except polyvinyl ethers):

Tack : AC ðunformulatedÞ > SBR ðunformulatedÞ


ð6:14Þ
> EVAc ðunformulatedÞ

The superior tack level of unformulated acrylic PSAs is illustrated in the


Table 6.1, comparatively to water-borne EVAc or CSBR. It can be seen that
acrylic PSAs possess a better tack at any given coating weight.

Factors Influencing the Tack


Tack is influenced by the nature and amount of the adhesive (coating
weight) and face stock material; as shown earlier test methods and
conditions also influence the measured tack value.
Influence of the Nature of the Adhesive. Tack may differ according to
the chemical composition or state of the adhesive. Pressure-sensitive
adhesives that have different chemical bases (e.g, natural/synthetic rubber,
acrylics, vinyl acetate copolymers, maleinates) exhibit different tack levels.
On the other hand, PSAs within the same class of monomers (e.g., acrylic
PSAs) may display different tack, depending on the specific monomer
characteristics, chain structures, and molecular weight. However, most
PSAs are formulated and the formulating additives (tackifiers, plasticizer,
etc.) change the tack. Water-based PSAs need special formulating additives.
Thus the nature of the uncoated adhesive (solvent-based, water-based, or
hot-melt) will also influence its tack. Generally, additives such as tackifiers
(resins or plasticizers) improve the tack; their influence depends on their
compatibility. As an example, polybutenes added as a solution in toluene
generally improve the rolling ball tack, whereas the use of Acronal 81 D
causes a marginal decrease in tack [38].

Table 6.1 Tack of Unformulated PSAs on Different Chemical Bases.


Rolling Ball Tack of Water-Based Acrylic, CSBR, and EVAc PSAs

Rolling Ball Tack (cm)

Coating Weight (g/m2) AC CSBR EVAc

22 1.5 5.8 17.0


40 1.5 5.8 8.5
AC, V 205 (BASF); EVAc, 1360 (Hoechst); CSBR, 3703 (Polysar).
Adhesive Performance Characteristics 245

Influence of Molecular Weight. Tack is a function of the molecular


weight. Fibrillation supposes a molecular weight higher than the
entanglement molecular weight. Previous studies [39,40] on PIB showed
that tack and adhesive fracture energy decrease monotonically for
MW > 45 kg/mol. Zosel [1] demonstrated that the tack of polyisobutylene
increases with increasing molecular weight, up to a maximum for
MW ¼ 5  104. Midgley [41] noted a slight tendency for tack to depend
on increasing molecular weight. According to Midgley, pressure-sensitive
adhesive-like blends containing longer chains do not show a strong
dependence of the work of adhesion (Wa) on MW, temperature, and
mixing ratios, especially as Mv exceeds 1000 kg/mol. This seems to be in
contrast to the pronounced tack maximum at Mw/Me around 4 observed
by Tobing et al. [42,43]. The useful range of molecular weight in which
good PSAs can be identified is limited. The upper limit of molecular weight
corresponds to adhesive failure caused by a lowered ability of the adhesive
to absorb energy before the interfacial bond is broken or by poor wetting
because of the lowered flow of the adhesive. For solvent-based acrylic
PSAs, the viscosity is dependent on molecular weight, and thus limited
by it. For water-borne acrylic PSAs there is no technological molecular
weight limit, but there is an optimum molecular weight for superior
tackification.
At a sufficiently short contact time the bond strength (tack) is due
primarily to the interdiffusion [17]. The self-diffusion rate is inversely
proportional to the second power of the molecular weight; thus tack
decreases with increasing molecular weight. On the other hand, adhesive
diffusion in the face stock material influences the coating weight and tack
depends on the coating weight. Therefore the molecular weight will have a
contradictory effect on the tack. Relatively low molecular weight base
elastomers and relatively high molecular weight tackifier resins should be
used for a good tack.
The dependence of the tack on the molecular weight is demonstrated
by the aging of SIS rubbers [44]. In this case chain scission occurs and
tack increases. In general, resins with a softening point below about 50 C
impart tack, but give poor cohesive strength, while resins with a softening
point above 70 C give good cohesive strength but poor tack [45]. Tack
properties at a given tackifier level are superior with the lower melting point
tackifier [46]. For a natural rubber/rosin ester mixture a resin with a lower
melting point (e.g., Staybelite E 10, melting point of 75 C) imparts the
maximum tack at a higher level (52%) than a higher melting point resin
(e.g., Pentalyn H, melting point of 95 C) at a 33% loading level. At the same
time, for maximum peel adhesion, an even higher tackifier level (70%)
is needed.
246 Chapter 6

For CSBR tackification, increasing the melting point of the resin


improves the quick stick of the adhesive [47]. New, high temperature
tackifiers for formulating PSAs were proposed [48]. Narrower MWD resins
contribute to a higher tack than broad MWD resins [49]. A detailed
description of the resin characteristics and the correlation resin melting
point/tack is presented in Chap. 8.
Effect of Crosslinking and Sequence Distribution. Crosslinking yields
high shear and lower tack properties [50]. As demonstrated for formulations
containing an electron beam crosslinkable rubber, the tack is good and
remains fairly constant for the Polyken probe and loop tack, regardless of
the applied electron beam dose. However, the rolling ball tack values tend
to become slightly poorer as the radiation dosage is increased. Generally,
crosslinking reduces the chain mobility with increasing Tg, thus decreasing
the tack. As an example, crosslinkable water-based acrylic PSAs (e.g.,
Acronal 29 D, 30 D) suffer a loss of tack after thermal crosslinking.
The tack of silicone PSAs depends on their coating weight and
curing [51]:

Tack ¼ f ðCoating weightÞ,ðCuringÞ1 ð6:15Þ

Lower peroxide levels and a greater adhesive thickness increase the


tack. High peroxide levels increase crosslink density and cohesive strength
[generally 0.5–3.0% benzoyl peroxide (BPO) is used]. The modification of
the tack by crosslinking is a function of the materials used, and of practical
experience. Despite the reactivity of Cymel resins, the tack of the modified
neoprene latex remains quite excellent [52]. Aqueous PUR dispersions were
used to improve the adhesion of acrylic adhesives, for better film formation
and mechanical properties. The use of nonionic PUR yielded higher shear
strength. The tack of formulation was also higher [53].
Tordjeman et al. [23] examined tack as a function of the manufacture
of the adhesive (batch, continuous feed with constant and with variable
comonomer ratio). Tack strength as a function of the time gives a peak for
batch and a bimodal curve (shoulder) for mixture. Tack is higher for batch.
This behavior can be correlated with the heterogeneous composition and
rather high cohesion of the batch type PSA. A very long plateau is observed
on the rheological master curve for batch type, which behaves as a
crosslinked elastomer.
Styrene block copolymers are physically crosslinked systems. In such
compounds the sequence distribution influences the tack. As stated by Roos
and Creton [18] the measured  max in probe tests decreases as the amount of
diblock increases.
Adhesive Performance Characteristics 247

Tack of Uncoated PSAs. Surface-active agents in water-based formu-


lations reduce the tack [41]. This influence appears to be less pronounced for
the other adhesive characteristics. With the exception of a slight change in
rolling ball tack, little difference was found between the same adhesives cast
from water- and solvent-based formulations. In fact, the different tack level
of solvent- or water-based PSAs may be observed mainly at low coating
weight (undercoating). At higher coating weight the tack level of water-
based acrylic PSAs approaches that of the solvent-based ones (Table 6.2).
The obtained tack level also depends on the nature and amount of the
surface-active agents [54]; fillers (e.g., silica) also decrease the tack [55].

Influence of the Coating Weight. The thickness of the adhesive layer


influences the flow conditions. Therefore the coating weight will influence
the tack; in general the tack is directly proportional to the coating weight.
The exact nature of this dependence also is influenced by the physical state
of the adhesive. Figure 6.4 illustrates the dependence of the tack on the
coating weight for a solvent-based and water-based adhesive, respectively,
namely, that tack increases with increasing coating weight of the PSA.
Solvent-based adhesives display a higher tack at lower coating weights.
The dependence of the tack on the coating weight is not linear; tack depends
on the nature of the adhesive, face stock, and substrate,

Tack ¼ f ðCw Þ ð6:16Þ

where  6¼ 1.

Table 6.2 Tack of Solvent-Based Acrylic PSA Vs. Water-


Based Acrylic PSA as a Function of the Coating Weight

Rolling Ball Tack (cm)

Solvent-based Water-based
Coating Weight (g/cm2) PSA PSA

8.0 4.30 6.9


10.0 3.20 6.0
12.5 2.80 5.2
15.0 2.40 4.0
17.5 2.20 3.5
20.0 2.10 2.8
22.5 2.00 2.3
25.0 1.95 2.0
27.5 1.30 1.9
248 Chapter 6

Figure 6.4 Dependence of the tack on the coating weight. Rolling ball tack as a
function of the coating weight for 1) water-based and 2) solvent-based PSAs.

Influence of Time/Temperature on Tack. Greenwood and Johnson


[56] have predicted that by increasing the velocity of separation, viscoelastic
effects extend the length over which the surface forces act, thereby increasing
the total force of adhesion. Probe tack values increase with the applied load;
even after a 100 sec dwell time the tack value is still increasing, indicating
that optimum contact has not yet been achieved [57]. Bates [58] proposed
that tack could be described by an equation of the Arrhenius type such that
the tack energy T:

T ¼ Aemx ð6:17Þ

where A and m are constants which involve the effect of load and dwell time,
and x is the separation rate of the probe.
Tack is known to go through a maximum around Tg þ 60 C [43].
Since the glass transition temperature of PIB is approximately 60 C,
tack measurements were performed at 10–25 C. Zosel [1] studied the tack
dependence of polyacrylates on the temperature and found that the
tack increased with increasing temperature going through a maximum at a
given temperature as a function of the acrylate structure. Branching pro-
duces a shift of the modulus towards lower temperatures, that is, branched
Adhesive Performance Characteristics 249

acrylates display the same modulus at lower temperatures. According to [18]


increasing the temperature from 22 C to 60 C leads to lower  max and "max
values in probe tack tests regardless of the percent diblock content of the
base polymer. At high temperature the interfacial crack propagation
becomes easier and debonding proceeds quickly by crack propagation.
The most useful signals in PSA analysis are G0 (the shear modulus) and
G (the loss modulus). G0 is a direct measurement of the tack strength, and
00

G00 is a direct measurement of the peel strength [59]. It is assumed that the
application of the adhesive (tack test) would occur at an angular frequency
near 0.1 rad/sec. It is further assumed that the peel test would approximate
the higher frequency of 10 rad/sec. The G0 range of 50,000 to 200,000 Pa is
accepted as an ideal PSA performance. Frequency dependent F-D measure-
ments (force-distance curves) with friction force microscopy (FFM), were
made as they represent a nanoscale analogous to the macroscopic probe
tack test; lateral modulation measurements are being developed, too.

Influence of Test Methods and Conditions. Different test methods give


rise to different tack values. On the other hand, labeling practice points
towards a strong influence of the labeling conditions (mainly of the
temperature) on the tack. As discussed in detail in [60] an ideal PSA gives
instantaneous tack and peel on a rigid substrate. Certain pressure-sensitive
products must adhere to soft substrates and provide enough adhesion after a
given time under special application conditions (temperature, pressure,
dwell time, etc.) only. For such products (e.g., protective films) ‘‘application
tack’’ characterizes better the adhesivity of the product. Chap. 2 gives a
detailed description of the tack dependence on the experimental conditions
(time/temperature).

Influence of Face Stock Material on Tack. As observed when


measuring the tack with different methods, the nature and dimensions
of the face stock material influence the value of the tack. Loop tack
measurements are sensitive to the face stock flexibility, while rolling ball
tack is influenced by the anchorage of the adhesive on the face stock (i.e., by
the face stock surface). Wet-out on the face stock material influences the
tack of the pressure-sensitive laminate; however, experimental results show
some discrepancies; Toyama et al. [61] found maximum tack for a surface
tension of the face stock equal to the surface tension of the adhesive. On the
other hand, Sherrif et al. [62] and Counsell and Whitehouse [63] found that
tack increases with increasing surface tension of the face material. The
surface influences the tack values measured as loop tack (Table 6.3) [64].
Tack characterizes the short bonding time and debonding behavior of
the adhesive. During bonding the pressure-sensitive laminate must conform
250 Chapter 6

Table 6.3 The Influence of the Face Stock Material on the Tack

Loop Tack (N)

Coating Weight (g/m2) Soft PVC Paper

22 7.3 5.7
24 5.8 5.5
43 3.3 3.2
46 2.5 3.8
56 3.3 4.2
A water-based EVAc PSA was used.

to the substrate in order to establish a maximum contact area. A high


debonding resistance is favored by a large contact area. Thus the flexibility
(i.e., the bulk properties) of the face material will strongly influence the tack
of the PSAs. This behavior is illustrated by the sensitivity of the tack test
methods to the flexibility of the face stock material.
In rolling ball tack tests the adhesive does not merely contact the ball,
but climbs up the back of the ball just prior to release, very much like a low
angle peel. The more flexible the face, the worse the situation becomes,
especially if the self-adhesive tape specimen is not firmly secured prior to the
test. Kaelble [65] showed that there is an area of compression just behind
the peel front. This results in automatic pressure application of the tape to
the test panel, no matter how minimal the initial application pressure. A
polyethylene (PE) film, unlike other film tapes, does not have a tack plateau
but continues to climb in tack value [3]. At maximum load the PE film con-
tributes to the value of tack obtained due to its extremely high extensibility.
In industrial practice the low adhesivity of special protective films
is the resultant of the simultaneous reduction of the coating weight and of
the thickness of the carrier film. Thus the coating weight can be decreased
without the loss of the peel and tack [60].
Tack Values. In adhesive practice there is a need for reference values
of the tack since the formulator and end-user have to compare the obtained
tack values with their target performance. This remains a difficult question
because tack values depend on the measurement methods and conditions,
the face stock material used, and the coating weight. Tack values obtained
for adhesives with a different chemical basis or physical state differ.
However, for those skilled in the formulation and use of PSAs, there are
some target values of the tack which need to be achieved. Table 6.4
summarizes some tack values for PSAs with different chemical bases and
measured with different methods.
Adhesive Performance Characteristics 251

Table 6.4 Tack Values for PSAs on Different Chemical Bases, Measured with
Different Methods

Chemical Basis Loop Tack (N/25 mm)


Rolling Ball
Base polymer Supplier Chemical nature Tack (cm) Glass PE

1360 Hoechst EVAc 17.0 6.9 3.4


FC88/PC80 UCB AC 7.0 13.1 7.6
FC88 UCB AC 7.5 13.3 8.0
V205 BASF AC 2.0 6.0 4.2
V208 BASF AC > 30.0 7.9 7.9
3703 Polysar CSBR 25.0 5.3 2.7
3703/V205 Polysar/BASF AC/CSBR 20.0 6.2 1.8
3703/V205 Polysar/BASF AC/CSBR 6.5 8.4 7.6
Tackified emulsions were used.

Improvement of the Tack


On the basis of theoretical considerations and practical experience one can
assume that hot-melt and solvent-based PSAs possess a higher tack than
water-based ones (cf. the lower molecular weight for the base polymers of
hot-melt PSAs and no poor tack additives for solvent-based PSAs); soft
base polymers (flexible, branched chain) yield better tack. Low molecular
weight polymers and uncrosslinked polymers (or with a low degree of
crosslinking) display higher tack levels. Using nonmigrating surfactants
(NMS) improvement in peel, tack, and shear were realized [66]. Latexes
prepared with NMS exhibited a loop tack increase from 1.67 to 4.0 (psi).
Since PSA bonding typically occurs at the plateau region of the viscoelastic
curve, bonding is facilitated by lowering the G0 at this region. G0 can be
effectively lowered by increasing the entanglement molecular weight of the
polymer; PSA debonding, however, happens at much higher frequency,
typically between 100–12,000 rad/sec. The loss modulus at this frequency is
in the transition region of the viscoelastic curve and its value is highly
dependent on the Tg of the polymer. Increasing the Tg results in higher G00
in this region. The tack properties will be improved. Increasing of the glass
transition temperature can be achieved by special branched comonomers,
e.g., vinyl neodecanoate and vinyl neododecanoate [67]. Tackified polymers
possess better tack as do plasticized polymers. Finally, higher coating
weights result in higher tack levels (see Fig. 6.4).
The coating weight also influences the peel adhesion and shear (see
Sections 1.2 and 1.3), as well as other characteristics like wettability (see
252 Chapter 6

Chap. 2). Therefore, before discussing other adhesive characteristics, a short


examination of the coating weight as a quality parameter is presented.

Role of Coating Weight and Parameters Influencing It


Benedek [68] summarized the influence of the coating weight on the PSA
label quality and its importance for the screening and testing of PSAs. The
coating weight influences the geometry and the quality of the PSA layer. Its
influence on the quality of the PSA layer is indirect, via drying [69]. The
coating weight influences the drying of the adhesive layer. The drying rate R
is given as a function of the mass transfer coefficient cm, the heat transfer
coefficient ch, and the latent heat of evaporation L [70] (i.e., as a function of
mass and heat transfer):

R ¼ f ðch  s  tÞ=L ¼ f ðcm  s  pÞ ð6:18Þ

where
t ¼ t  ts
p ¼ ps  p
ts ¼ surface temperature
ps ¼ vapor pressure at ts
p ¼ partial pressure
s ¼ surface area
For the drying of aqueous adhesive layers, one can assume that the
drying rate R also depends on the layer thickness h, concentration c, c, and
specific gravity [69]:

R ¼ ðh  t  cÞ=ð  L  hÞ ð6:19Þ

For solvent-based adhesives, the drying rate depends on the diffusivity


D, concentration c, and layer thickness h [69]:

R ¼   D  c=4h2 ð6:20Þ

Generally, the drying rate is inversely proportional to the layer


thickness (i.e., to the coating weight Cw):

R ¼ f ð1=Cw Þ ð6:21Þ

As shown in Fig. 6.5 for low coating weight values, the exponent is
smaller than 1, whereas for high coating weights, the exponent should be
larger than 1 (i.e., for drying PSAs with a high coating weight, the layer
Adhesive Performance Characteristics 253

Figure 6.5 Dependence of the drying speed on the coating weight. Weight loss (%)
as a function of the time, for different coating weights, and (wet adhesive thickness,
m: 1) 60; 2) 100; 3) 200) during drying of a water-based PSA.

thickness is the most important rate determining parameter). The


dependence between the wet coating weight, solids content, and drying
time was studied by Hansmann [70].
The strong influence of the coating weight on the drying has an
indirect effect on the adhesive characteristics. The equilibrium water content
of the paper and water-based PSAs layer depends on the drying degree (i.e.,
on the coating weight). On the other hand, residual humidity affects tack,
peel, and shear (see Section 1.2).
The direct influence of the coating weight on the peel and shear will be
discussed in Chap. 10. As a general statement, the coating weight has a
positive effect on tack and peel, and a negative one on the shear and drying.
In a first approach tack T and peel P are proportional to the coating weight,
shear S and the drying speed R are inversely proportional. Thus:

T ¼ f ðCw Þ ð6:22Þ

P ¼ f ðCw Þ ð6:23Þ
254 Chapter 6

S ¼ f ðCw1 Þ" ð6:24Þ

R ¼ f ðCw1 Þ ð6:25Þ

In summary the dependence of the adhesive properties Padh on the coating


weight (Cw) can be represented as follows:

Padh ¼ f ½ðCwþ Þ  ðCw"þ Þ1  ð6:26Þ

namely, the coating weight influences the versatility of the adhesive in a


complex manner. Evidently, the development of the PSA chemistry leads
to better adhesives (i.e., with a higher tack and peel adhesion) at a lower
coating weight.
The most important factors influencing the coating weight are the face
stock material, the substrate to be adhered to (adherent), the permanent/
removable character of the adhesive, and the coating conditions. In general
it can be stated that the optimum coating weight value depends on the
chemical composition and on the end-use of the PSAs. On the other hand,
the coating weight must be correlated to the uniformity of the coated layer
and to the coating (film-forming) quality, depending on the combination of
the following parameters:
Machine characteristics: speed and sense of rotation of the
metering cylinders, concentricity of the rolls, gap widths, properties
of rolls (e.g., surface finish, stability of material, deflection),
configuration, number and combination of rolls, hydraulic
pressure, and temperature control
Adhesive characteristics: cohesive strength, flow, and rheological
properties of the adhesive.

Chemical Composition and Coating Weight. According to their


chemical nature, synthesis, or formulation, different PSAs possess a different
adhesion-cohesion balance. Tack and peel may be improved by increasing
the coating weight. Thus, the chemical nature of the PSAs will influence the
required coating weight. On the other hand, as discussed in detail in [71]
regulation of the crosslinking degree allows changes in the peel and tack for
the same coating weight and adhesive.
Influence of Coating Conditions on the Coating Weight. The most
important parameter is the coating technology (i.e., the direct/indirect
nature of the coating operation). Direct coating supposes depositing the
Adhesive Performance Characteristics 255

adhesive on the face stock and thus adhesive bleeding through the porous
surface remains possible. Hence the measured coating weight could be lower
than the one being deposited. A detailed analysis of the influence of the
coating conditions (coating technology) on the coating weight will be given
in Section 2.2 of Chap. 9.
A lot of parameters influence the coating weight [68]. For laboratory
tests it is important to describe the dependence of the coating weight on the
knife opening (clearance) and viscosity. For a coating device with a blade,
the pressure difference p depends on the viscosity , width of the blade b,
web velocity Vw, and blade angle as follows [72]:

p ¼ ð4  Vw Þ=ðb  Þ ð6:27Þ

For the simplest case of a parallel blade, the blade angle may be replaced
by the blade opening (i.e., the distance between the blade and the cylinder).
In a first approximation a linear correlation is found between coating weight
and blade opening (Fig. 6.6).
On the other hand, the dependence of the coating weight on the
viscosity appears more complex. The coating weight or layer thickness Hw

Figure 6.6 The dependence of the coating weight on the clearance (blade opening).
Unfilled, water-based PSA formulation.
256 Chapter 6

for a Newtonian fluid depends on the surface tension ST, blade opening d,
viscosity , and specific gravity [73]:

Hw ¼ f ðST ,d,, Þ ð6:28Þ

There is a complex dependence of the coating weight on the viscosity


(Fig. 6.7). Unthickened dispersions behave quite differently from thickened
ones; thickened dispersions give a higher coating weight at higher viscosities
than unthickened ones (Fig. 6.8).
The viscosity also depends on fillers and pH; adjusting the coating
weight with fillers (viscosity) or through the pH also is possible. Table 6.5
demonstrates that filled dispersions do not display a linear dependence of
the coating weight on the blade opening.
Surfactants may change the viscosity of the aqueous dispersions and
hence the coating weight. Storage time and temperature also influence the
viscosity and thus the coating weight. In fact practical coating conditions
differ quite a bit from laboratory ones. Shear thinning and thickening and
flow under different centrifugal forces may influence the coating weight.
Laboratory coating conditions may be regarded as static whereas industrial
conditions are dynamic. Here the web velocity and shear conditions are the

Figure 6.7 The dependence of the coating weight on the viscosity. Water-based,
unthickened PSA formulation.
Adhesive Performance Characteristics 257

Figure 6.8 Coating weight dependence on the viscosity. Water-based, thickened


PSA formulation.

Table 6.5 Dependence of the Coating Weight on the


Clearance (Blade Opening) for Water-Based Acrylic
PSAs Containing Filler

Coating Weight (g/m2)

Filler content
(parts/100 parts, wet weight)

Clearance 0 5 10 20 30

60 25.5 20.0 33.0 44.0 29.0


100 43.0 36.0 40.0 38.0 45.0

most important parameters of the coating weight for a given adhesive and
coating device geometry.
Practically, during coating, in order to keep the coating weight
tolerance within maximum 5% the viscosity of the emulsion must be kept
at a constant level [74]. For an emulsion PSA on a reverse gravure coater,
it is ideal to have a high solids content with a low viscosity. Low viscosity
258 Chapter 6

means a high coating weight; conversely, high viscosity means a low coating
weight. The minimum level of the viscosity also depends on the surface
tension of the adhesive; common viscosities for reverse gravure range from
17–24 DIN sec.
Pressure-sensitive adhesive solutions or dispersions that have different
chemical bases may have a different solids content and viscosity. Thus, for
the same viscosity and wet coating weight, a quite different dry coating
weight and tack level may be obtained. This is illustrated by Fig. 6.9.
For aqueous dispersions the coating weight also depends on the
porosity of the web (paper). Therefore direct or transfer coating will result in
different coating weights (i.e., for the same coating weight the thickness of
the adhesive layer on the coated side will differ) (Table 6.6).
The obtained coating weight depends on the type and characteristics
of the metering device used. A gravure cylinder with a line screen of
40, 70 m depth, yields a coating weight of 5–7 g/m2 [75]. When coating
a 50%-solids adhesive with adequate viscosities, the rheology of the
system is such that to achieve a 3–4 g/m2 solid coating weight, a 35-m
gravure cell is suggested, while 2–2.2 g/m2 solids requires a 22-m deep
gravure cell.

Figure 6.9 Tack dependence on the coating weight. Rolling ball tack as a function
of the coating weight for water-based PSAs on different chemical bases: 1) CSBR
tackified; 2) CSBR; 3) acrylic.
Adhesive Performance Characteristics 259

Table 6.6 Dependence of the Coating Weight and Peel on the Coating Method

Coating Method

Peel on glass
PSA Coating Weight (g/m2) Direct Transfer (N/25 mm)

1 18.0  17.0
20.0  17.0 PT
2 19.5  21.0
21.0  21.0 PT
3 15.0  12.0
21.0  19.0 PT
4 10.0  7.0
15.0  9.0
Water-based acrylic PSA coated on paper face stock material. PT ¼ paper tear.

For a metering cylinder, the coating weight will depend on the blade
opening between blade and cylinder, the machine speed, and the film
transfer from the cylinder to the substrate [76]. Machine speed influences
the coating weight; higher velocities yield higher coating weights [77]. As
discussed in [60] the coating device affects the nominal coating weight value
also. For instance, using a Meyer Bar a lower effective coating weight of
1.8 g/m2 gives the same adhesivity obtained with the gravure cylinder and
2.4 g/m2 adhesive.
Influence of the End-Use of the Label on Coating Weight. As discussed
in detail in [71] the product class (e.g., label, tape, protective film, etc.)
influences the coating weight. Protective films are manufactured with 1–15 g/
m2 coating weight, labels use 15–25 g/m2 PSA, tapes have generally coating
weight values above 20 g/m2. For the same product class the end-use affects
the coating weight value also. The end-use of the label requires PSAs with a
permanent or removable character. On the other hand, the end-use of the
label determines its application and labeling technology. This also depends
on the adhesion/cohesion balance. Labeling guns require very tacky adhe-
sives; for decals a high tack is less important. Peel is a function of the
coating weight. In Chap. 2 it was shown that removability depends on peel
adhesion and on the coating weight. According to the permanent/removable
character of the designed label, one can use a higher or lower coating weight.
Mounting, insulating, and splicing tapes possess coating weight values
higher than 30 g/m2. Removable tapes like protective films are coated with
less than 10 g/m2 adhesive. The coating weight values used for different
pressure-sensitive products are listed in [71].
260 Chapter 6

Influence of the Face Material on the Coating Weight. The dry


coating weight (i.e., the weight of the dry adhesive applied per unit surface
area) can vary substantially depending upon the porosity and irregularity of
the face material and of the substrate surface to which the face stock is to be
adhered. For instance, higher adhesive loadings are preferred for adhering
porous, irregular ceramic tiles to porous surfaces, while lower adhesive
loadings are required to manufacture tapes, films, and other articles from
relatively nonporous smooth-surfaced materials such as synthetic polymer
films and sheets. When the adhesive is applied on nonporous polymeric or
metallic face materials intended for adhesion to nonporous polymeric or
metallic surfaces, adhesive loadings of about 8–85 g of dry adhesive per m2
are generally adequate. High adhesion in tapes manufactured from
continuous sheet polymeric substrates can usually be achieved with dry
adhesive coating weights of about 16–32 g/m2, while coating weights of
about 14–28 g/m2 of adhesive are usually employed for paper-backed tapes,
such as masking tapes. Textile carrier materials for labels or tapes need
coating weight values above 40 g/m2. The anchorage of the adhesive on
the carrier material influences the coating weight also. As discussed in [71]
primed removable products work with less coating weight. The conform-
ability and deformability of the face stock material affects the coating weight
also. For instance, soft and thin LDPE carrier for protective film requires
lower coating weight values than HDPE or PET.
Influence of the Substrate on Coating Weight. The adhesive flow is
influenced by the nature of the contact surface (liner as adherent). Chemical,
physical, and mechanical interactions with the solid state component may
hinder the adhesive flow. Therefore the chemical nature (e.g., affinity and
polarity) of the surface and the physical nature (e.g., porosity and
roughness) will influence the interaction of the adhesive with the substrate.
The adhesive flow influences the bonding and debonding characteristics.
According to Tordjeman [19] in probe tack tests the actual contact
area increases with the thickness and then applatizes. The contact area Ac is
given as a function of the contact force Fc and the apparent modulus Ea as
follows:

Ac  Nhai2 þ Fc =Ea ð6:29Þ

where the apparent modulus Ea takes into account the competition between
the thickness effects and the roughness hai:

Ea ¼ Eð1 þ hai2 =2 h2 Þ ð6:30Þ


Adhesive Performance Characteristics 261

The coating weight determines the real thickness of PSAs and the
thickness of the adhesive layer which can flow freely. Thus the coating
weight will depend on the nature of the substrate. As a practical example,
adhesion onto HDPE is generally lower than adhesion onto LDPE [78];
therefore the coating weight should be increased for adhesion to HDPE.
The influence of the carrier material and of the substrate on the peel will
be discussed in detail in the Section 1.2.

Dependence of Coating Weight on Adhesive Quality. The coating


weight may also depend on the quality of the adhesive. Under certain con-
ditions the emulsion polymerization of water-soluble monomers may yield
suspensions (i.e., the monomer is homopolymerizing) with bulky, non-
dispersed polymer particles remaining (coagulum, grit). This phenomenon
is especially encountered with acrylnitryl copolymers; Acronal 81 D may
contain so much coagulum that a coating weight of at least 100 g/m2 is
recommended (i.e., more than the diameter of the largest solid particle in
the fluid adhesive layer) [79].

Coating Weight Values. There are a lot of factors influencing the


value of the coating weight. On the other hand, the development of
the PSA chemistry has led to adhesives with improved tack and peel
adhesion (i.e., with lower coating weight). Some years ago coating weight
values of 30–100 g/m2 were recommended [80]; actually 15–20 g/m2 are more
common. Table 6.7 gives some typical values of the coating weight used
for labels on different face stock materials and for different end-uses.
Techniques for measuring the coating weight and tolerances are covered in
Chaps. 8, 9, and 10.

1.2 Peel Adhesion


Peel adhesion is the force required to remove a PSA-coated flexible material
from a specified test surface under standard conditions (e.g., specific angle
and rate). It gives a measure of adhesive or cohesive strength, depending on
the mode of failure [88]. Johnston [57] stated as an analogy that the adhesive
consists of many little springs, as depicted in the Maxwell and Voigt models
to simulate viscoelastic behavior, and that these springs range from strong
to weak. The former contribute to the adhesion (peel), the latter contribute to
tack. In fact, the measured peel energy includes a significant contribution
from energy dissipation within the bulk of the polymer. Thus for a given
extensible polymer, the higher the work of adhesion (due to the interaction
at the surface), the higher the dissipation energy (due to polymer
deformation) and the higher the peel energy [89]. As discussed earlier, in
262 Chapter 6

Table 6.7 Coating Weight Values of Industrial Labelstock

Adhesive characteristics

Nature Adhesive layer

Coating Coating Face


Chemical Physical weight (g/m2) thickness (m) stock Refs.

Acrylic Hot-melt — 25 Paper 46


Acrylic Water-based 25 — Paper 81
Acrylic Solvent-based 20 — Paper 81
— — 30 — — 82
— — 26–29 — — 83
Acrylic Water-based 27–29 — Film 2
Acrylic Water-based 25 — Paper 84
Acrylic Water-based 25 — Paper 85
CSBR Water-based 25–28 — Film 38
Acrylic Water-based — 20 Paper 86
Acrylic Hot-melt — 25–150 Paper 87
Acrylic Solvent-based/water-based — 25 — 45

Chap. 3, Section 3, based on contact physics and mechanics, debonding is


related to cavitation and fibrillation of the PSA, and its modulus.

Measurement of the Peel Force


Peel is measured as the force required to remove a PSA-coated material
which was applied to a standard test plate, at a specific rate and angle. In
some cases not only the value of the peel force, but also the failure mode
(deformation or destruction of the components of the pressure-sensitive
laminate or of the substrate) has to be examined in order to obtain a correct
appraisal of the removability of PSAs.
The results of a peel test are very useful in comparing different
adhesives, but due to a complicated stress distribution, they lack the ability
to quantitatively investigate the molecular origins of adhesion. For this
reason asymmetric adhesion tests are favored in studying the interfacial and
bulk contributions to adhesion. Due to the disadvantages of the flat punch
(alignment difficulty and fibrillation upon pull-off ) Crosby and Shull [30]
use the spherical geometry. This test provides adequate data to generate tack
curves. Pocius et al. [90] studied the release behavior of acrylic tapes from
release coating based upon long alkyl side-chain polycarbamates by the
Adhesive Performance Characteristics 263

methodology of contact mechanics and peel strength data (90 peel of PSA
tapes measured on glass plate coated with release). The extrapolation of the
contact mechanics data (at low rate) matches the peel data (at high rate).
They studied the adhesion energy as a plot of crack speed (nm/s). On the
basis of fibrillation, the debonding force is computed by Verdier and Piau
[91] using the following correlation:

G ¼ F=l ð6:31Þ

where F is the force and l is the bond width.


Z "max
G ¼ ½a ðL  eÞ= "max  ð","0 ,TÞ d" ð6:32Þ

where "max is the maximum elongation of the filament, l is the periodicity of


the filaments, a is the filament width, L is the filament height, and e is the
adhesive thickness. The elongation rate ("0 ) is taken to be V/e;  is the stress
given by Lodge’s model. According to Lin and Hui [15] the fibrillated zone
length is determined by the fracture criterion h ¼ hc and s ¼ so. The
dependence of the normalized peel velocity V/E*A on the normalized peel
force (F/E*I1/3) is shown for different values of normalized hc; normalized
peel velocity is proportional to normalized peel force; A is the rate constant
associated with fibril deformation. The normalized peel height (L/2I1/3) is
plotted against normalized peel force for different hc/(2I )1/3 values. The peel
height decreases with increasing peel force. In addition the dependence of
the normalized peel height on the normalized peel force is independent
of the critical fibril length. For a given peel force the fibrillated zone size
increases with critical fibril failure length.
According to [92] the peel force P relates to tensile strain " and stress 
by the following equation:
Z "
P ¼ ðbl=2Þ ð"Þ d" ð6:33Þ
0

where b is the width and l is the thickness of the adhesive film. The specific
work of PSA film tensile straining (A) is identical to the work of adhesion:
Z
A ¼ ð"Þ d" ð6:34Þ

As follows from Eqs. (6.33) and (6.34), the mode of adhesive joint
failure is determined by the ratio between the two limiting tensile stress
264 Chapter 6

values  a (which characterizes the strength of the adhesive bond) and


 c) which relates to the cohesive strength of the PSA polymer). For  a >  c
the failure is cohesive, whereas an adhesive type of debonding occurs at
 c >  a. The slip-stick point position corresponds to the condition of
 a   c. (Slip-stick is the phenomenon when the adhesive separation front
can move faster than the rate of testing.) In the case of interfacial failure
the peel curve is noisy but approximately constant [93]. According to [94]
when failure is cohesive, the peel force is rather steady; for uncrosslinked
PSA the peel force drops when the mode of failure changes from cohesive to
adhesive [95].
The 180 peel test is a combination of tensile strength and shear, while

a 90 peel is of tensile strength only [96]. Approximating the deformation
of adhesive in the peel front by uniaxial elongation, the peel force is given
by integration of the tension across the peel front [97]:
Z xb
P¼ o ðxÞ dx ð6:35Þ
0

where P is the peel force per peel strip width,  o is the tension at the
adhesive/substrate interface, x is the horizontal position in the peel front,
and xb is the position at which the adhesive detaches or breaks cohesively.
Christensen [97] compares the experimental and theoretical peel forces and
demonstrates the validity of Eq. (6.35) and of the assumption of uniaxial
deformation.
Kendall [98] compares the adhesion force for different kinds of
adhesive joints and debonding tests, e.g., peel test (with flexible carrier
materials at large angle; with rigid carrier-disk sample; with semiflexible
carrier Obreimoff [99] geometry for cleavage) lap joint failure, and
small angle peel. As stated by Johnson [100] for contact of elastic
spheres, the elastic modulus (E ) does not affect the pull force (F ) (see
Chap. 3, Section 3). According to Kendall [98] the sphere adhesion result is
rather similar to that of the peel test. In the peel test the shape, and
therefore the strain energy, remains constant for an elastic material, the
adhesion force turns out to be independent of the elastic modulus just as
in the JKR (Johnson–Kendall–Roberts) solution. For a disk sample (rigid
cleavage) the force increases in proportion to (EW)1/2, the interface
toughness. This supports the idea that stiffening adhesive joints usually
leads to more strength. That means that by cleavage peel the force depends
on the E:

F ¼ f ðEWÞ1=2 ð6:36Þ
Adhesive Performance Characteristics 265

Using Obreimoff geometry, to pull a cantilever strip from the surface the
force is given by the following correlation:

F ¼ bðWEdÞ1=2 ðd 2 =6c2 Þ1=2 ð6:37Þ

where b is the width and c is the length of the joint. For the lap joint the
strength increases with sheet thickness (d ) and E.

F ¼ bð4WEdÞ1=2 ð6:38Þ

For low angle peeling:

F ¼ bð2WEdÞ1=2 ð6:39Þ

which is the equation for lap failure of a flexible film in contact with a rigid
surface. That means that E does not influence F for 180 peel but it
influences it for other kinds of debonding. The strain energy release rate G,
during debonding of a cantilever beam sample, containing at its midline a
thin PSA layer, was utilized to quantify the adhesion of PSA by Taub and
Dauskardt [101]. One might expect the value of the adhesive fracture energy
(Gc) from the peel tests to be significantly lower than that determined from
CDB (Double Cantilever Beam) specimens, as was observed in the case of
values of Gc deduced using the FEA-CZM (cohesive zone model) analysis of
the peel tests [102].

Factors Influencing Peel Adhesion


Bonding plays an important role in the peel resistance. Parameters
influencing the adhesive flow during the bonding step will influence the
peel adhesion. In a manner similar to tack, peel is influenced by the PSA
nature and geometry, and by the nature of the face stock material. The
influence of the adhesive thickness, viscosity, backing rigidity, and peeling
velocity in the cohesive regime was stated [103]. Test methods and
conditions modify the peel value as well. More pronounced for the peel
force are the influence of the dwell time and of the adherent.
Influence of the Adhesive’s Nature on Peel. Similar to tack, peel
resistance is a function of the chemical nature and macromolecular
characteristics of the base polymers. On the other hand, and in contrast
to tack, peel adhesion increases with the cohesion (i.e., with the molecular
weight) up to a limit. A quite different adhesion/cohesion balance is required
for good peel adhesion onto polar or nonpolar substrates. Regarding the
influence of the adhesive’s nature, one should differentiate between the
266 Chapter 6

chemical composition of the base elastomers, the chemical composition of


the formulated adhesive, and the technology dependent composition of the
coated PSAs; one should also note the influence of aging, the chemical
composition of the PSA, and solid state components of the laminate,
because of the environmental and reciprocal interactions between the
adhesive and the neighboring components (e.g., face stock and release liner).

Influence of the Chemical Composition of the Base Elastomer on the


Peel. The anchorage of the adhesive on the substrate is governed by a
physical and a chemical component. Chemical interactions between the adhe-
sive and the substrate result in adhesion build-up; the chemical composition
of the base elastomers strongly influences this adhesion build-up. Small
portions of highly polar monomers (e.g., acrylic acid) produce adhesion
build-up. On the other hand, PSAs with a vinyl polymeric backbone and with
grafted polysiloxane moieties produce initially repositionable labels [104].
The inclusion of polar carboxylic groups in the adhesive may influence
its peel value and change its affinity towards polar surfaces [105]. In a similar
manner such groups influence the peel of the adhesive through water
sensitivity. Built-in carboxyl groups may produce a change in the peel value
from special substrates through pH changes [106]. Introducing polar,
functional groups into ethylene vinyl acetate copolymers improves their
adhesion [107]. The role of the carboxyl sticker groups on the peel adhesion
strength (G) of linear polybutadiene to aluminum surfaces coated with
amine and other receptor groups was investigated in [108]; G went though a
maximum at some optimal concentration of the sticker groups (ca. 3 mol %).
According to [109] the acrylic acid content dramatically increases adhesion
and adhesion hysteresis of acrylic copolymers. The addition of acrylic acid
also decreases a critical crack propagation speed above which a much
stronger rate dependence of the adhesion energy is observed. The contri-
bution to adhesion strength by these groups is evaluated by surface energy
measurements and by peel strength measurements (energy dissipation due to
polymer deformation).
Earlier studies [110–112] have shown that the modulus G0 value at low
frequencies can be related to the wetting and creep behavior (bonding) of the
adhesive and the modulus at high frequencies can be related to the peel or
quick stick (debonding). A general correlation was developed between tape
properties (release and peel adhesion). Good adhesion for tapes is given
when G0 is low at low frequency rates, i.e., 0.1 rad/sec and a relatively high
slope exists for G0 as the frequency is increased [113]. The peel strength is
proportional to the ratio of G00 and G0 at the respective debonding and
bonding frequencies. Gordon et al. [114–116] presented an empirical corre-
lation between the bulk viscoelastic properties of commercial adhesive and
Adhesive Performance Characteristics 267

the release profile observed from peel force vs. peel velocity measurements.
It was shown that the bulk viscoelastic properties of the adhesive play a
dominant role in controlling release profiles in a 180 peel test. The general
shape of the release profile was consistent with the shape of the adhesive’s
viscoelastic function—specifically the loss tangent (tan delta) as a function
of the dynamic frequency—in the peel window previously defined:

! ¼ 2=h ð6:40Þ

where ! is the angular frequency in the viscoelastic test,  is the delamination


velocity in the peel test, and h is the thickness of the deformed region of
the viscoelastic material. It was evident, that the viscoelastic properties of
the release influence the magnitude of the release force at any given peel
velocity. A power law relationship was observed in coating formulations in
which the linear PDMS chain segments were greater than the entanglement
limit [117]. Adhesives with lower creep compliance (higher G0 ) will have
better cohesive strength (creep resistance). However, if the modulus is too
high, the adhesive may not rapidly wet the substrate. The real area of
contact of an elastic film on a rough surface depends on two parameters: the
elastic modulus at the bonding frequency (G0 ) and the relaxation properties
of the polymer which govern the change of the real area of contact with time
for a given imposed displacement. Increasing the modulus G0 results in
adhesives with higher adhesion and lower tack. In the correlation proposed
by Bikermann [118] for the energy release rate, the Young moduli of the
backing and adhesive respectively are included. According to Fukuzawa and
Uekita [119] the increase of the elastic modulus of the PSA caused a decrease
of the peel adhesion. As discussed in Chaps. 2 and 3 the rheological
properties depend on the chemical composition and molecular weight of the
base polymers.

Influence of the Chemical Composition of the Formulated Adhesive


on Peel
The effect of the formulating additives on the peel will be reviewed in
Section 2.9 of Chap. 8. Some of these influences are caused by the chemical
interaction of the viscous and elastic components of the formulation (e.g.,
elastomer, tackifier, plasticizer) and exist in a freshly coated adhesive
layer, but others are due to the interaction of adhesive/environment or
adhesive/laminate components. Changes in the peel value given by adhesive/
environment interactions may be irreversible (aging) or reversible, such as
the influence of the atmospheric humidity on the peel of tackified (with
aqueous resin dispersions) or untackified water-based PSAs.
268 Chapter 6

Tackification remains the most pronounced formulating effect on the


peel. Peel adjustment through tackification will be discussed in a detailed
manner in Chap. 8. Here it should be mentioned that for removable (low
peel adhesives) no tackifier or a low level of tackifier should be used. More
than 50% tackifier will make the adhesive nonpeelable; fillers also influence
the peel adhesion. Generally, they improve the modulus, but decrease the
contact area and diffusion rate. Therefore over 10–15% filler concentration
(on wet/wet basis) decreases the peel [106]. In hot-melt PSA formulations
peel values sharply decrease with the use of fully saturated mineral oils [120].

Effect of Crosslinking on Peel Adhesion. The 180 peel strength of


PSA formulations based on crosslinkable rubber (e.g., SIS, electron beam-
curable) remain fairly constant over the entire range of electron beam doses
[50]. Yarusso et al. [95] studied as a model PSA, anionic polyisoprene. The
effects of the initial molecular weight and of the degree of crosslinking were
studied. As stated, high gel content could compensate for low MW. This was
not true for peel behavior. Three different failure modes were observed
in the adhesion testing. Adhesion failure with no apparent change in the
appearance of the stainless steel plate was favored by high molecular weight,
high gel content, and high peel rate. Cohesive failure leaving tacky adhesive
both on the tape and on the panel was favored by low molecular weight, low
gel content, and low peel speed. When the molecular weight was low, but the
gel content was high, and the peel speed was low, we observed a very thin
residue on the stainless steel plate, which was often of a bluish or brownish
color and was nontacky. Generally, crosslinking exerts a more complex
influence on the peel. A low degree of crosslinking improves the cohesion
and thus the peel, whereas a high crosslinking degree lowers the tack and the
peel. As an example one can consider the increase of the peel for Cymel
crosslinked neoprene latex and the dramatic reduction of the peel upon
adding zinc oxide to the formulation [52].
For the crosslinked adhesive more complex three-dimensional flow
patterns are obtained by peeling. At very high velocities stick release is
observed [91]. Peel data for moderately crosslinked, epoxidized natural
rubber (ENR) are shown to be time-temperature superposable, while those
for a lightly crosslinked ENR are not. Strain crystallization, because of the
resultant hysteresis, could provide the additional dissipative mechanism and
hence the anomalously high peel strength at elevated temperature for the
lightly crosslinked ENR [121].
Tobing [122] has shown that for gel free, low-Tg acrylics (based on
2-EHA) there was no significant difference between emulsion vs. solvent-
borne PSA film adhesive performances. However, commercial products
have a high level of gel and low shear. Interlinking of the microgels
Adhesive Performance Characteristics 269

by covalent bonds in the film (using a functional monomer, e.g., IBMA) is


needed. In special acrylic emulsions 0.5–1.0 wt% IBMA (isobutoxymethy-
lacrylamide) was copolymerized to provide interlinking of the microgels in
the film upon heating. When the MW of the linear polymer was greater than
2Me and the Mc of the microgels was greater than Me, a synergistic effect
was obtained.
Influence of the Molecular Weight of the Adhesive on the Peel. The
influence of the molecular weight of the polymer components of the PSAs
on the rheology of the PSAs was examined in Chap. 3. The rheology of the
adhesive influences its adhesive characteristics; therefore, the peel also is a
function of the molecular weight [123]. The increase of the modulus with the
molecular weight reinforces the peel; on the other hand, it reduces the chain
mobility and diffusion rate, and decreases the peel [124]. Physical adhesion
can develop from adequate wetting at a well-defined interface or by
diffusional interpenetration of segments across the interface, when this is
thermodynamically possible. The fracture energy G of a joint follows a
relation of the form [125]:

G ¼ f ðD  tÞa ð6:41Þ

where D is the diffusion coefficient, which is inversely proportional to the


molecular weight, and t is the time. For adhesive formulations tackified with
butyl rubber, peel adhesion increases with the molecular weight of the
tackifier. On the other hand, when tackifying CSBR, higher melting point
resins decrease the adhesion on polyethylene. The tendency to transfer
adhesive or residue to the substrate is especially sensitive to the initial
molecular weight even at high gel content.
The dispersion of the molecular weight plays an important role in
debonding also. Commercial acrylic adhesives, like rubber very broad
spectrum of relaxation times. And in particular low MW fractions provide
the fast relaxation times necessary to achieve a large real contact area within
a short contact time (typically 1 sec). In the case of monodisperse polymers
these low MW fractions are absent and the contact area depends mostly on
the elastic properties of the polymer at bonding frequency. The distinctive
feature of these linear monodisperse samples is that the transition from
cohesive to adhesive failure with increasing strain rate always corresponded
to a maximum on the value of the "max vs. strain rate [126]. However, the
strain rate at which the cohesive-adhesive transition and the maximum in
"max occurred increased when the molecular weight decreased. If failure
proceeds through disentanglement of the polymer chain, both the fibril
stress  fibril and "max increase continuously as VdebM3 increases because
270 Chapter 6

disentanglement becomes more difficult. If the failure mechanism is not


disentanglement but a debonding process the critical stress  sustainable
by the fibril-punch interface is the rupture criterion. On expects  r to be
constant so in this regime "max decreases with VdebM3 since: "max /  r/
VdebM3. The two limiting cases are interfacial dewetting (Vdeb <<1) and
brittle interfacial fracture (Vdeb >> 1). Experimental results have shown
that a large spectrum of relaxation times is a necessary condition for the
performance of PSAs. If short relaxation times are absent (i.e., no low MW
fraction) the contact with the substrate is not achieved within short contact
times, while if long relaxation times are absent, a fibrillar and adhesive
fracture cannot be achieved.

Influence of the Adhesive Geometry on the Peel Adhesion. Adhesive


geometry includes the dimensions and form of the adhesive layer; the former
depend on the coating weight. The coating weight strongly influences the
value of the debonding force. In general peel adhesion is a function of the
thickness of the adhesive [127] and, like tack, peel is very sensitive towards
the coating weight. Generally, at low coating weights, the peel increases with
the coating weight, implying a debonding through adhesive failure.
Concerning the influence of the coating weight on the peel resistance,
the peel force P depends on the peel stress s, sample width b, layer thickness
h, and creep modulus Ec as follows [105]:

P ¼ ðs2  b  hÞ=Ec ð6:42Þ

As can be seen in Fig. 6.10 there exists a complex dependence of the peel on
the coating weight; generally there is an inflection point in the plot. Up to
this point peel increases rapidly with the coating weight; after this point peel
depends less on the coating weight. The shape of the plot depends on the
adhesive and face material. In order to avoid quality consistency deviations
caused by a variation of the coating weight, the coater should use a coating
weight above this critical value. It should also be mentioned that in the
upper domain (over the critical value), the peel dependence on the coating
weight is less dependent on the face stock. Generally, the dependence of the
peel on the coating weight may be formulated as follows:

P ¼ f ðCwb Þ ð6:43Þ

where b 6¼ 1 and possesses quite different values in the over and under critical
domains. As discussed in Chap. 2 for given application conditions, surface
roughness, and carrier deformability, the dependence of the peel force on
Adhesive Performance Characteristics 271

Figure 6.10 Dependence of the peel on the coating weight. 1) Peel on


polyethylene; 2) peel on glass.

the coating weight can be described by a simplified plot [see Eq. (2.43)],
which is a linear dependence (like the Kaelble equation), valid over the
critical coating weight.
Taub and Dauskardt [128] shifted the so-called R curves—energy
release rate G (J/m2) ¼ f (PSA thickness)—towards lower values of G with
increasing layer thickness (in the range of 30–250 m) until a critical layer
thickness is obtained. For thicknesses greater than this critical thickness, the
R curves become relatively insensitive to changes in the layer thickness. This
insensitivity to adhesive layer thickness is consistent with data reported in
the literature for cohesive failure in PSA peel tests [129].
By the analysis of PSA fracture mechanics in the course of peeling,
Chalykh et al. [92] established that the peel force P relates to tensile strain
" and stress  by an equation which includes b (the width) and l (the
thickness) of the adhesive film. The peel force P, to rupture the adhesive
bond obeys the modified Kaelble equation (3.15) [130,131].
Bikermann [118] also proposed a derivation of the energy release rate
as follows:

F ¼ 0:3799beð1=4Þ ð2hÞð3=4Þ ðE=Þð1=4Þ ð6:44Þ


272 Chapter 6

where F is the force (F ¼ bG), b is the adhesive width,  is the rupture stress,
e is the adhesive thickness, 2h is the backing thickness, and E and Y are the
Young moduli of the backing and adhesive respectively.
The adhesion energy calculated from indenter tests is given by the
follo-wing correlation [18]:
Z
Wadh ¼ ho ð"Þd" ð6:45Þ

where ho is the initial film thickness,  the nominal stress, and " the strain.
As discussed earlier, tack also is measured as peel adhesion. Tack
increases with the coating weight. Figure 6.11 illustrates the simultaneous
increase of tack and peel with the coating weight; however, if the debonding
occurs by break in the adhesive layer (cohesive failure) the peel does
not depend on the coating weight [132]. The peel strength depends on
the thickness of the adhesive layer and of the face stock material.
Peel increases with increasing thickness of the adhesive. This increase of

Figure 6.11 The influence of the coating weight on the tack and peel. The
dependence 1) of the rolling ball tack and 2) of the peel on polyethylene on the
coating weight for a water-based acrylic PSA formulation.
Adhesive Performance Characteristics 273

Table 6.8 Dependence of the Peel on the Thickness of


the Adhesive Layer

Peel (N/25 mm)


Wet Adhesive Coating Weight
2
Layer Thickness (m) (g/m ) PVC Paper

60 22 7.3 —
24 — 5.6
70 24 5.8 —
25 — 5.5
100 43 3.3 —
46 — 3.2
130 46 2.5 —
47 — 3.8
160 56 3.3 —
67 — 4.2
A 200-mPa.s EVAc water-based PSA was coated on soft PVC via
an 80 g/m2 paper (transfer coating).

the peel with the adhesive thickness is illustrated in Table 6.8. The influence
of the thickness of the adhesive layer on the peel was also studied by
Schlimmer [133], who found that peel adhesion does not depend on the layer
thickness.
The use of large amounts of adhesive leads to a relatively strong bond
between the adhesive and the target substrate, thus making the product less
peelable. Generally, one finds that lower coating weights make peelability
of the product in use easier to achieve, but make a uniform coating more
difficult to obtain. For paper coating with acrylic PSAs the coating weight
amounts to at least 12 g/m2, and usually 14–20 g/m2 depending on the nature
of the adhesive. The influence of the coating weight on the peel was studied
by several authors [3,123,134]. The 90 peel adhesion test (FTM 2) results
in a higher (usually more than double) peel adhesion from PE than the
180 peel test [47] (see Section 3.1 of Chap. 10). For the 90 peel a nonlinear
dependence of the peel on the coating weight was established; for 180 peel
adhesion a linear dependence of the peel force on the coating weight is given.
Peel dependence on the coating weight also is a function of the
substrate, face stock, and adhesive. For silicone-based adhesives, in most
cases a coating weight-independent adhesion value on steel is obtained at
coating weights over 4–6 g/m2. Table 6.9 shows that peel adhesion increases
with the coating weight; after a certain value of the coating weight (which
depends on the adhesive and substrate) peel adhesion increases at a slower
rate and eventually reaches a plateau.
274 Chapter 6

The existence of a critical coating weight value plays an important role


in the end-use of pressure-sensitive products with low nominal coating
weight, e.g., protective films, where the coating weight may attain the critical
value. Figure 6.12 illustrates the dependence of the peel resistance on the
coating weight for two different adhesives. At low nominal coating weight
values (0.5–1.5 g/m2) changes of 25–35% of the coating weight lead to
pronounced (50–60%) reduction of the peel value. As seen, the reduction

Table 6.9 Dependence of the Peel on the


Coating Weight (Peel on Glass and Polyethylene
as a Function of the Coating Weight)

Peel (N/25 mm)


Coating
2 Weight
(g/m ) Glass PE

9.4 17.0 PT 15.0


18.8 17.0 PT 16.0
33.5 22.0 PT 14.0
39.6 25.0 17.0
61.8 25.0 19.0
A tackified water-based acrylic PSA was used.
PT ¼ Paper tear.

Figure 6.12 Dependence of the peel resistance on the coating weight and adhesive
nature: (1) and (2) are different adhesive grades.
Adhesive Performance Characteristics 275

of the coating weight decreases the peel resistance and below a critical
coating weight value (which is adhesive and substrate dependent) joint
failure may occur.
Dependence of the Peel on the Shape of the Adhesive Layer. In
general, defects in an applied continuous PSA layer should be avoided for
common adhesive coatings. However, there are special coatings with a
discontinuous character of the adhesive layer, mostly in order to avoid the
build-up of high peel resistance. On the other hand, the structure of the
continuous adhesive layer also influences the peel resistance [132]. As
discussed in [60] the coating device affects the shape of the coated adhesive
layer. Crosslinked adhesives coated with a gravure cylinder give peel values
which differ from those obtained with a Meyer Bar. Regulating the shape of
the PSA allows the control of the adhesion for removable adhesives too
[135]. See Chap. 2 for a more detailed discussion of the influence of the
adhesive form on the peel.
Influence of Coating Technology on Peel. Direct coating achieves
better anchorage of the PSAs on the face stock, and therefore better
removability. For optimum anchorage, sheet-stock PSAs are often directly
coated on to the face material, rather than transfer coated via the release
liner. The state of the adhesive during the coating operation also influences
the peel. Water-based PSAs contain different water-sensitive and hygro-
scopic components; hence they may contain less or more humidity than the
laminate. The residual water content acts like a plasticizer and reduces the
peel [78]. Hyde and Yarusso [136,137] discovered that under certain hot-
melt coating conditions PSA blends of NR and miscible aliphatic tackifying
resins can also undergo strain-induced crystallization. This phenomenon
would not be observed in coatings prepared from solvent- or water-based
adhesives. In this case the molecular orientation of the rubber phase
imparted by the strain-induced crystallization process leads to a reduction of
tack and an anisotropy in the peel adhesion performance. The peel adhesion
is substantially lower in the coating machine direction compared to the
transverse direction. This property allows for repositioning of the tape.
Influence of Face Stock Material on the Peel. The rheology of the PSA
depends on the solid state components of the laminate. Peel is a function of
the rheology of the adhesive and it also depends on the face stock material.
Both the bulk properties of the face stock material and its surface quality
affect the peel resistance. Peel is determined by the deformability not only of
the adhesive, but also of the face stock [138]; the stiffness (flexural modulus)
of the face stock influences the peel angle and thus the peel value [132]. The
effect of the face material becomes part of the result; as an example one can
276 Chapter 6

consider a 180 peel adhesion test, where work is being done not only to
separate the adhesive from the applied surface but also to bend the face stock
through 180 [3]. The same adhesive at the same coating weight can display
a variety of peel values when coated onto different face stock materials.
Depending on the substrate, adhesives may be classified as removable or
permanent. The same substrate material may have different surface proper-
ties; for example, yellow plastic items may be labeled less easily than white
ones because of the fillers used [139].
The bulk characteristics of the face stock material, its stiffness, and
plasticity/elasticity affect the transfer of the debonding forces (see Chap. 2).
On the other hand, the chemical composition of the face stock material
influences the interaction between the adhesive and the face stock material.
This influence should be taken into account especially for PVC foils [132].
The geometry and dimensions of the face stock can affect the peel resistance
as well. Mechanical effects in the peel test were studied by Kim et al. [140].
The peel strength measured by the peel test method is a practical adhesion
(an engineering strength per unit width) and does not represent the true inter-
facial adhesion. The measured value is a combination of the true interfacial
adhesion strength and forces required for plastic and elastic deformation of
the adhesive film and substrate. The major controlling factors in the peel
strength are thickness, Young’s modulus, yield strength, the strain harden-
ing coefficient of the adhesive film, compliance of the face material, and
interfacial adhesion strength. The mechanical characteristics of the face
stock material, its flexibility, and elasticity (plasticity) influence the transfer
of the debonding energy to the adhesive layer (see Chap. 2).
The decrease or damping of the peel force is caused by the deforma-
tion (elongation) of the face stock material under tension, changing the
direction or distribution of the force (change of the debonding angle), and
by the stiffness of the material; these phenomena decrease the peel force [3].
Thus, one can write:

Peel ¼ f ½ðStiffnessÞ1 , ðPlasticityÞ1  ð6:46Þ

The stiffness depends on the flexural modulus Ef whereas elongation


depends on the elongation modulus Ee of the face stock. Thus:

Peel ¼ f ½1=ðEf ,Ee Þ ð6:47Þ

This equation is confirmed by practical data from the labeling industry.


Stiff aluminum foils used as face stock material exhibit a lower peel value.
Similarly, opaque soft PVC gives a lower peel value than clear (softer) PVC.
Adhesive Performance Characteristics 277

For some adhesives, in spite of the weak adhesion, there is a possibility


of achieving high peel strength values if the adhesive film exhibits a low
modulus [141]; this means that:

Peel ¼ f ðEf ,Ee1 Þ ð6:48Þ

According to [142] a disadvantage of the peel testing is the significant


plastic deformation that the peel arm undergoes. The extent of plastic defor-
mation is a function of peel angle and peel rate and therefore the measured
Gc (fracture energy) values are said to be geometry dependent.
Bikermann [118] proposed a derivation of the energy release rate see
[Eq. (6.44)] where h (the backing thickness) and Y (the Young moduli of the
backing) take into account the influence of the carrier material.
In the formula discussed by Kaelble [143] the factor Eh3 takes into
account the backing rigidity. Piau et al. [103] stated that for steel as carrier,
lower peel and lower increase of the peel with the velocity. In the peeling
master curve they represent the quantity
3/7G as a function of V, where

¼ Ve2/(Eh3) is a dimensionless number representing the effect of the


adhesive viscous forces compared to the rigidity effects of the backing.
Fukuzawa and Uekita [119] suggested a theoretical formula of peel
adhesion (P):

P ¼ b½Wa þAX 2 ðeB=X þ B=X  1Þ


ð6:49Þ
þ 2Es Is ð!  sin !Þ=ð1  cos !Þ=ð1  cos !Þ

which includes in its parameters A and B the work of adhesion Wa, between
the PSA and the adherend, the work of deformation of the PSA (Wd), and
the work of bending of the backing Wb. The modulus of the PSA (Ea), the
modulus of the backing Es, and the moment of inertia of the backing Is
(Is ¼ bt3s /12; where b is the width of the PSA tape, ts is the thickness of
the backing) are taken into account also. An increase in the modulus of the
backing causes an increase in the peel adhesion. If the modulus of
the backing attains 39  E 9 dyn/cm, peel increases.
Fracture energy values measured from the SLBT were compared to
values obtained using the pull-off test (and were of the same order of
magnitude for different substrates). However, for high energy substrates
results from a 90 peel test were an order of magnitude greater than the
SLBT, probably due to the plastic deformation of the peel arm [142]. For
thin films the effects of plastic deformation and yielding may be significant
given the small load bearing capacity, associated with the small thickness,
and the relatively strong adhesion [144].
278 Chapter 6

According to Williams et al. [145] the current challenge is to model


accurately any extensive plastic deformation which may occur in the flexible
peeling arm, since if this is not accurately modeled then the value of Gc
deduced may suffer a high degree of error. Since the cover sheet width is
much greater than its thickness, the deformation during peeling is
independent of the out-of-plane coordinate. The cover sheet is assumed to
be linearly elastic with Young modulus E and Poisson’s ratio  respectively.
Plastic yielding of the cover sheet will not occur as long as it is sufficiently
thick [146], so that:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
6E FR=H <  ð6:50Þ

where F is the peel force per unit width,  is the yield stress of the cover sheet
in shear, and E* ¼ E /(1–2). The cover sheet is modeled as elastic, so that:
M ¼ E*Ik, where I ¼ H3/12, k is the radius of curvature of the cover sheet at
P  (x, y), and M is the moment at P [15].
As discussed above, reduction of the coating weight to values under
the critical weight causes joint failure. In principle such reduction may be the
result of carrier (and adhesive) deformation also [147].
Comparative industrial experiments with protective films having
carrier films with standard and reduced thicknesses for the same adhesive
and adhesive coating weight, confirm the decrease of the peel resistance with
the reduction of the thickness of the carrier film too (Table 6.10). Films
with strongly reduced thickness show very low adhesion. The mean peel

Table 6.10 The Influence of the Carrier Thickness


on the Peel Resistance of Adhesive Coated Films

Peel Resistance
Adhesive Carrier on Stainless Steel
Code Thickness (m) 180 (N/10 mm)

1 50 0.15
25 0.05
2 80 2.50
40 0.90
3 80 2.50
45 0.80
4 45 1.10
35 0.80
5 40 0.40
30 0.01
Adhesive Performance Characteristics 279

resistance value of about 0.6 N/10 mm is reduced to 0.1–0.06 N/10 mm. That
means that such films, although they possess a standard adhesive coating
weight, do not work like standard products. The adhesion of such protective
films with very thin carrier material is extremely low and does not allow
their standard use. Tensile stresses in the carrier film may produce bond
failure. The deformability of the carrier (reduction of its thickness
by stretching) depends on its thickness. The deformability of the carrier
produces two effects. In general, reduction of the carrier thickness causes
the decrease of the peel force. In the extreme case, an excessive reduction of
the carrier thickness leads to bond failure. Astonishingly this phenomenon
occurs at a carrier thickness which is sufficient for self-supporting or
processing of the laminate. That means, that the pressure-sensitive laminate
cannot be used although its components (the adhesive and the carrier) work.
The adhesive film can be laminated on the product to be protected and the
adhesion between the film and substrate is sufficient if no in-plane tensioning
of the carrier appears, but stretching of the film causes debonding. Such
pressure-sensitive products with very low (dynamically tested) adhesion
behave under statical use adequately. However, their end-use quality cannot
be evaluated by means of standard tests. As stated, bond failure appears at
a critical carrier deformability (Fig. 6.13). It is supposed that in this
case the simultaneous, carrier-induced deformation of the adhesive leads to

Figure 6.13 Dependence of the peel resistance on the maximum tensile force (Fm)
in the carrier film.
280 Chapter 6

an apparent reduction of the coating weight which causes bond failure


(Fig. 6.14). For thin plastic films the maximum tensile strength (Fm) is
preferred as an index of the carrier deformability. Thus, the peel resistance
can be described as a function of the coating weight (Cw) and the parameter
(Fm) of the carrier deformability which affects the coating weight according
to the following correlation:

P ¼ Cw Fm ð6:51Þ

The dependence of the peel resistance on the carrier thickness is given in a


general form by the following correlation:

P ¼ f ðFm  k2 Þ ð6:52Þ

where the coefficient k2 takes into account the minimum film thickness
required for dimensional stability. The correlation between adhesive coating
weight, carrier strength, and peel resistance can be written in the following
form:

P ¼ f ½ðCw  k1 Þ,ðFm  k2 Þ ð6:53Þ

Figure 6.14 Bond failure and deformation of the removable, adhesive coated,
thin plastic carrier.
Adhesive Performance Characteristics 281

There is a critical deformability which leads to a critical coating weight:

Cwcr ¼ f ðFmcr Þ ð6:54Þ

Our experience suggests that the critical value of Fm (situated between


8–9 N) is attained by mono-extrudates at a film thickness of about 35 m.
Co-extrudates display such values at a lower thickness, of about 20 m.
A theoretical minimum peel is observed for a 180 peel angle. In
reality, minimum peel values were obtained for lower peel angles. These
discrepancies may be explained by the low flexibility of the face stock
material; however, gradually increasing face stiffness and thickness
decreases the peel angle and so the peel adhesion [3]. On the other hand,
the extra work needed to bend the face material will raise the observed
adhesion value. In order to overcome this, a ‘‘reverse adhesion test’’ can be
performed by mounting the tape to be evaluated, adhesive surface up, onto
a standard test panel, to eliminate the backing effect. The 180 peel adhesion
values (under-layered) show a linear dependence on the thickness of the face
material [3]. Table 6.11 contains typical 180 and 90 peel adhesion values
for labels with different face stock materials and peel angle.
As can be seen from Table 6.11, the stiffness of the face stock exerts a
strong influence on the peel values. The dependence of the peel value on the
peel angle and dwell time is shown in Table 6.12. In fact, not only the
stiffness but [as discussed above, see correlation (6.53)] also the resistance to
elongation influence the peel value (Table 6.13).
In these measurements, the transfer of the peel force was achieved with
the aid of a different material (i.e., the face stock had a combined paper/
material construction) where the paper layer was contacting the adherent
and the material layer (paper, film, etc.) was subjected to tension. Thus by
using different force transferring materials the peel force was more or less

Table 6.11 Peel Adhesion Values as a Function of the Face Stock Material and
Peel Angle (Peel Force, N/25 mm)

Face Material Peel Angle ( )

PSA Paper Film 90 180

Tackified, water-based acrylic x — 8.3 25.0 PT


Water-based acrylic — x 8.0 12.0
Tackified, water-based CSBR x — 8.4 25.0 PT
Rubber-resin solvent-based x — 12.5 23.0
Rubber-resin solvent-based removable x — 4.6 13.0
282 Chapter 6

Table 6.12 Dependence of the Peel on the Peel Angle and the Dwell Time (Peel
Value, N/25 mm)

PSA Face Material Peel Angle ( )


Dwell
Chemical nature Physical state Paper Film 90 180 Time

Tackified acrylic Water-based x — 8.3 25.0 PT 0


Acrylic Water-based — x 8.0 12.0 0
CSBR acrylic Water-based x — 8.4 25.0 PT 0
Rubber-resin Solvent-based x — 12.5 23.0 0
Rubber-resin Solvent-based x — 4.6 13.0 0
Acrylic Water-based — x 5.7 5.2 0
Acrylic Water-based — x 6.0 12.0 15 sec
Acrylic Water-based — x 10.0 14.0 24 hr

Table 6.13 Dependence of the Peel on the Bulk


Properties of the Face Stock Material. The Influence of
the Mechanical Properties of the Face Stock Material
(Maximal Tensile Strength of a PE Film) on the Peel Force

Ultimate Tensile Strength (N/cm) Peel (N/cm)

8.0 0.13
9.0 0.23
10.0 0.30
11.0 0.40
12.0 0.54
13.0 0.60
14.0 0.70
16.0 0.87
18.0 1.07
A nontackified acrylic PSA was used. 180 peel adhesion on
stainless steel was measured.

dampened according to the plastic/elastic character of the material. Soft,


plastic materials (e.g., film) exhibit less peel strength than resistant ones
(e.g., paper) (Table 6.12). As an example for the peel dependence on the face
stock stiffness or rigidity, aluminum-laminated paper is removable without
primer (using an adequate PSA), and a solvent-based silicone release liner
should be used in order to get an appropriate release force.
Peel strength increases with the thickness of the face stock material,
according to a power law dependence [148]. With gradually increasing face
Adhesive Performance Characteristics 283

material thickness one needs to consider two additional features that affect
the result, namely, the rapidly decreasing peel angle, which will drop the
observed peel value, and the extra work needed to bend the face material,
which will raise the observed peel adhesion value [3]. As an example PET
face film shows a drop in peel adhesion from around 2 mil onwards,
aluminum shows a continual rise, and polyethylene film shows little change
or a slight decrease. To overcome this, a ‘‘reverse adhesion test’’ can be
performed [3]; a standard PET film strip is applied and rolled down to
standardize the face material effect. A plot of adhesion vs. face thickness
can then be drawn which can be extrapolated (for different PET thicknesses)
indicating the value of adhesion without the added influence of the face
stock.
The influence of the carrier rigidity is evidenced by special debonding
tests which are affected by the stiffness (flexibility) of the carrier material.
Such a characteristic test is the shaft loaded blister test (SLBT). The shaft
loaded blister test may have both the ease and convenience of peel testing
and the plastic deformation characteristic for a pressurized blister test [142].
The theoretical framework for measuring the fracture energy using the shaft
loaded blister test was developed by Wan and Mai [144]. For small angle of
deflection the following relation exists between P (the shaft load), wo (the
central shaft displacement), a (the crack length), h (film thickness), and
material properties (E):

Ehw3o =8 Pa2  0:5 W ð6:55Þ

where the interfacial adhesion (W) can be calculated as a function of P.


Unfortunately, the experimental results are contradictory. Samples consist-
ing of stacked films (either 1, 2, or 4 plies thick) were investigated utilizing
the shaft loaded blister test to determine if the backing’s tensile rigidity had
any effect on the measured G (apparent strain energy release rate) [149].
The results suggest that the number of plies had no effect on the measured
G when an experimental compliance calibration is incorporated into the
calculations. G values measured from the SLBT were compared to values
obtained using the pull-off test (and were of the same order of magnitude for
different substrates). In a similar manner the (calculated) value of W is not a
function of the number of stacked films, n. On the other side the sensitivity
of W to the film thickness h reflects that the actual system is a viscoelastic
bimaterial made up of the polymer backing and PSA, whereas the theoreti-
cal system is a single elastic material.
In the peeling master curve Piau et al. [103] represent the quantity

3/7G as a function of V, where


¼ Ve2/(Eh3) is a dimensionless number
284 Chapter 6

representing the effect of the adhesive viscous forces compared to the


rigidity effects of the backing. Miyagi et al. [150] relate the stiffness of the
laminate to the shear modulus of the adhesive. The stiffness is defined as P/
where P is the load applied at the center of the support and  is the resulting
displacement of the support. Their analysis provides the following
expression for stiffness:

P= ¼ ð4E= Þ fb½2ðt þ hÞ3 =L3 g ð6:56Þ

where is a function of the adhesive thickness (t), carrier thickness (h),


length (L), and sample width (b); E and G are the tensile and shear moduli
of the carrier, and Ga is the shear modulus of the adhesive. The stiffness
increases with the adhesive shear modulus, as a function of t/h.
The surface properties of the carrier material affect the adhesive
performances of the pressure-sensitive products too. When a laminate is
being debonded, there are at least four possible modes of failure, including:
adhesive failure with debonding of the adhesive from the web, adhesive
failure with debonding of the adhesive from the substrate, cohesive
failure or splitting of the adhesive itself, and paper/film tear or failure of
the web itself. The failure mode is influenced by the surface quality of the
face stock [151,152]. Clay-coated papers exhibit paper tear if the surface
strength of the layer is not sufficient [153]; water-based dispersions of
polymethylacrylate can be used for removable paper labels, but they are not
removable from PVC and lacquered surfaces [84]. The substrate nature
influences the peel and removability; for example, a two-year warranty was
given for Fascal 1800, except from PMMA, nitrolacquer, and polystyrene
surfaces.
In an ideal case, the correlation between the anchorage of PSAs to
the face stock material AF, cohesion of the adhesive C, and adhesion of the
PSAs on the substrate As should be valid [138]:

AF > C > AS ð6:57Þ

In the case of permanent adhesion As should be higher than the


mechanical resistance of the face stock material or of the substrate. For
removable adhesives, As should be less than the minimal forces producing a
deformation or deterioration of the face stock material or of the substrate.
Energy considerations imply an adequate anchorage of the PSA onto the
face stock material, but this will depend on the surface quality of the face
stock material. In fact the surface characteristics of the face stock material
influence the anchorage of the PSAs on the face stock material through the
Adhesive Performance Characteristics 285

wetting of the surface and the anchorage of the coated, wetted adhesive on
this surface. The first step in the bonding process is the wet-out of the
surface by the adhesive; surface roughness and energy affect this wettability.
Wettability is an initial prerequisite to high adhesion. Adhesion is at a
maximum at zero interfacial energy. The influence of the surface properties
of the carrier material and of the adherent on the adhesive performances will
be discussed in the next sections together with the effect of the experimental
conditions.

Influence of the Release Liner on the Peel. The surface and bulk pro-
perties of the release liner influence the peel resistance of the PSA. The
structure, nature, and thickness of the adhesive layer affect the peel value
[138]. Silicone transfer to PSAs, contamination of the adhesive layer through
crosslinking catalysts, or mechanical damage of the PSA layer (as a replica of
the release layer) may negatively influence the peel resistance.
At a given peel rate release force decreases as coating thickness is
increased (1.4–4.2 m) [117]. This is consistent with earlier understandings
of varying adhesive thickness. The increasing thickness h of the deformable
layer within the PSA laminate results in shifting the viscoelastic window to
lower frequencies. Traditional high release agents (HRAs) used in release
coatings ‘‘freeze out’’ interfacial slippage resulting in increased adhesive
deformation. The HRA reduces the segmental mobility of the PDMS chain
within the cured network leading to increased shear stress and an increase in
the practical peel force.
In a manner similar to the face stock the bulk properties of the liner
influence the peel of a label from the release liner. When increasing the
thickness of the release liner, the high speed release values are approxima-
tively doubled when the liner is removed from the label. This confirms that
the nature of the release liner (probably its stiffness) has a considerable
influence on the high speed release force.

Influence of Label Laminate Construction on Peel Adhesion. The


energy absorption during peeling decreases the value of the peel force.
Therefore, multilayer labels, with more deformable liquid adhesive layers
and with the ability to absorb the debonding energy by viscous flow, show
lower peel values than single-ply labels (Table 6.14).
The peel resistance depends on the rheology of the adhesive layer R
and on the peel force F (see Chap. 2). The rheology of the adhesive layer
depends on its thickness h. In the case of multilayer labels, the adhesive
thickness is the sum of the different layers or:

h ¼ hi ð6:58Þ
286 Chapter 6

Table 6.14 Peel Dependence on the Laminate Structure. Stiffness and Peel Values
for Mono- and Bilayer Laminates, with the Same Face Stock and Different
Adhesives

Laminate Structure Characteristics

Stiffness Peel on glass


Face stock PSA (mN/cm) (N/25 mm)

First layer Soft PVC Untackified acrylic — —


Second layer Paper Tackified water-based acrylic 698 17.4
Tackified water-based 480 17.8
acrylic/CSBR
Tackified water-based acrylic 515 16.0
Tackified solvent-based 483 10.5
rubber-resin
The untackified acrylic adheres to the test substrate.

Therefore the increased layer thickness improves the flow properties of the
adhesive (i.e., its energy absorption capacity). Thus it is assumed that the
peel is inversely proportional to the number N of the adhesive layers,

Peel ¼ f ðN 1 Þ ð6:59Þ

On the other hand, increasing the layer number increases the stiffness of the
laminate (i.e., it changes the debonding angle) and thus decreases the peel
too (see the influence of the face stock on the peel). Therefore, it is more
accurate to assume a pronounced effect of the number of layers through
viscous flow (higher adhesive thickness) and peel angle (higher laminate
stiffness), or:

Peel ¼ f ð1=N  Þ ð6:60Þ

where  > 1. As shown in Table 6.15, this hypothesis is confirmed by


experimental data. Increasing the number of layers increases the stiffness
of the laminate and at the same time decreases the peel value. Increased
stiffness requires more energy for the flexure of the face stock material (i.e.,
the peel force increases). Thus, the peel will depend on the number of layers
according to a more complex correlation.

Peel ¼ f ðN =N  Þ ð6:61Þ
Adhesive Performance Characteristics 287

Table 6.15 Dependence of the Peel on the


Number of Layers

Laminate Characteristics Peel (N/25 mm)

Number Stiffness
of layers (mN/cm) On glass On PE

2 18 4.3 2.9
4 60 3.9 2.6
6 100 3.6 2.6
8 152 3.7 1.5
Water-based acrylic formulation. Film face stock
material.

where  >> . A multilayer film-based label may increase the energy


absorption by the viscous flow of the solid state components (face stock
included) (i.e., the peel decreases again). Thus, one can assume that:

Peel ¼ f ½ðN Þ=ðN   N  Þ ð6:62Þ

where  takes into account the adhesive slip and the change of the
debonding angle, denotes the increase of the mechanical resistance with
the number of layers, and  accounts for the deformation of special film-
based face stock materials. For multilayer filmic labels the dependence of the
peel on the number of layers may be summarized by the following equation:

Peel ¼ f ðN =N  Þ ð6:63Þ

where N ¼ N Ng (i.e., N > N). It may be assumed that the peel decrease
with increasing number of layers is more pronounced for film labels than for
paper labels. This hypothesis is confirmed by the data in Table 6.16. Paper
laminates show a less pronounced peel decrease with increasing number of
layers (Table 6.17). In a similar way paper layers in hybrid (paper/film)
labels impart increased peel adhesion.
In some experiments the influence of the stiffness deformation
resistance of multilayer laminates was studied. In these tests a film label
was reinforced with paper labels coated with different PSAs. The
distribution of the peel forces depends on the deformability of the film
face stock, or on the slip between the labels (i.e., the adhesive nature). As
seen from Table 6.16, reinforcing the film label with an elongation-resistant
288 Chapter 6

Table 6.16 The Influence of the Number of


Layers on the Peel Adhesion of Film Laminates

Number Stiffness Peel on Glass


of Layers (mN/cm) (N/25 mm)

2 39 23.5
4 236 19.5
6 408 12.0
8 644 8.0
Water-based PSA coated on soft PVC.

Table 6.17 Peel Dependence on the Number of Layers for a Multilayer


Paper/Film Laminate

Laminate Stiffness Peel on Glass


Laminate Construction (mN/cm) (N/25 mm)

PSA Film – – – – 46 8.4


PSA Film PSA Film – – 130 6.4
PSA Film PSA Paper – – 245 8.0
PSA Film PSA Film PSA Paper 483 7.8

material (paper) improves the peel value. On the other hand, the peel value
of multilayer laminates increases with the internal rigidity (no motion) of the
sandwich (i.e., with the adhesion between the laminate face components).
The slight anchorage of a rubber-resin adhesive on soft PVC dramatically
decreases the peel value of the sandwich. Control of the viscous flow can be
achieved by superimposing adhesive layers.

Influence of Experimental Conditions on Peel. The surface tension,


surface structure (roughness), and thickness of the substrate influence the
peel resistance [138]. Test conditions such as dwell (contact) time, contact
pressure, debonding rate, debonding angle, and temperature also affect the
peel resistance. At lower speed, viscous flow is the predominant influence on
adhesion (peel value), while at higher speed, the elasticity determines the
adhesion; at lower speed, peel adhesion increases. Some of these parameters
were discussed earlier (see Chap. 2).
Concerning the stress rate and its influence on the peel, an increase of
the temperature allows increased molecular mobility resulting in increased
tack and reduction in shear resistance. A high rate of stress is equivalent to a
Adhesive Performance Characteristics 289

drop in temperature. Adhesive transfer may occur at higher unwinding force


and rate; faster peeling rates may transfer the adhesive or break the label.
Physical and chemical interactions between the adhesive and the sub-
strate result in a build-up of the adhesion (i.e., the peel increases with the
dwell time) [138]. This is a general behavior observed for PSAs with different
chemical compositions (e.g., acrylics or neoprene latices) [58]. The build-up
of the peel depends on the chemical composition and time [105,106,154].
Because of the slow flow of the adhesive, adhesion tests require at least 0.5
sec of dwell time [155]. The surface of a PSA can be affixed to a substrate to
which it will be adhered, but once affixed, adhesion builds to form a strong
bond. Removable adhesives can be formulated so as not to display adhesive
build-up, that is, their adhesion to the substrate does not increase to the
point where the label (tape) cannot be removed cleanly, even after exposure
to heat in a drying oven. Adhesion build-up requires a long time [3].
Different peel values are obtained when testing immediately after applica-
tion (PSTC 1), after a 20 min dwell time (ASTM D-1000), and after a 24 h
dwell time. The adhesion can still be climbing after 30 days. Peel adhesion
is a function of the dwell time [156]; after 10, 30, 60, 180 min, and 24 h
dwell times the following peel values were obtained: 4.0, 4.6, 5.3, 5.5, and
8.0 N/2 cm. Different methods specify different dwell times: AFERA 4001
requires 10 min, PSTC 1 and FINAT specify 20 min and 24 hr, respectively
[139]. In order to test adherent fiber tear, tape is applied to cardboard with a
2 lb roller, immediately pulled off at a 90 angle, and rated for fiber tear.
Roughness of the carrier surface and of the adherent influences
bonding and debonding also. As discussed earlier dwell time depends on
the surface quality too. Time dependent peel adhesion is commonly assumed
to be due to the limited rate at which the rough adhesive layer can conform
to the flat substrate; the applied label retains a hazy look for a substantial
time when applied to clear transparent substrates. Generally the substrate is
rough too. According to Elmendorp et al. [157] the paper backing although
highly calendered, is never perfectly flat for two reasons: the distribution of
wood fibers in the paper (the formation) is never perfect, and the humidity
cycles the paper undergoes during adhesive coating and drying can lead
to microcockling, a slight waviness. Surface roughness with an amplitude of
several microns with a wavelength of 2–5 mm is typically seen. Nardin and
Ward [158] discovered a simple relationship between the surface roughness
(i.e., the depths of the pits or valleys on the fiber surface) and the degree
of interfacial adhesion due to the mechanical interlocking. Hui et al. [159]
showed that when the dimensionless parameter
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ¼ ð2= Þ 2W =E > 1 ð6:64Þ
290 Chapter 6

where  is the aspect ratio of a typical surface profile (or asperity) and is
the number of asperities per unit length, the surfaces will jump into contact
with each other with no applied load, and the contact area will continue
to expand until the two surfaces are in full contact. Self-adhesion can be
avoided by increasing the aspect ratio of the asperities, by decreasing
the work of adhesion, by decreasing the density of the asperities, or by
increasing the modulus of the polymer. In the coating and testing practice,
adhesive contact is needed, i.e., the PSA has to conform onto substrate or
carrier.
According to Tordjeman et al. [19] the actual contact area is a function
of the contact force (Fc) and the apparent modulus (Ea) which takes into
account the competition between the thickness effects and the roughness
[see correlation (6.29)].
Other phenomena can compete with adhesive conformation during the
dwell time also. The study of the adhesion of butyl rubber to rigid counter
surfaces showed [160] that after 1000 hr of contact at 80 C the strength of
adhesion become about ten times larger than the value measured after short
contact times, and still showed no signs of reaching an equilibrium level.
This increase is attributed to slow molecular rearrangements at the interface
rather than to development of specific bonds, because it took place to a
similar degree for butyl rubber and chlorobutyl rubber, against different
substrates. A similar but somewhat smaller increase in adhesion also
occurred for vulcanized layers.
For tackified adhesives the overall work of adhesion was measured as
a function of the storage time [161]. Generally it decreases with increasing
storage time; the resin migrates to the surface. The increase of the domain’s
size with increasing time suggests that at long times there will be a surface
layer of tackifier-enriched material, making a more rigid surface that will
inhibit wetting and degrade the overall adhesive performance.
The strength of an adhesive bond formed with PSA is generally
recognized to be defined by two major contributions: specific interaction
forces across the adhesive-substrate interface and the work of viscoelastic
deformation of the PSA polymer to rupture the adhesive bond. Within the
frameworks of the diffusion theory of adhesion, the well-known effect of
adhesive strength enhancement with the time of adhesive-substrate contact
is treated as the time required to provide the penetration of PSA
macromolecules into the substrate [162,163]. One of the contributing
mechanisms to polymer adhesion is said to be diffusion of chain elements
across the polymer/polymer interface. The concept makes sense when
polymers in close contact are above their respective Tg. The common
features of various PSA polymers are high diffusivity and translational
molecular mobility as well as fibrillation of the polymer at cohesive failure
Adhesive Performance Characteristics 291

under applied debonding stress. Diffusion across the interface contributes to


bond strengths in joints involving flexible polymer chains, where dispersion
forces dominate at the interface, and where favorable acid-base interactions
abet the dispersion forces [164]. Chalykh et al. [165] studied the kinetics
of interfacial zone formation in the closure of adhesive bonding to a
polyethylene substrate and showed the lack of PSA interpenetrating into the
substrate and vice versa: no interdiffusion across the interface occurs.
During the stage of adhesive bonding the molecular mobility of PSA
polymers contribute to the process of polymer relaxation, which in turn
accounts for the increase of the strength of adhesive joints with PSA-
substrate contact times. There is no need to advocate the diffusion theory of
adhesion to explain the relevance of enhanced molecular mobility for the
adhesive behavior of PSA polymers [163].
As discussed in Chap. 2, the time-temperature strongly affects the
peel. When increasing the temperature the viscosity of the adhesive
decreases and its anchorage to the substrate will be improved [105]; on
the other hand, adhesion increases and thus peel reaches a maximum for
a given temperature. Temperature influences the effect of roughness also.
As demonstrated in [166,167] the debonding mechanism of SIS-based PSA is
strongly influenced by roughness; at normal temperatures surface roughness
is detrimental for good contact. High numbers of defects are trapped at
the interface, relaxation is incomplete. At higher temperatures roughness
prevents the cavities from rapidly expanding in the plane of the film and
increases the adhesion energy. At low temperatures adhesion decreases and
the peel adhesion decreases as well [106]. One should take into account
the poor peel adhesion properties of most current PSA labels at low
temperatures (deep-freeze labels). The rate of separation and the tempera-
ture are related to the adhesive strength in a complicated manner, and these
factors can dramatically influence whether or not adhesive or cohesive
failure will occur. Peel decreases with decreasing temperature [3]. Cohesion
and auto-adhesion of styrene-butadiene elastomers were studied by Hamed
and Hsieh [17], through peel tests at different peeling rates and
temperatures. Peel resistance depends on the difference between the
application temperature and the Tg of the adhesive. For instance, at 0 C
the peel resistance value of an untackified acrylic adhesive is about 210%
lower than the peel resistance value of the same adhesive at 23 C [168]. By
unwinding of tapes so-called adhesive inversion may occur also [169]. In this
case the separation (debonding) by unwinding of the tape does not occur at
the interface of adhesive/printed carrier but between adhesive and primed
carrier. Such behavior can be observed especially at temperatures lower than
10 C and at high unwinding speeds. The effect of temperature on the 180
peel force in the temperature range between 20 C and þ40 C could be
292 Chapter 6

predicted from the profile of the velocity spectrum of the rolling adhesive
moment [170].
The dependence of the peel force on the peel rate is complex. It has
been shown that several regimes can be obtained as a function of the peel
rate [167]. The first regime is the cohesive regime and it can be divided into
three different ones (as a function of the peeling rate). A two-dimensional
one (at very low speeds), a three-dimensional one showing regularly spaced
ribs, and finally a three-dimensional one, showing the breaking of ribs at
the edge of the free surface [103]. Assuming that one can add the cohesion
energy wc, Piau et al. [103] propose the following relation for the first regime
for low velocities:

G ¼ wc f1 þ 0:1ðEh3 =Ve2 Þ3=7 ðV=s Þg ð6:65Þ

For the dependence of the peel force on different parameters Kaelble


[143] proposed a law using Bikermann’s approach, for a Newtonian
adhesive:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
F=b ¼ 2 2 ðEh3 =eÞ1=21=2 ðK=tm Þ1=2 ð6:66Þ

where  is the viscosity of the adhesive, K/tm is a peeling velocity factor


which is not so well defined, but tm possibly will vary like e/V.
Fukuzawa and Uekita [119] proposed a theoretical formula of peel
adhesion see [see Eq. (6.49)] as a function of the peel angle and peel rate. At
low peel rates the peel adhesion depends mainly on the work of adhesion,
but at high peel rates it is influenced by the work of deformation.
According to [92] the maximum of the curves (peel force-peel rate)
corresponds to the change in mode of failure from cohesive (at lower peel
rates, < 300 mm/min) to adhesive at higher debonding rates. At low tensile
rates the adhesive fracture occurs at comparatively small tensile stress
values, but at significantly greater values of elongation, ". With the increase
in extension rate, the critical magnitude of tensile stress at fracture ( f)
grows rapidly, approaching a limiting value which corresponds to a
maximum cohesive strength of the PSA film ( c). At the same time the
work of PSA tensile straining varies, passing through a maximum.
As confirmed in [160] also the peel strength is strongly dependent on
the effective rate of peeling. At low rates the elastomer layer could not be
detached cleanly. Instead it split apart leaving rubber behind (cohesive
failure). Under these circumstances the peel strength is the same as the tear
strength measured by tearing apart a thin sheet of the elastomer. As the peel
Adhesive Performance Characteristics 293

rate was increased, however, an abrupt transition occurred to interfacial


failure at much lower peel forces, only about one twentieth as high. This
transition is attributed to failure of entanglements to flow apart at a critical
peel rate. After the transition the peel strength again increased with peel rate
paralleling the cohesive strength of the polymer itself but now only about
one tenth as high. This increase is attributed to increasing viscous losses in
peeling away the elastomer as it approaches the glassy state.
According to [94] the difference between stick and slip peel force
diminishes as the rate increases. There are two interfacial debonding zones
as a function of the rate. The flatness of the peel force vs. peel rate response
is optimal at intermediate molecular weight with very high molecular weight
leading to poor adhesion at low rates [95]. In the correlation used by Gent
and Schultz [21] and later by Carré and Schultz [171]:

G ¼ wo gðMc Þ ðVÞ ð6:67Þ

where wo is the cohesive energy, g(Mc) is a molecular dependent term, and 


is a dissipative function of the velocity, the function (V ) has been deter-
mined in a few cases to be a power law of the velocity. Using a statistical
software package, peel strengths were measured for PSA tapes, according to
the following correlation [172]:

Load ¼ 0:55 þ 0:88 rate þ 1:17  103 time ð6:68Þ

It has been demonstrated, that single pull peel testing does not always
predict the joint failure that occurs as a result of cyclic loading at subcritical
loading levels [173–175]; adhesive joints when critically loaded at levels
below the load at first yield in the single pull experiment, can open up at
appreciable rates. The application conditions of pressure-sensitive products
and their influence on the end-use properties are discussed in detail in [176].
Peel Force from the Release Liner. A special case of peeling is the
separation of the label from the release liner (i.e., labeling). As discussed
above the peel resistance depends on the peeling rate and angle. As stated by
[177] the peel decreases with the angle: 90 peel is about twice of that of 180
peel. In a similar manner the peel force to separate the PSA from the release
liner depends on the peeling rate and angle, and adhesive/release nature [177]:

Peelforce ¼ f ðV, Þ ð6:69Þ

where V and  denote peeling rate and angle. Separation rates of 200 m/min
are common during labeling. At these rates condensation silicone release
294 Chapter 6

layers display an increase of the peel. In the FTM 10 test method a much
lower test rate is used, while the FINAT method is designed for a rate of 10–
300 m/min with the first test apparatus built in 1984. For testing purposes a
sample of 2.5  70 cm or 5  70 cm was proposed; samples have to be stored
under pressure (70 g/m2, 20 hr, 23 C) before testing. Performance charac-
teristics after aging should be tested upon storage at 40 and 70 C,
respectively [177], because the release nature and age also influence the
peel. There are some new materials that are supposedly insensitive to release
nature and age (e.g., Nacor 80). Acrylic PSAs require a separate study for
each formulation with regard to the level of the release force and its stability.
For current water-based acrylic PSAs on solventless release liners, release
forces average 0.05–0.06 N/25 mm (FTM 3). Interaction between water-
based acrylic PSAs can increase this value to more than 0.1–0.15 N/25 mm
(e.g., Vinacryl 4512). Increasing the concentration of the methylsilicate
HRA, increases release force. However, the largest impact of the HRA
on release force is typically at low to intermediate peel velocities (0.005
to 0.17 m/s) while increasing the molecular weight between crosslinks mainly
influences release at high peel rates [117].
Textiles, nonwovens, and surfaces coated with alkyd-, PUR-, acryl-,
epoxy-, nitro-, and powder-lacquers, and plastisols be used as contact
surfaces for testing purposes. Crystal glass is proposed as substrate by
FINAT, steel by AFERA, and glossy stainless steel by ASTM or BWB-TL
[178]. These materials are used in order to eliminate the influence of the
adherent on the peel.
If the only factor governing the peel force were the total extent of
contact (i.e., the number and type of interactions across the interface) then
the scale of the contact zone would not influence the measured peel strength
[17]. In general this is not true, particularly for soft vulcanized elastomers,
whose strengths are a measure of dissipative energy losses that occur while
they are being peeled away. Adhesive strength to paper and stainless steel is
greater than to polyester film [83]. Therefore the bond fails due to breakdown
within the adhesive layer when paper is used. When polyester film is used as
face material it peels away, leaving the adhesive layer virtually intact on the
steel plate. The influence of the adherent is illustrated by PSA labels applied
on PVC (where the peel value changes in plasticized PVC) or by removability
from fragile substrates (e.g., lightweight papers) [179]. Peel and tack should
be tested on different substrates (e.g., cellulose, polyamide, glass, cardboard,
metal) [180]. Temperature resistance, the lowest application temperature,
and the suggested operating end-use temperature should be determined. The
influence of the adherent surface on the peel is shown in Table 6.18.
The mechanical strength of the substrate influences the peel resistance
too. Paper possesses an anisotropic structure. Cellulose fibers which have the
Adhesive Performance Characteristics 295

Table 6.18 The Influence of the Adherent Surface (Substrate) on


the Peel. Peel Values of PSAs with Different Chemical Bases, as a
Function of the Substrate Used

Chemical Composition Peel (N/25 mm)

Base components Supplier Glass Cardboard PE

1360 Hoechst 12.0 PT 2.0 11.0


V205 BASF
80D BASF
CF52 A&W
FC88 UCB 15.0 PT 11.0 PT 13.0
PC80 UCB
CF52 A&W
FC88 UCB 18.0 PT 11.0 PT 12.0
CF52 A&W
V205 BASF 18.0 PT — 15.0
80D BASF
CF52 A&W
V205 BASF 14.0 PT 10.0 13.0
V208 BASF
80D BASF
CF52 A&W

typical dimensions of 4  20  2000 m, are layered with most of the fibers
lying like ribbons in the plane of the paper sheet. Therefore paper is very
resistant to in-plane stresses and very sensitive to out-of-plane or z direc-
tional stresses. It is susceptible to delamination. Such delamination occurs at
higher peel rate [181]. The peel force initially rises until the peel force applied
to the paper surface is high enough to remove fibers after which the force
drops to a low steady-state value, corresponding to delamination (paper
failure). Therefore paper delamination is characterized by a high initiation
force and a relatively low propagation force. Microscopic observation
reveals that paper failure starts with the lifting of an individual fiber.
Bikermann and Whitney [182] reported that paper is much more likely to
delaminate in peel if the tape overlaps the edge of the paper. In peeling the
actual lifting forces on the paper surface are less than the measured forces
because of viscous dissipation in the PSA layer. The actual forces were
calculated on the paper surface in peel from the observed stretching of paper
in 90 peel. Comparing peel force and the actual force at the paper surface,
as a function of the peel rate the two types of forces were equal at low peel
rates, whereas the forces experienced by the paper were about 2/3 of the
296 Chapter 6

measured force at high peel rate. Yamauchi and coworkers reported peel
forces and failure modes as a function of peel rates and paper types [183].
Pelton et al. [184] studied the peel force as a function of log (peel rate),
and found quite different interdependences for glassine, copy paper, and
filter paper. Glassine (tracing) paper, a smooth, nonporous substrate gave
interfacial failure only and peel force increased with peel rate. Filter paper
showed paper failure and low peel forces. Copy paper showed increasing
peel forces with peel rate until the failure mode changed to paper failure
giving low peel forces. Filter paper, which is strong by conventional
measurements, gave only paper failure. Thus roughness, porosity, and the
extent of adhesive penetration dictate whether paper failure occurs with peel.
According to [185] the deformability of the substrate must be taken
into account. For small values of the adherend thickness, the assumption of
bending-dominated crack propagation becomes invalid when the opening
forces become significant. At the other extreme the peak stress in the
interface region affects the plastic deformation of the adherends. Thus, peel
force depends on the adherend thickness. For a given macroscopic energy
release rate the plastically deforming adherends do not allow high levels of
opening stresses to be developed within the adhesive. Thus the appropriate
stress levels for development of a microcracking zone are reached before the
adhesive fails cohesively at the crack tip. In contrast the elastic adherends
induce much higher levels of opening stress over an extended region in the
adhesive layer. Thus before cohesive failure at the crack tip occurs, large
enough stress are developed ahead of the crack tip to allow a microcracking
zone to be developed. This change in mechanism means that toughness of
joints may not be the same for substrates that remain elastic as opposed to
cases where they deform plastically. The deformability of the substrate has
to be taken into account by the choice of the peel test methods also. For
instance T-peel is used for flexible adherends [186,187] (see Chap. 10 too).
Chalykh et al. [92] investigated the strength of the PSA joints
on different substrates (PE, PP, PVC, PET, PA, and hydrate cellulose). The
strength of the adhesive bond for the substrate was found to be ordered like
the corresponding magnitudes of the thermodynamic work of adhesion.

Wa : HC > PA > PVC > PP > PE ð6:70Þ

Piau et al. [103] measured for steel as carrier, lower peel and lower increase
of the peel with the velocity. The adhesion energy of the acrylic adhesive on
silicone release coating is 10 times smaller than that on steel. Moreover the
stress-strain curves have different shapes on steel [188].
Adhesive Performance Characteristics 297

In peeling tests described in [91] different substrates (PyrexTM,


PlexiglasTM and PMMA) were used. Four mechanisms of cohesive failure
have been found as the reduced velocity increased. The peeling mechanism
on Plexiglas also led to the creation of another mechanism not observed
on Pyrex. Peel values on various substrates (stainless steel, cardboard, and
polyethylene) for different adhesives are listed in [189].

Other Factors. Peel adhesion does not depend on the number of


passes with a standard application roller and it does not depend on the
solvent used to clean the test plates. On the other hand, peel increases with
the application pressure and temperature [132].
Rheology is frequency dependent and strain frequency influences the
peel; repeated, cyclical peeling off reduces the peeling force (see Chap. 3).
Readherence onto a paper substrate decreases as the number of peeling
trials grows. In a lot of experiments the peel of an experimental PSA from
newspaper as substrate was tested. The following data for the peel force
were collected as a function of the number of peel trials:
Peel from paper, first trial: 110 g/12 mm.
Peel from paper after 50 peel trials: 85 g/12 mm.
Peel from paper after 100 peel trials: 75 g/12 mm.
The readherability of a pressure-sensitive product was tested after repeated
lamination and delamination [190].
The dependence of the peel on the cohesion is very complex. For some
adhesive formulations based on natural rubber and SBR a direct
proportionality between peel and cohesion was established [191].
Humid environments can adversely affect bond strength. Soaking
conditions, even high humidity, often cause irreversible loss of adhesive
strength or failure. According to Czech [192] for splicing tapes the peel
adhesion strongly depends on the relative humidity.
Release transfer to the adhesive surface can reduce the peel resistance
also. Druschke [193] found for a noncovered PSA layer a peel value of
12.3 N/20 mm which changed to 9.7 N/20 mm for the same adhesive after
release paper contact. The subsequent adhesion, i.e., the peel resistance
value after contact with release paper depends on the initial peel resistance
of the adhesive and on the substrate [194]. Another problem is peel retention
after aging. The requirements concerning peel retention after aging differ as
a function of the product build-up and end-use.

Improvement of Peel Adhesion


Improving the peel means optimizing it in such a way that it meets the end-
use requirements. Regulating means increasing or decreasing the peel value
298 Chapter 6

and changing the character of the bond rupture related to the requirements
of a permanent (paper tear) or removable (adhesion break) bond.

Permanent/Removable PSAs. Pressure-sensitive adhesive-coated


materials are functionally divided into two broad classes: permanent and
removable [79]. The first of these is represented by the so-called permanent
materials in which the properties of PSAs are selected so as to form an
adhesive bond with the target substrate which is strong and, apart from
degradation, does not weaken significantly with time. The second broad
class consists of ‘‘removable’’ or ‘‘peelable’’ PSA-coated materials in which
the PSAs form an adhesive bond, of functionally adequate strength with the
target, but which after an extended period of adhesion (typically days to
months) can be peeled away from the substrate, without damaging it,
without leaving any residue of PSA on the target substrate, and without the
adhesive-coated material tearing itself apart. To be peelable PSA-coated
materials require a combination of properties. The PSAs must form a
bond of adequate strength with the target substrate, but which does not
subsequently either dramatically weaken, in which case the adhesive-coated
material might fall off the substrate, or significantly increase in strength, as
this would tend to limit peelability. From these considerations it follows
that the adhesive should possess an adequate cohesive strength in order to
minimize any tendency of the adhesive film itself to rupture, as well as a high
bonding strength to the face material (anchorage) onto which the adhesive is
coated. It is clear that the formation of a strong bond between a PSA and
its face material and a relatively much weaker bond between adhesive and
target substrate are in conflict. Consequently, many of the peelable PSAs
currently on the market do not fully meet these contradictory requirements;
effectively, they are only less permanent than typical permanent PSA-coated
materials. The reduction of the peel is a relatively simple procedure which
(at least theoretically) can be achieved by increasing the viscous component
of the PSA, that is, the component converting the high speed bond-breaking
force in permanent adhesive deformation (viscous flow). Unfortunately this
viscous flow may exhibit adhesive break (i.e., after bond breaking adhesive
residue is left on the adherent). Such a peel reduction is inadequate to obtain
a convenient removable PSA.
Summarizing, a limited peel value and a high shear strength are
required for removable adhesives [86]. A typical value for removable
pressure-sensitive tapes is 10 N/25 mm [195]; lower values (2.5–5 N/25 mm)
are required for labels. Shear values of at least 100 min at room temperature
were suggested [196]. In fact these values should be considered as guidelines
only since paper tear occurs at 20–25 N/25 mm. Removable PSAs should
exhibit a peel adhesion value of 1.5 N/25 mm [195]; 1–6 N/20 mm are
Adhesive Performance Characteristics 299

suitable values for removable labels [197]. In general, the same removable
water-based acrylic PSAs for labels may be used for tapes. Removable PSA
labels are a special type of permanent label, where low peel force levels,
as a function of dwell time, and a pure adhesive break from the substrate
are assumed. Removable behavior depends on energy dissipation which
influences the debonding (Fig. 6.15).
The adhesive (its bulk properties, formulation, nature, structure, and
molecular weight), laminate components, laminate geometry, and debond-
ing conditions influence the removability. During peeling tests at least four
possible modes of failure may occur [198]:
1. Adhesive failure (primary): debonding of the adhesive from the
face material (i.e., not removable).
2. Adhesive failure (secondary): debonding of the adhesive from the
substrate (i.e., removable).
3. Cohesive failure: splitting of the adhesive itself (i.e., not
removable).
4. Web or film tear: failure of the face material itself, hence not
removable.
Energy dissipation during debonding depends on such parameters as
interfacial adhesion, polymer rheology, rate of peel, temperature, angle
of peel, face stiffness, and coating weight. Several requirements have to
be fulfilled in order to achieve good removability [199]. These include

Figure 6.15 Schematical presentation of stress dissipation as a function of the time


for permanent and removable PSAs.
300 Chapter 6

self-adhesion on the substrate (tack), peel adhesion (in order to assure


bonding), cohesion in order to avoid legginess or stringing, and superior
adhesion to the face stock rather than to the substrate (anchorage). For
suitable peelable labels adhesive failure from the substrate is desired. As
discussed earlier removability implies lower stresses, and lower stresses are
possible using stress-restricting polymers, fillers, or a primer coating [200].
Peel depends on the nature and geometry of the bonding components (i.e.,
the self-adhesive label and substrate). For a given adhesive, face stock, and
substrate, the interaction of the adhesive with both solid state components
of the joint will depend on the interface adhesive/solid state component. The
type of face material is governed by the end-use of the label; therefore
the label-maker should optimize the interface between the face stock and
the PSA. The most important parameters of the face stock influencing
the removability of the PSA label are its flexibility and the anchorage
(adhesion) of the adhesive to the face stock. Effective bonding between the
face and the adhesive is critical to the performance of the PSAs; for example,
given an adhesive with internal integrity, the adhesive properties exhibited
by that system will be proportional to the affinity of the adhesive for the face
material. Face stock interactions bring the definition of ‘‘permanent’’ and
‘‘nonpermanent’’ PSAs much more into focus by emphasizing the fact that if
a permanent bond is formed between the label and the substrate, the failure
mode will be face oriented and the result is 100% transfer of the adhesive to
the substrate. Conversely, if the bond to the substrate is nonpermanent, then
a clean release from that substrate is encountered and the adhesive remains
on the face material.
Repositionability and readherability are performance characteristics
that depend on the peel build-up. Both allow the debonding of the PSP after
its application and its rebonding. Repositionability is an aging-related
characteristic; it needs a slow build-up of the adhesion on the substrate. In
this case peel build-up is delayed in time but is not limited as an absolute
value. Repositionable labels may also be permanent.
According to a simplified theory bonding and debonding can be
controlled through the value of G0 and through the dependence of this value
on the frequency. For removable products G0 possesses the same (small)
value at different frequencies. Thus a main possibility for regulating the peel
resistance is the control of the G0 of the PSA.

Use of Primers. Primers are necessary when using removable PSAs


[201,202]; they modify the face stock-PSA interaction. Polyolefin face
stock materials are generally top coated to improve printability; in a similar
manner primer coatings are applied on polyolefin face stock materials to
increase the anchorage of the adhesive [203]; PVC-based tapes also need
Adhesive Performance Characteristics 301

a primer coating [204,205]. Soft PVC should be primed when used with
rubber-resin-based PSAs [206]. There are a few studies describing the use
of primers in order to obtain removable adhesives [207–209]. The use of
the primer generally decreases the peel value [210]. Primers are generally
used to promote a stronger bond between PSAs and face materials, to absorb
stress, and to strengthen the face stock, especially when paper labels are
concerned.
There are very different recipes for primers. As in the case of adhesive
base elastomers, primer coatings use more crosslinking agents, harder resins,
resin solutions, or self-crosslinked elastomers. A primer coating for
polyethylene face stock material typically is composed of a chloroprene
rubber (20–60 pts, 60% chlorine), an EVAc copolymer (40–80 pts, with 25%
VAc), and chloropolypropylene (1–15 pts with 25–50% chlorine) [209].
While stearic acid is effective in preventing adhesive build-up, it leaves a
greasy deposit referred to as a ‘‘ghost’’ which could stain the substrate; the
metallic stearates do not do this [210]. Contact cements (polymers with high
cohesive strength and a strong tendency for two surfaces of the polymer to
adhere to one another when placed in contact) are preferred as primers;
isoprene polymers and polychloroprene are particularly good as contact
cement. The use of a crosslinked contact cement as a primer improves
peelability and increases bonding on the carrier from 0.5 to 2.4 kg [211,212].
More about primer formulations can be found in Chaps. 7 and 8. A detailed
discussion of primer formulations for different pressure-sensitive products is
given in [213]. One approach is to use a key coat including particulate
matter, so as to make the surface of the coating rougher, and provide a
better key for PSAs; primers provide a bond between adhesive and face
stock [214,215]. Primed face stock ensures more uniform peel values and
thus better removability (Table 6.19).
In order to achieve a maximum anchorage of the adhesive to the face
material, a silicone release coating may be used as a primer for silicone
adhesives [51]. In making PSA tapes, polyacrylamide may be used as the
binder between the hydrophilic cellulose and the hydrophilic butadiene-
styrene latex. The primer also improves the anchorage for natural rubber-
based PSAs [216]; the primer modifies the separation energy of the surface.
Primers are somewhat flexible and generally have high anchorage to
the face material. Thus, they can absorb impact stresses without adhesion
failure [200]. There are removable PSA formulations with a primer coating
that contain a contact cement with the same contact cement in the PSA as
well. The flexible sheet label has on one side a primer coating firmly bonded
to the face stock, which includes a contact cement, and above the primer
coating is a peelable PSA (including the same contact cement). The contact
cement serves to firmly bond the primer coating and the adhesive coating
302 Chapter 6

Table 6.19 The Influence of the Primer on the


Removability of PSA Labels

Removability

Substrate With primer Without primer

Steel 2 3
Copper 2 2
Aluminum 1 3
Glass 2 4
PE 2 5
HPVC 2 3
PMMA 2 2
PA 2 5
PET 1 5
Average 1.8 3.4
Removability was evaluated subjectively. The following
removability indices were used: 1— very low peel, no
residues; 2 — low peel, no residues; 3 — low peel, legging;
4 — medium peel, no residues; 5 — high peel, no residues.

together. Contact cements are a class of adhesives which function by


exploiting the properties of certain polymers, which exhibit a high cohesive
strength and a natural tendency for two surfaces of the polymer to adhere
when placed into contact. Examples include polymers such as isoprene,
natural rubber, polychloroprene, and one class of acrylic polymers. A low
Tg is a property common to polymers which are used in contact cements.
The stress absorption capacity is a function of the primer’s own
softness. As shown in Table 6.20, adhesives with a different ‘‘hardness’’ (peel)
act as a primer coating in different manners. The soft (removable) ones do
not cause any peel decrease; the peel adhesion decrease is proportional
to primer stiffness. In these experiments different adhesives were used as
the primer (first) coating. The second layer (i.e., the PSA contacting the
substrate) has always been the same (a permanent, tacky PSA). The necessity
of stress absorption (i.e., energy absorption) is illustrated by the data of
Table 6.21. It can be seen from Table 6.21 that the peel value of the
removable adhesive layer increases with the coating weight of the primer
coating. A primer coating weight equivalent to 5–10% of the permanent
adhesive weight is able to change the peel of removable PSAs.
The peel force will be influenced by the material (PSA) in the non-
contact area as well [17]. When the elastomer is more viscous, the strength
will be less affected by flaws (noncontacted zones) compared to an elastomer
Adhesive Performance Characteristics 303

Table 6.20 The Influence of the Softness of the


Primer on the Peel Value/Removability of PSA

Primer Peel (N/25 mm)

Soft hot-melt PSA 20.0


Removable rubber-resin 23.0
solvent-based PSA
Hard hot-melt PSA 24.0
Permanent rubber-resin 24.0 PT
solvent-based PSA
None 24.0 PT
A tackified water-based acrylic PSA was used.

Table 6.21 Influence of the Primer Structure/ Geometry on the


Peel Value/Removability

Removability

Relative thickness of the layer (%)

Substrate 100 75 50 0

Steel 2 1 1 3
Copper 2C 1 1 2
Aluminum 1 2 2 2
Glass 2 1 1 3
PE 2 2 2 5
PMMA 2 1 1 2
PET 2 1 1 5
Average 1.85 1.3 1.3 3.1
Removability was evaluated subjectively. The following removability
indices were used: 1 — very low peel, no residues; 2 — low peel, no
residues; 3 — low peel, legging; 4 — medium peel, no residues; 5 — high
peel, no residues; C — cohesive failure.

that responds more elastically. As discussed above, primers may improve


stress distribution. Another stress-reducing possibility is the choice of the
adequate polymer (PSA) formulation or the use of adequate fillers [200].
The primer coat typically is present on the coated face stock in an
amount ranging from 2–20 and more often 5–10 g/m2. The lower limit is
determined by the requirement to ensure complete coverage of the coated
face material, with a good coating pattern, and the upper limit is determined
304 Chapter 6

primarily by costs as there appears to be no significant benefit in increasing


the coating weight of the primer beyond the limit indicated; generally
2–5 g/m2 of primer is used [86]. For tapes with an adhesive layer of
16–20 m, the thickness of the primer coating should be 1–1.5 m [217].
The thickness of the primer (manufactured via an electron beam) from a
monomer blend should be less than 5 m, preferably less than 1 m, and
ideally less than 0.5 m. Indeed it is believed that thicknesses approaching
a single molecular layer would function efficiently [218].
The use of polyaziridine as a primer shows the effect of the cross-
linking on the removability of PSAs. Peel adhesion of the crosslinked PSA
(after 24 hr) is 0.51 kg/cm versus 0.67 kg/cm for the uncrosslinked one. The
crosslinking of PSAs shows similarities with the use of lightly crosslinked
rubbers in thermoplastics for controlled soft-phase dispersions.
Readhering adhesives permit removal of the coated face material and
allow readherence to another surface. Paper substrates which were
precoated with removable, readhering adhesives are commercially available.
One commonly known product of this type is marketed under the trade
name ‘‘Post It’’ as note pads [219]. In order to achieve a good removability,
many different procedures are used, including the use of hard (Tg > 10 C)
suspension particles, crosslinking the matrix, and low coating weights
(7 g/m2) [199].

Stress-Resistant Polymers. Stress-resistant polymers are those which


develop a controlled crosslink density or which are internally soft. This
means that crosslinking agents or plasticizers may improve the removability.
Like craze deformation for plastic-elastomeric blends, PSAs are subject to
energy dissipation via deformation work; thus shear deformation is
necessary [200].

Flexibilizers. Polymers exhibit two kinds of fracture mechanisms:


ductile and brittle. Ductile fracture involves overall yielding of the specimen
and since the volume of yielded material is large, this failure mechanism
absorbs a great deal of energy; however, it is slower than brittle fracture.
Removable PSAs undergo ductile fracture in that they absorb all the break
energy.
A number of superimposed adhesive layers having different gradients
of shear/creep compliance can meet the requirement of releasable adhesion
to plastic surfaces such as polyethylene. Flexibilizers are plasticizer com-
pounds which react with the polymer; because of their volume they are able
to keep the chain segments apart. These plasticizers increase tack, but lower
cohesion [200]. Adding a 15% butyl dispersion to a water-based PSA
formulation may improve its removability properties [220]. Flexibilizers, like
Adhesive Performance Characteristics 305

plasticizers introduce creep, which is a function of molecular weight and


plasticizer loading, or

Creep ¼ f ½ðPlasticizerÞ, ðMolecular weightÞ1  ð6:71Þ

A high molecular weight and a low plasticizer level reduce creep. Plasticizers
may be incorporated in the peelable PSAs to soften the adhesive and
thus improve peelability. Care should be exercised as some plasticizers
can have a tackifying effect on adhesive polymers and this may limit
the amount to be used; however, it is possible to use quite large quantities
of plasticizers, even up to 50 wt% on the solid adhesive, but more often
they are used at 10–20 wt%. Generally, plasticizers impart a fast tack
increase, a slower peel increase, and very fast decrease of the cohesive
strength.
Fillers. Stress reduction by fillers (i.e., reducing peel adhesion) is
well known in the tape industry (e.g., 0.5–10% filler, particle size 150 m,
in rubber-based PSAs) [221]. Peel adhesion decreases in formulations
containing fillers [106]. Rubber dispersions also may be used as fillers for
water-based PSAs [220].
Strengthening of the Face Stock. Paper face materials can lead to
paper tear, when peeling off the label. Reinforcing the fiber structure of the
paper face stock with a primer can strengthen the surface layer in order to
avoid fiber tear. High strength face stock materials (e.g., rubber latex-
impregnated papers) are often used. In some cases the primer strengthens
the face stock material [216].
Reduction of the Contact Surface. Peel adhesion depends on the
surface area between the label and the adherent. Reducing the contact
surface between the label and the adherent improves peelability or
removability. Peel resistance and removability can be regulated by
modifying the ratio between the adhesion surface and the application
surface [222]. There are different methods to reduce the contact area
between the adhesive and the substrate, including: reducing the coating
weight of the adhesive, increasing the stiffness of the face stock (or
substrate), applying a discontinuous coating on the face stock material, and
removing contact points or areas between adhesive and substrate.
Discontinuous adhesive coatings (i.e., ungummed areas) reduce the
contact area since in some cases, it suffices to coat only a minor portion
of the face stock surface with adhesive [223–226]; a number of adhesive
and nonadhesive zones may be used [219,227]. There are several patents
concerning the use of filled PSAs that decrease the adhesive/adherent
306 Chapter 6

contact area. Some use glass beads as filler [228]; microspheres also may
be used [229]. The reduction of the contact surface may be achieved by
overlapping the contact surface of the PSAs with powdery materials [230].
As an example, adhesive tapes applied for sealing windows and bonding side
moldings onto cars use UV-cured acrylic adhesives containing 2–15 parts
per hundred (phr) hydrophilic silica-glass microspheres [231]. Adding an
expandable filler to the formulation reduces the peel resistance. Regulation
of the contact surface can be achieved also by crosslinking the whole
adhesive surface and making it sufficiently fragile to allow crack migration
of the uncrosslinked adhesive.
Coating Weight Adjustment. As far as the PSA component is
concerned the properties which contribute to peelability are principally a
limited tack, a low build-up factor (lack of age hardening), and a relatively
soft adhesive. These properties can be achieved conventionally, by omitting
or only using a low level of tackifier, or by including tack deadeners, such as
waxes, and by including plasticizers. The most important parameter is
the adjustment of the coating weight. A way to reduce the peel (i.e., to get
a removable PSA) is the reduction of the coating weight. Reducing the
coating weight from 20–25 to 15–10 g/m2 allows removability for PSA
formulations that are based on acrylic, vinyl ether, and plasticizer. For PSAs
coated on paper the coating weight for repositionable labels may be as low
as 7 g/m2. In this case an initial peel value of 110 g/12 mm (180 peel at
300 mm/min) is enough to resist a vertical application (5 hr debonding
method) [199].
Shear Strength for Removable Adhesives. Permanent adhesives
generally possess a higher cohesive and shear strength than removable
ones. On the other hand, several high-shear adhesives used for tapes also
are proposed for removable labels (e.g., Ashland, Aroset 2530-W-50). For
removable adhesives a shear value of 50–1000 min at 70 C or 50–1000 min
at 175 F is suggested [232]. The minimum level of shear for a removable
PSA also depends on the coating weight. From a good removable PSA no
adhesive transfer can be accepted.
Problems Concerning Removability. The most frequently encoun-
tered problems for removable PSAs are too high peel forces, the change of
the peel force in time (build-up), adhesive residue left on the substrate, low
tack, and pronounced bleeding, smearing, and migration. It is difficult to
find removable labels which can be removed without leaving any adhesive
residues behind. Generally, the migration or bleedthrough of removable
PSAs is more pronounced than for permanent PSAs. This will be discussed
in more detail in Chap. 8.
Adhesive Performance Characteristics 307

Peel Adhesion Values


Improved peel adhesion requires a higher tackifier loading than tack. Higher
melting point resins are suggested. Permanent and removable labels/tapes,
labels for hand-applied labeling or gun labeling, labels for room
temperature or high temperature use, labels for use on plastics or metals,
etc., have quite different peel adhesion values. Table 6.22 summarizes some
of the performance characteristics specified by manufacturers of PSA
laminates. Detailed lists with peel values for various pressure-sensitive
products are given in [71,176,233].

1.3 Shear Resistance (Cohesion)


Pressure-sensitive adhesives possess typical viscoelastic properties which
allow them to respond to both a bonding and a debonding step. For
permanent adhesives the most important step is the debonding one; the
adhesive should not break under debonding (mainly shear and peel) forces
(i.e., permanent adhesives must provide a higher level of cohesive or shear
strength than removable adhesives). This is an inverse requirement with
respect to Dahlquist’s criterion (see Section 2.3 of Chap. 3) for a minimum
value of the compressive creep compliance to achieve tack. Creep (fluidity
under low forces) results in edge bleeding, migration, and poor die-
cuttability; high resistance to fluidity allows low instantaneous flow (i.e.,
low tack). The cohesion (shear) is important as an index for label-processing
characteristics (bleeding, migration, die-cutting) and also end-use properties.
Shear resistance is measured as a force required to pull the pressure-
sensitive material parallel to the surface to which it was affixed with a

Table 6.22 Typical Peel Values

Adhesive Nature 180 Peel on Steel

Chemical Physical End-use of the laminate Value Units Ref.

Acrylic Hot-melt PSA Label 90–105 oz/in 46


Acrylic Solvent-based Transfer tape >7 lb/in 234
— — Highly aggressive label/tape 60–100 N/100 mm 235
SIS Hot-melt PSA Packaging tape 10–18 N/25 mm 236
SBS Hot-melt PSA Double-sided carpet tape 8.5 N/cm 237
— Hot-melt PSA Permanent label 69–77 oz/in 238
— Hot-melt PSA Removable label 11–14 oz/in 239
Acrylic Water-based Application tape/film 1.3–5.0 N/25 mm 240
Acrylic Water-based Protective film 100–200 cN/cm 241
308 Chapter 6

definite pressure. Practically one measures the holding time under standard
conditions. Unfortunately there are no standard values of the time until
shear failure occurs.

Measurement of the Shear Resistance


Test conditions influence the shear resistance also. Holding power is
basically a viscosity effect. It was found that the shift factors used to plot
shear measurements at several temperatures on a master curve were almost
identical to the shift factors used for G0 and G00 master curves. The shear
resistance of PSAs may be measured statically or dynamically. Currently,
static shear test methods use a constant load at longer test times; they
generally show poor reproducibility and need very long measurement times.
Better results are obtained with the hot shear test, where the cohesion of
the sample is measured at an elevated temperature. Dynamic shear tests
measure the cohesion of the sample in a tensile tester under increasing load
(force). It has previously been indicated, that molecular disentanglement can
take place at a testing rate of 0.0100 /0.25 mm per minute [96]. A dynamic
shear test can be set up on an adhesion tester at this slow speed. While the
width of the sample can be of any standard width, the height of the sample
needs to be limited, so that it does not become a tensile test of the backing.
If a hot shear test is carried out in such a way that the test temperature is
gradually raised, and if the temperature at which the bond fails is taken as a
characteristic value, specialists speak about a dead load hot strength test
or shear adhesion failure temperature (SAFT). There also are shear test
methods that measure neither the time until debonding nor the temperature
of the debonding but rather the deformation of the sheared sample (i.e., the
slip of the sample after a given time is measured) [234].
A combined peel and shear strength of the adhesive can be tested (e.g.,
20 hold strength of the adhesive to corrugated board). The important shear
properties of adhesives are the shear strength or the shear stress at failure,
the shear modulus or the shear stress/shear strain ratio, the creep modulus
or the shear stress/shear strain ratio at time t, the shear stress or the load/
bond area ratio, and the shear strain or the shear slip/adhesive thickness
ratio [241]. A detailed description of the shear measuring systems is included
in Chap. 10. Here the importance of hot shear measurements only will be
highlighted. They exhibit a lower discrepancy of the test values and allow
extrapolation of the data for long end-use periods.
Room temperature shear values display variations ranging from
100–1000%. Hot shear values normally do not vary more than 100–200%.
On the other hand, better die-cuttability requires hot shear values to be
higher by an order of magnitude (1000%); hence, a real improvement of
Adhesive Performance Characteristics 309

Figure 6.16 Interdependence: room temperature shear/hot shear. 1) Test using


heated panel; 2) test using climatized atmosphere.

the cohesion can only be evaluated via hot shear measurements. Figure 6.16
illustrates the measuring errors for the most commonly used hot shear test
methods (relative to room temperature shear tests); the preheated plate
method is better. Note that the value of shear or hot shear alone is not
important, but rather it is the interdependence between them.
At times the hot shear values may be converted into room temperature
data. Using the Williams–Landel–Ferry time/temperature superposition shift
factors developed for SIS triblocks the 80 C data may be extrapolated to
yield a shear holding time at 25 C of 387 days for a pure triblock versus
86 days for a commercial product.
The cohesion is a real measure of the internal structural resistance
of the polymer. Generally, the mechanical properties of a polymer depend
on its cohesion. Different methods exist to determine the mechanical
resistance of a polymer, such as the measurement of the tensile strength,
elongation at break, etc. Generally, adhesives are used as thin layers, where
the adherents undergo shear deformation during debonding and delamina-
tion. Therefore, the measurement of the shear was accepted as a criterion
for the cohesion.
Cohesion also can be measured by the Williams plasticity [242]. A
measure of the Williams plasticity is given by the thickness of an adhesive
310 Chapter 6

pellet (mm) after 14 min compression at a fixed temperature and under a


fixed load. Peel adhesion measurements also provide information about the
cohesion. Shear may be also measured as tension. For polymers obeying
Hooke’s and Newton’s laws, shear is directly proportional to tensile
strength [243]. The evaluation of the viscosity is an index for holding power,
telescoping, oozing, and die-cutting properties [244].

Interdependence of Adhesive/Cohesive Properties. Generally, the


practical use of PSA labels requires an adequate adhesion/cohesion balance,
that is, a certain aggressiveness of the adhesive, characterized by instant
tack, a time-dependent final adhesion (bonding strength) characterized by
peel and shear, according to the method of debonding (i.e., the end-use
stresses). The fast, instantaneous flow required for high tack does not allow
less flow and lower deformation when debonding (i.e., shear is inversely
proportional to tack and peel is partially so). On the other hand, the
adhesion/cohesion equation is more sophisticated. A good peel adhesion
supposes a certain flow and a low to medium cohesion, but a high peel needs
high cohesion when debonding. There is an internal relationship between
peel adhesion and shear strength (Fig. 6.17). For PSA specification and
current production control purposes both values should be measured. For

Figure 6.17 Interdependence: hot shear/peel. Tackified water-based acrylic PSA.


Adhesive Performance Characteristics 311

Acronal V 205 minimum and maximum hot shear values (70 C, coated on
PET) are 25 min and 100 min, respectively, which correlates with a peel from
steel of 6–8 N/25 mm. The adhesion-cohesion balance and its regulation for
various pressure-sensitive products is discussed in detail in [232].

Factors Influencing Shear Resistance


Like tack and peel adhesion, shear is influenced by the characteristics of the
adhesive and of the laminate, as well as by the debonding conditions.

Nature of the Adhesive. Special built-in functional groups, cross-


linking, and high molecular weight can ensure high shear and low tack
properties. The use of harder monomers also improves the cohesion. For
example, a higher acrylic acid level in solvent-based acrylic PSAs yields
higher shear values [245]. Cohesive strength is independently controlled by
the hard monomer content and increases as the concentration of this hard
monomer increases [2]. Cohesive properties of a polymer can be affected by
introducing polar groups that interact by forming secondary bonds; such
interactions are classified as hydrogen, dipole-dipole, and dipole-induced
bonding. Monomer-reinforced copolymers are suitable as PSAs, since they
possess high shear strength [246]. Minor amounts of unsaturated olefin
carboxylic acids and/or sulfoalkyl esters of such carboxylic acids improve the
cohesive strength of PSAs; it is preferred that the polymer contains at least
about 0.1 wt%, usually about 0.1–10 wt% of these components. The higher
the percentage of iso-octyl acrylate, the tackier the adhesive; conversely, the
higher the percentage of acrylic acid, the higher the shear properties.
Permanent adhesives generally provide a higher level of cohesion shear
strength than removable adhesives. Rubber-resin adhesives ensure a compro-
mise between high cohesive strength for conformance to curved containers
and good quick stick required in automatic labeling [247]. Acrylic adhesives
provide a high level of adhesion to surfaces that are difficult to adhere to.
However, most acrylic-based adhesives do not have a cohesive strength
equal to normal rubber-resin systems and may display cold flow. Showing
up as a halo of adhesive around the die-cut label, cold flow can inhibit
efficient labeling. All adhesives used in automatic labeling systems require a
level of quick stick that will provide for efficient adhesion with a minimal
application pressure.
It is the general finding of the industry that for the same tack and peel
values acrylic emulsion PSAs show much lower shear holding power
compared to that of their solvent-based counterparts [248]. The difference
in shear holding power between emulsion vs. solvent-based PSAs is about
1–2 orders of magnitude, depending on the monomeric composition used.
312 Chapter 6

Tobing and Klein [248] have shown that for gel free low-Tg acrylics (based
on 2 EHA or BuAc) there was no significant difference between emulsion vs.
solvent-borne PSA film adhesive performances. However, commercial
products have a high level of gel and low shear. The four-fold decrease of
shear by tackifying was attributed to the lack of entanglement between
the linear polymer chain and the microgels. In a novel acrylic emulsion
0.5–1.0 wt% IBMA (isobutoxymethylacrylamide) was copolymerized to
provide interlinking of the microgels in the film upon heating.

Shear Dependence on Coating Weight. Shear depends on the coating


weight. A suitable test for the measurement of shear properties of an adhesive
is the thick (6 mm) adherend shear test (TAST) since it is conducted using a
tensile testing machine. However, finite element analysis shows that the stress
distribution in the thick adherend shear test is not pure shear. The values of
maximum shear strain gained from the thick adherend shear test are not
comparable to those obtained from tests in pure shear [249].

Shear Dependence on Molecular Weight. Cohesion increases with the


molecular weight. The ultimate tensile strength of a polymer reaches a
maximum at about 500–1000 main chain atoms. Longer molecules do not
impart higher strength because their various segments act independently of
each other. The influence of the molecular weight on the shear resistance is
illustrated by the changes of the shear resistance due to adhesive aging (i.e.,
destruction and depolymerization). SIS block copolymers undergo oxidative
degradation at elevated temperatures by a mechanism which leads
predominantly to scission of the polymer chains. This leads to a fall in
molecular weight and a resulting decrease in viscosity and holding power
[250]. The composition of the adhesive (i.e., its nature and molecular weight)
influences the shear; shear measured as SAFT is directly dependent on the
softening point (i.e., the molecular weight of the resin) [251]. The dependence
of the shear on the molecular weight is illustrated by the manufacture of hot-
melt PSAs. When high shear mixers are used the degradation depends on the
mixing environment. In contact with open air, degradation occurs rapidly
and polymer breakdown is observed. A dramatic reduction in melt viscosity
and holding power results from even immediate degradation [252]. Polymer
viscosity and shear are interdependent [253]. The latter depends on the
molecular weight and its distribution. A broad MWD polymer may have a
lower cohesive strength than one with a narrow MWD and lower molecular
weight. The dependence (increase) of the cohesive strength as a function
of the molecular weight was described by Schrijver [254]. For HMPSA, the
effect of MW and aromatic content of the resin on the shear resistance was
observed. The number average molecular weight showed a strong correlation
Adhesive Performance Characteristics 313

with shear adhesion values [255]. The holding power has been shown [256] to
be controlled by zero shear viscosity which is related to Mw, Me, and Tg by
the following equation:

Log o ¼3:4 log Mw  2:4 log Me þ A=ðT  Tg þ 70Þ þ B ð6:72Þ

where A and B are constants. This equation shows that while o is reduced by
Me it is also increased by Tg of the polymer. Copolymers containing vinyl
neoester units exhibit higher Me and Tg, thus one can expect that peel, tack,
and holding power can all be improved. As illustrated in [67] shear for EHA-
based copolymer was 3.9–1.6 hrs (100  0.500  1 kg); the neo-based polymer
gave shear values of 16–21 hr. The shear resistance is improved by the
increase of the melting point of the resin [47]. Gordon et al. [257], using
different Escorez (manufactured by Exxon Chemical) resins with different
softening points, list the SAFT data of 60/40 adhesive formulas. The SAFT
improves with increasing softening point of the resin. In a similar manner
shear increases with the increase of the ring-and-ball softening point of
Bevitack (Bergvik-Arizona) resins used as tackifier for hot-melt PSAs [258].
For tackification of Neoprene Latex 102, the 32 C holding power increases if
the 49 C softening point (SP) resin is replaced by a 72 C SP resin. Table 6.23
illustrates the increase of the shear resistance with the increase of the
molecular weight.

Influence of Crosslinking on the Shear. Crosslinking affects polymer


properties in a manner similar to molecular weight increases but in a more
pronounced way [50]. Crosslinking imparts high shear and lower tack [83].
The phenomenon is well known from the field of plastics where shear is
proportional with the degree of crosslinking for polyethylene (i.e., the
peroxide concentration) [259]. Similarly, increasing the amount of the resin
in a silicone adhesive increases its crosslink density and results in greater

Table 6.23 Shear Dependence on the Molecular Weight

Shear [min at temperature ( C)]

Molecular Weight of
the Base Elastomer 50 60 70 50 60 70

1 M1 20 2 1
1.5 M1 420 120 60
3 M1 > 1200 > 1200 780
A CSBR was used as base elastomer.
314 Chapter 6

cohesive strength and higher peel adhesion values [260]. Physical cross-
linking has the same influence as chemical crosslinking. The beneficial
influence of a 100% pure triblock in hot-melt PSA formulations is demon-
strated in creep testing. Here the PSA is subjected to a dead load shear, and
if every molecule of the rubber is able to fully participate in the associative
crosslinking shear, bond holding times are excellent.
Reactive resins with methylol groups in the fluid state are efficient
tackifiers but also crosslink with the neoprene rubber used as elastomer,
and lead to high strength bonds [45]. Aqueous PUR dispersions were used to
improve the adhesion of acrylic adhesives for better film formation and
mechanical properties [53]. The molecular interaction between acrylic
and polyurethane not only influences the product stability but also governs
the adhesion performance. The use of nonionic PUR yielded higher shear
strength (from 28 kN/m2 to 65 kN/m2), but the use of anionic PUR resulted
in very high shear strength (700 kN/m2). Anionic PUR behaved like a
crosslinker. Shear resistance in acrylic formulation improved as crosslinker
concentration increased. Electron beam-cured crosslinkable rubbers are
used for PSAs also; the SAFT increases for these polymers with increasing
gel content, but the relationship between gel content and SAFT is a gradual
slope [50]. A polymer gel content greater than 0.8% is required to increase
the high temperature holding power above 300 min. Crosslinking increases
the modulus; therefore crosslinked adhesives have the highest shear holding
power (Table 6.24).
Unfortunately crosslinking is associated with the loss of tack (and
peel) and thus it may not be suitable for PSA modification, except under
strictly controlled conditions (e.g., solvent-based adhesives). Ulrich [261]
addressed the need to improve both the adhesion and the heat resistance
at 70 C of an acrylate/acrylic acid adhesive when applied to low energy
surfaces; this can be achieved by employing a tackifying resin for better

Table 6.24 Influence of the Crosslinking on the Shear

Adhesive Nature Crosslinking Agent Shear

Level Temperature
Chemical Physical Nature (%) Value Units ( C) Ref.

Acrylic Solvent-based AZ 0 0.1 hr RT —


Acrylic Solvent-based AZ 1 1.0 hr RT —
Acrylic Solvent-based — — 1.0 min 70 9
Acrylic Solvent-based — 1 5000 min 70 9
AZ, aziridine; RT, room temperature.
Adhesive Performance Characteristics 315

adhesion and by crosslinking the tackified PSA for better heat resistance
at 70 C. The new class of acrylic hot-melt PSAs that contain reactive
methacrylic end groups that can be crosslinked, imparts good temperature
stability [253]. Conventional natural rubber adhesives do not have the
required temperature resistance for masking tape; to impart the required
degree of heat resistance to the adhesive, it must be crosslinked.
Shear Dependence on Composite Status. Generally, the composite
structure of the PSA layer contains technological (polymerization or
formulating) additives necessary for converting the liquid adhesive, and
additives for improving the convertability of the laminate. Additives related
to the water-based nature of the adhesive (e.g., surfactants, defoamer, and
water) generally decrease the cohesion level of the adhesive. On the other
hand converting additives (fillers) improve the cohesion level. Thus:

Cohesion ¼ f ½ðFillersÞ,ðWater, SurfactantsÞ1  ð6:73Þ

The resistance of separating liquid-bonded plates depends on the layer


thickness and viscosity [see Eq. (2.14)]. Concerning the influence of the
coating weight on the shear, the flow (creep) limit (FL) may be formulated
as a function of the adhesive modulus, the elastic work, and the dimensions
of the sample. Shear (defined as the limit of the slip) also depends on the
layer thickness. Experimental data reveal a complex dependence of the shear
on the coating weight. Above a minimum coating weight level the shear
values decrease with increasing coating weight. Therefore the following
dependence between coating weight and shear S can be written:

S ¼ f ðCwd Þ1 ð6:74Þ

where d 6¼ 1. Since the coating weight influences the drying rate, and vice
versa, therefore the shear strength depends on the drying as well (Fig. 6.18).
Generally, the thickness of the adhesive layer influences the strength of
the adhesive joint [133]:

 ¼ tan  ¼ v=D ð6:75Þ

where g is the shear resistance,  is the displacement of the component of the


joint, and D is the thickness of the adhesive layer. The strain rate is given
by the following equation:

 0 ¼ d=dtðk=DÞ ð6:76Þ
316 Chapter 6

Figure 6.18 The dependence of the shear on the drying of the PSA coating.

On the other hand, the mechanical resistance of adhesive joints


depends on the length of the overlap [262]. For adhesives obeying Newton’s
and Hooke’s laws, the shear is inversely proportional to the thickness of the
adhesive layer (i.e., the coating weight) [243]. Shear really depends on the
adhesive’s intrinsic cohesion and on the adhesive/label geometry (i.e., shear
resistance measurements should be carried out with well-defined samples).
Influence of Substrate on Shear. Shear depends on the substrate
surface also. Therefore for packaging tapes cardboard shear (in hours) and
cardboard adhesion (%FT) are tested too. The influence of the substrate on
shear is illustrated in Fig. 6.19 [217].
Shear Resistance and Face Stock Material. The evaluation of the
shear strength should be carried out relative to the face stock used.
Unmodified PSAs are used mostly for film coating, while tackified PSAs are
used mainly for paper labels. Consequently, film and paper laminates
display different shear levels. Shear values for the same adhesive differ when
coated on different face stock materials. Common, untackified acrylic PSAs
display less cohesive strength when coated on paper than on polar films.
The shear resistance of the joints depends on the surface tension of the
substrate [263]. A surface treatment (i.e., an increase of the surface tension)
Adhesive Performance Characteristics 317

Figure 6.19 The influence of the substrate nature on the shear. Shear values for
samples on glass, polyethylene, stainless steel, and PVC as adherend.

Table 6.25 Dependence of the Shear on the


Surface Nature of the Face Stock Material

SAFT Values ( C)

PSA Formulation PET Paper

A 90 60
B 60 45
C 70 55
D 75 60

may improve shear characteristics. Table 6.25 illustrates the dependence of


the shear resistance (measured as SAFT) on the face stock surface.
Because of the static nature of the shear tests in general, the
deformability, the elasticity of the face stock material is less important,
but may nevertheless influence the test results. Therefore shear tests are
carried out with nondeformable, dimensionally stable face stock materials
(like paper or PET) or by reinforcing the plastic label on the back side
(see Section 3.1 of Chap. 10). The applied coating technology influences the
318 Chapter 6

coating weight and smoothness of the adhesive layer, and thus the shear
resistance.

Improving Shear Resistance


The values of the shear resistance depend on the cohesion of the adhesive
and on the laminate geometry. Cohesion may be improved with
hard monomers and by increasing the molecular weight (crosslinking).
The laminate geometry influences the adhesive flow. Upon improving the
anchorage of the adhesive to the face stock the adhesive bond fails via
cohesive break (i.e., within the adhesive layer). Therefore a low coating
weight (an optimum as a function of the roughness of the solid state
components of the label) will improve the shear resistance. Cold flow may be
limited by use of fillers; in some cases the use of reactive fillers improves
the shear resistance also. The homogeneity of the adhesive strongly
influences the resulting shear strength (see Chap. 8).

Shear Values
There are many data available about shear strength. Unfortunately most
of them are static shear values at room temperature. For these the method
used (i.e., the sample, dimensions, weight, adherent) influences the measured
absolute values. On the other hand, and in a quite different manner
from tack and peel values, shear test results for PSAs with different chemical
bases do not allow a real comparison of their cohesion-related properties
(e.g., cuttability); they are used more as an internal check of the adhesion/
cohesion balance for small changes in a given formulation. Test methods
and test conditions strongly influence the measurements. For instance
the SIS block copolymers have such high room temperature shear
values, that only highly plasticized formulations can be evaluated in a
convenient period of time [255]. For SIS copolymers the shear properties
were measured in the linear viscoelastic regime (small strain) on a parallel
plate rheometer. Differences due to the change of the SI content occur in
the range 5  103–5  101 rad/sec, where an increase in the SI content
leads to a decrease in the elastic modulus G0 and to an increase in the
loss modulus G00 [18]. Shear values obtained for hot-melt PSAs are generally
lower than those for solvent-based PSAs (Table 6.26). Dispersion-based
PSAs may exhibit lower or higher shear values as a function of
the formulation. High shear, high molecular weight solvent-based acrylic
PSAs are superior in shear to rubber-resin-based PSAs; CSBR exhibits
higher shear than common acrylic PSAs. Crosslinking improves the
cohesion even more.
Table 6.26 Typical Shear Values

PSA

Chemical Basis Physical State Shear Resistance

Temperature
Styrene Water- o
Acrylic Rubber-resin rubber-resin based Hot-melt Note Value Unit ( C) Method Ref.

— — x — — x Electron beam- 300 min 95 PSTC 7 50


cured SIS
x — — — x — 300 min RT PSTC 7 50
Adhesive Performance Characteristics

x — — — x — Crosslinked > 20 hr RT PSTC 7 —


x — — x — — Self-crosslinked 260 hr RT — 264
— — x — — x 4.6 d RT — 265
— — x — — x Kraton GX/ > 5000 hr RT l00  l00 1000 g 266
Regalrez 100/150
x — — — x — — 5–10 min RT — 267
x — — — x — — 0.5–2 min 70 — 267
x — — x — — Crosslinked > 200 min RT — 267
x — — x — — Crosslinked 20–100 min 70 — 267
— — x — — x — 100 hr 88 — 268
x — — x — — — > 16 hr RT 0.500  0.500 , 1000 g 46
x — — x — — — 0.7–2 hr RT 0.500  0.500 , 500 g 2
x — — — — — — 83 hr 70 1.27  2.54 cm, 1000 g 269
x x — — — — Removable 16–83 hr 70 — —
— x — — — — — 4–5 hr RT — —
319
320 Chapter 6

2 INFLUENCE OF ADHESIVE PROPERTIES ON


OTHER CHARACTERISTICS OF PSAs
2.1 Influence of Adhesive Properties
on the Converting Properties
Wettability depends on the adhesion-cohesion balance; high shear, low tack
adhesives exhibit better wettability than soft, tacky ones. Printability of the
laminate depends on its stiffness, porosity, and surface characteristics; all
these properties depend on the stiffness, anchorage, and migration of the
adhesive (see Section 2.2 of Chap. 7 and Chap. 8).

2.2 Influence of Adhesive Properties on End-Use Properties


The dependence of the removability on the adhesive properties is well
known. High peel adhesion and peel build-up do not allow removability.
On the other hand, low shear may cause legging or adhesive residue on the
substrate after peeling.
Labeling properties depend on the label adhesion to the release liner
(i.e., on the release or peel force). For gun labeling they also depend on the
tack and peel. Repositionable adhesives need low initial peel adhesion,
whereas tamper-proof ones require a high initial peel adhesion level.
A detailed discussion of the influence of the adhesive properties on the
end-use properties for the main pressure–sensitive products (e.g., labels,
tapes, protective films etc.) is given in [176].

2.3 Influence of Peel Adhesion


The influence of the peel on the end-use properties is evident. For
permanent/removable or tamper-proof labels the peel value remains the
main property. In a similar manner peel influences the labeling process,
where the release force (peel from the release liner) has to be different for
labels not used in dispensers (where a high release force would not be of
great consequence), and those used in dispensing guns (where a high peel
force of the label from the release would indicate poor conversion
properties, or poor labeling ability). The build-up of the peel resistance as
a function of the time is very important for the design of removable (no
build-up), repositionable (slow build-up), and tamper-proof (fast build-up)
labels. Different peel levels are required for labels with a multilayer
construction (i.e., similar to application and masking tapes). The peel level
has to be quite different for smooth, polar surfaces (e.g., glass or steel) or
rough ones (e.g., cardboard).
Adhesive Performance Characteristics 321

2.4 Influence of Shear


Low cohesion limits the applicability of the label/decal PSA grade with
respect to PVC film shrinkage both on the release liner during storage and
when applied to a substrate, with regard to wing-up (flagging) on curved
surfaces (mandrel hold), stringing, and edge ooze during converting. Shear
strength for film laminates is important for removable adhesives in order to
avoid the build-up of adhesive residues; for permanent adhesives cohesion
ensures good die-cutting properties and avoids oozing and bleed through the
face material.

3 COMPARISON OF PSAs ON DIFFERENT


CHEMICAL BASES

It is evident that on the basis of their chemical composition and rheological


behavior PSAs with different chemical bases can display quite different
properties. Next a comparative examination of PSAs on a different chemical
basis is presented. First the classical rubber-resin solvent-based PSAs will be
compared to acrylic PSAs.
In a separate discussion the advantages of water-based acrylic PSAs
over other water-based PSAs is examined. Although the whole range of
properties is covered, it remains clear that the most important properties of
a PSA label are those correlated with the adhesive characteristics.

3.1 Rubber-Based Versus Acrylic-Based PSAs


Table 6.27 summarizes the most important differences (i.e., the general
adhesion properties, the special adhesive properties, and the stability)
between rubber-resin or acrylic-based PSAs. As can be seen from Table 6.27,
rubber-resin adhesives possess very good general adhesive properties, but
remain defensive in shear. Acrylic PSAs display better specific adhesion
properties.
Generally, the advantages of acrylic systems include chemical and
water resistance, UV and oxidative stability, heat resistance, tack at varying
temperatures, and the balance of adhesive and cohesive properties. Acrylic-
based PSAs offer a unique combination of performance advantages relative
to hydrocarbon-based rubber-resin adhesives, and are used extensively in
end-use markets that demand excellent color and clarity, weatherability,
durability, and plasticizer migration resistance (i.e., overall versatility of the
adhesive). Acrylic PSAs do not exhibit yellowing and display good chemical
resistance [270]. Rubber-resin adhesives provide a compromise of high
322 Chapter 6

Table 6.27 Comparison of Rubber-Resin PSA Versus Acrylic PSA

Chemical Basis of the PSA

Acrylic Rubber/resin
Performance
Characteristics Solvent-based Water-based Solvent-based Hot-melt

General
Tack Fair Fair Good Good
Peel Fair Fair Good Good
Shear Good Good Good Fair
Specific
Cuttability Good Fair Fair Insufficient
Resistance to migration Good Fair Good Insufficient
Aging
Thermal stability Good Good Fair Insufficient
UV stability Good Good Fair Insufficient

cohesive strength for conformance to curved containers and good quick stick
required in automatic labeling, but they exhibit a varying batch-to-batch
consistency and can cause staining problems. Staining and softening usually
result when using rubber-based PSAs on PVC substrates. Low energy face
stock surfaces and soft rubber-based PSAs bond fairly well (anchorage), but
they are difficult to convert, whereas acrylic PSAs convert well, but exhibit
low bond strength (anchorage). Properties that are correlated with the
chemical instability of the rubber-resin basis (e.g., temperature and light
resistance) and the pronounced liquid character of the rubber-resin mixture
(e.g., cold flow and shear) remain inferior for rubber-resin PSAs as
compared to acrylic PSAs.

3.2 Acrylics and Other Synthetic


Polymer-Based Elastomers
In order to make a competitive evaluation of the properties of acrylics used
as raw materials for PSAs, the advantages and disadvantages of acrylics
as well as their main technical features and adhesive properties should be
examined. This study deals with the evaluation and examination of the
adhesive and end-use properties of acrylic-based PSAs as compared to other
PSAs. First tack, peel adhesion, cohesion, and aging resistance are used
as examination criteria. The main adhesive properties (tack, peel, and
cohesion) depend on the state of the polymer (pure or formulated), the
coating weight, and on the face stock used.
Adhesive Performance Characteristics 323

Tack
Common acrylic PSAs are not tacky enough and special acrylic PSAs
are not cohesive enough. Therefore acrylic PSAs need compounding or
tackifying. The tack of unformulated acrylic PSAs is lower than
that of tackified rubber-based PSAs (their main competitor) but higher
than that of the other untackified adhesives, except polyvinyl ethers [see
Eq. (6.14)].
The better tack level of unformulated acrylic PSAs, compared to other
unformulated PSAs (e.g., EVAc and CSBR) is illustrated in Table 6.28.
These data show that acrylic PSAs possess a better tack at any coating
weight level. In order to evaluate the tack of formulated acrylic PSAs,
the following examination criteria will be considered: ease of tackification,
the tackifier level, and the tack level.

Ease of Tackification/Compatibility. Acrylic PSAs display good


overall compatibility with common tackifiers (plasticizers or tackifying
resins). Both solvent-based and water-based tackifying resins can be fed into
water-based acrylic PSAs. It is not possible to add molten tackifiers to
water-based acrylic PSAs.

Tackifier Level. A level of 30–40% (by wet weight) tackifier resin


is commonly used in formulated acrylic-based PSAs and a tackifier level
of 10–20% suffices for tack improvement of acrylic-based PSAs (Table 6.29).
This is very important because adhesive formulations for soft films should
contain a minimum resin level in order to avoid resin/plasticizer interaction.
As discussed in detail in [271] acrylics need a lower tackifier concentration
for the maximum tack than CSBR or vinylacrylics and the tack maximum
(as a function of the tackifier concentration) is broader.

Tack Level. In general, the ease of tackifying and the obtained


tack level depend on the chemical composition, structure, and molecular
weight of the acrylic PSAs. With respect to the chemical composition and
structure of the polymer, there exists a broad range of possibilities. For
solvent-based acrylic PSAs viscosity is dependent on molecular weight; thus
the molecular weight constitutes a limiting factor. For water-borne acrylic
PSAs there is no technological molecular weight limit, but the molecular
weight should be limited because of compatibility reasons. There is an
optimum molecular weight in order to achieve good tackification. Table 6.30
illustrates the low tackifying response of high molecular weight acrylic
dispersions.
Acrylic-based PSAs need lower tackifier levels than CSBR and
EVAc copolymers. The tack level of tackified acrylic PSAs is superior
324

Table 6.28 Comparison of Acrylic PSAs Versus EVAc and CSBR PSAs. Tack and Peel of Unformulated WB PSA on Different
Chemical Basis, as a Function of the Coating Weight

Chemical Composition

AC EVAc CSBR

Peel (N/25 mm) Peel (N/25 mm) Peel (N/25mm)


Coating RB tack RB tack RB tack
Weight (g/m2) Glass PE (cm) Glass PE (cm) Glass PE (cm)

10 — — — — — — 10.0 PT 3.0 17.0


15 — — — 10.0 PT 1.0 18.5 10.0 PT 5.0 8.0
20 22.0 20.0 1.5 10.0 PT 1.0 17.0 12.0 PT 12.0 5.5
25 27.0 20.0 1.5 — — — 16.0 PT 14.0 4.9
30 — — — 14.0 PT 1.0 17.0 — — —
35 — — — 13.0 PT 1.0 15.0 — — —
40 21.0 14.0 1.5 15.0 PT 1.0 7.0 21.0 14.0 6.0
50 25.0 15.0 1.5 — — — — — —
55 35.0 PT 17.0 1.5 — — — — — —
60 35.0 PT 22.0 1.5 — — — — — —
AC: Acronal V205 (BASF); EVA: 1360 (Hoechst); CSBR: 3703 (Polysar); RB: rolling ball.
Chapter 6
Adhesive Performance Characteristics 325

Table 6.29 Tack as a Function of the


Tackifier Level

Tackifier Level (parts Tack (quick


per 100 PSA wet weight) stick; g/cm)

20 270
40 570
60 710
80 800
100 880
Quick stick measured for tackified water-
based PSA (25 g/m2 on PET).

Table 6.30 Tack Dependence on the Molecular Weight. Influence of a High


Molecular Weight Acrylic PSA on the Tack of a CSBR-Based, Tackified
Water-Based PSA Formulation.

Chemical Composition of
the PSA (parts, wet weight)

Compound code 3703 20094 CF 52 V205


Supplier Polysar Syntho A&W BASF RB Tack
Nature CSBR Acrylic Tackifier Acrylic (cm)

32 33 35 — 25.0
32 23 35 — 20.0
32 13 35 — 6.5
32 — 35 33 2.5
RB, rolling ball.

to that of tackified EVAc, but lower than the tack of tackified CSBR,
namely:
Tack : CSBRðformulatedÞ > ACðformulatedÞ > EVAc ðformulatedÞ
ð6:77Þ
as far as the tack properties are concerned. As discussed in Chap. 2, adhesive
properties also depend on the face stock material. A short examination of
the performance of acrylic-based PSAs in comparison with other adhesives
coated on paper and film, follows next.
Permanent/Removable Paper Label Applications. Generally, remova-
ble adhesives exhibit low tack. Common acrylic-based removable PSAs are
326 Chapter 6

untackified formulations. Rubber-based, solvent-based, or water-based


removable formulations, generally are tackified and thus they are tackier
than acrylic-based ones (Table 6.31). On the other hand, the removability
assumes a primer coating and/or a low coating weight.

Film Application. SBR has low light and aging stability and EVAc has
a limited water resistance. These disadvantages make them inadequate for film
coating applications. On the other hand, both need tack improvement (i.e., a
tackifier). The use of soft PVC as the main face stock material implies no or a
low tackifier level. Thus, only acrylic-based PSAs meet the whole range of
quality requirements for foil coating; they show the best unformulated tack,
another important feature for film-coating adhesives (Table 6.32).

Adhesion to Polar/Nonpolar Surfaces. Acrylic-based PSAs are the


best unformulated materials in the whole range of required peel forces
for polar substrates. Unfortunately, the peel of untackified acrylic PSAs on
polar surfaces is not sufficient (Table 6.33). The peel of other PSAs on
nonpolar substrates is low too, except for the new class of ethylene-vinyl
acetate-maleinate and ethylene-vinyl acetate-acrylic copolymers [272].

Peel of Tackified Acrylic PSAs


Tackification improves adhesion to polar and nonpolar surfaces. Polymers
like SBR and EVAc need more tackifier than acrylic PSAs (Table 6.34).
A detailed discussion about tack and required tackifier level is included in
Chap. 8. Formulations with less than 30 wt% tackifier do not provide
enough peel on polyethylene for common acrylic formulations (Table 6.35).
On the other hand, there are some new acrylic PSAs which display
adequate peel adhesion on polyethylene at low (20 wt%) tackifier loading
levels (Table 6.36).

Table 6.31 Rolling Ball Tack (cm) of Rubber-Resin-Based


Formulation Versus Acrylic PSA. Dependence of the Tack on
the Coating Weight

Chemical Composition of PSA

Coating Solvent-based Water-based


Weight (g/m2) rubber/resin tackified acrylic

8.6 4.0 6.4


18.4 2.1 2.4
28.0 1.9 1.9
Adhesive Performance Characteristics 327

Table 6.32 PSAs used for Film Coating

Chemical Composition Adhesive Characteristics

Parts Peel RB tack


Basis Code Supplier (wet) (N/25 mm) (cm)

Acrylic V205 BASF 100 0 5.0


V205 BASF 70 15.0 2.5
V208 BASF 30
PC80 UCB 100 6.0 3.5
FC88 UCB 100 8.0 3.5
ECR555 Exxon 100 5.0 1.0
FC88 UCB 80 — 3.5
PC80 UCB 20
FC88 UCB 70 10.0 6.5
880D BASF 30
V205 BASF 80 11.0 4.5
20084 Syntho 20
FC88 UCB 70 11.0 4.5
2313 Nobel 30
EVAc 1360 Hoechst 100 4.0 6.0
1360 Hoechst 70 — 3.0
1370 Hoechst 30
Acrylic-EVAc blend 913 UCB 50 — 7.0
1360 Hoechst 50
FC88 UCB 30 10.0 3.5
V205 BASF 30
1360 Hoechst 30
EVAc-acrylic EAF60 Wacker 100 — 10.0
VU895 Wacker 100 11.0 2.5
EVAc-maleinate Nacor 90 National 100 7.0 8.5

Table 6.33 Typical Peel Values of Untackified PSA on Polyethylene

Code Supplier Nature Peel (N/25 mm)

V205 BASF Acrylic 20.0


80D BASF Acrylic 3.8
85D BASF Acrylic 10.3
1360 Hoechst EVAc 5.0
EAF60 Wacker EVAc 7.0
3703 Polysar CSBR 7.0
20094 Syntho Acrylic 0.2
Instantaneous peel was measured; coating weight 20 g/m2.
328 Chapter 6

Table 6.34 Required Tackifier Level for AC PSA Versus CSBR PSA. The
Dependence of the Peel and Tack on the Tackifier Level

AC Formulation CSBR Formulation


Tackifier Level Peel (N/25 mm) Peel (N/25 mm)
(parts/100 RB tack RB tack
wet weight) Glass PE (cm) Glass PE (cm)

20 20.0 PT 15.0 2.5 14.0 PT 10.0 3.0


33 22.0 PT 17.0 2.5 15.0 PT 11.0 2.5
43 22.0 PT 20.0 PT 2.0 17.0 PT 16.0 PT 1.5
50 22.0 PT 19.0 1.5 19.0 PT 14.0 PT 1.5
AC: an 80/20 wet weight parts blend of Acronal V205/80D; CSBR: 3703 (Polysar); tackifier:
CF52(A&W); PT: paper tear; RB: rolling ball.

Table 6.35 Required Tackifier Level for AC PSA Formulations. Peel Dependence
on the Tackifier Level (Peel Adhesion on PE Film and Plate)

Peel (N/25 mm)

PE film PE plate
(coating weight, g/m2) (coating weight, g/m2)
Tackifier Level
(% wet weight) 10 20 10 20

16 6.0 2.0 23.0 6.0


20 15.0 5.0 20.0 7.0
24 20.0 5.0 15.0 5.0
29 17.0 8.0 20.0 PT 5.0
33 15.0 8.0 21.0 PT 20.0 PT

Table 6.36 Peel on Polyethylene for Low Level Tackified Special WB PSA

Chemical Composition, Parts (wet weight)

Code 5650 ECR-576 Exoryl 2001 SE-351


Supplier Hoechst Exxon Exxon A&W Peel on PE

Nature AC AC AC Tackifier (N/25mm)


100 — — — 16.5 PT
100 — — 20 19.5 PT
— 100 — — 17.5 PT
— 100 — 20 18.0 PT
— — 100 — 5.2
— — 100 20 15.4
Adhesive Performance Characteristics 329

Proposals from tackifier suppliers include the use of 25–50% tackifier


resin for aggressive, soft, acrylic dispersions on an ethyl hexyl acrylate basis
(Fig. 6.20). Other base elastomers require higher tackifier levels.
Figure 6.21 and Table 6.37 illustrate the low tackifying response of
CSBR-based PSAs as compared to acrylic-based PSAs. For common water-
based PSA formulations with standard adhesive performance character-
istics, EVAc and SBR require higher tackifier loadings than acrylic PSAs
(Fig. 6.21). In a similar manner tackified acrylic PSAs show a more
pronounced response to tackification than ethylene vinyl acetate copolymers
(Table 6.38).
Generally, higher peel adhesion is obtained with less tackifier for
acrylic PSAs; the absolute value (level) of the achieved peel also is higher
(Table 6.39). One can conclude that the peel of tackified acrylic PSAs on
polar or nonpolar surfaces is superior:

Peel : AC ðTackifiedÞ > CSBR ðTackifiedÞ >> EVAc ðTackifiedÞ


ð6:78Þ

Figure 6.20 Suggested tackifier level for PSA formulation on different chemical
bases. 1) Acrylic; 2) vinylacrylic; 3) SIS; 4) SIS; 5) SIS; 6) acrylic; 7) SBR; 8)
vinylacrylic; 9) EVAc.
330 Chapter 6

Figure 6.21 Ternary composition diagram. Common PSA formulation for


1) CSBR, 2) EVAc, and acrylic formulation. Water-based tackified PSA.

Table 6.37 Tackifying Response of Water-Based CSBR PSA Formulations

Chemical Composition, Parts (wet weight) Adhesive Characteristics

Peel (N/25 mm)


PSA 3703 Tackifier CF 52 RB tack
Polysar (A&W) On glass On PE (cm)

50 50 19.0 PT 15.0 PT 1.5


57 43 17.0 PT 16.0 PT 1.5
67 33 15.0 PT 11.0 2.5
80 20 14.0 10.0 3.0

Cohesion of Acrylic-Based PSAs


Except for some special adhesives with improved low temperature (and
room temperature) fluidity (e.g., deep-freeze adhesives), common unformu-
lated acrylic PSAs possess enough intrinsic shear strength to be converted
and processed without problems. Other polymers (e.g., EVAc, CSBR)
have better cohesion (measured as shear) than acrylic PSAs, but their
Adhesive Performance Characteristics 331

Table 6.38 Tackifying Response of Water-Borne EVA-Based PSA Formulations

Chemical Composition, Parts (wet weight) Adhesive Properties

Code 1360 1370 CF52 Peel RB tack


Supplier Hoechst Hoechst A&W (N/25mm) (cm)

50 50 100 7.0 >30


50 30 100 3.0 >30
70 50 100 15.0 PT 20
50 50 70 6.0 >30
50 30 70 3.0 >30
50 30 50 8.0 >30

Table 6.39 Peel/Tackifier Level Correlation for Water-Based PSAs on Different


Chemical Bases

PSA
Tackifier Peel
Code Supplier Nature Level Level (N/25 mm)

3703 Polysar CSBR 50 50 19.0 PT


3703 Polysar CSBR 66 33 15.0 PT
3703 Polysar CSBR 77 23 14.0 PT
Nacor 90 National EVAcM 50 50 20.0 PT
Nacor 90 National EVAcM 66 33 20.0 PT
Nacor 90 National EVAcM 77 23 18.0 PT
FC88 UCB AC 50 50 16.0
FC88 UCB AC 66 33 8.5
FC88 UCB AC 77 23 7.0
V205 BASF AC 40 50 19.0
80D BASF AC 10
V205 BASF AC 49.5 33 10.5
80D BASF AC 16.5
V205 BASF AC 57.75 23 9.0
80D BASF AC 19.25
3703 Polysar CSBR 25 50 27.0 PT
80D BASF AC 25
3703 Polysar CSBR 35 33 25.0 PT
80D BASF AC 35
3703 Polysar CSBR 38.5 23 23.0 PT
80D BASF AC 38.5
1360 HOE EVAc 50 50 15.0
1360 HOE EVAc 66 33 12.0
1360 HOE EVAc 77 23 6.5
The concentration (level) of the components is given in parts wet weight. Tackifier: CF52 from
A&W.
332 Chapter 6

shear-correlated properties like cuttability are not superior. As an example


for the upper limit, crosslinked, solvent-based acrylic PSAs exhibit shear
levels in excess of 260 hr [150]. On the other hand, soft, high tack acrylic
PSAs show 15 min shear times. Table 6.40 summarizes characteristic shear
values for PSAs on different chemical bases; it shows that acrylic PSAs
display the best shear values in the range of unformulated polymers with
similar tack. This statement is very important for film label applications,
where new ethylene-maleinate multipolymers display very good nonpolar
peel adhesion, but low cohesion.

Shear Strength of Tackified Acrylic PSAs


Common tackifier levels lead to a decrease in cohesion. This is a function of
the nature of the acrylic, the nature of the tackifier, and the mixing ratio.
Shear loss is less pronounced for CSBR and EVA. Table 6.41 illustrates
the variation of the shear through tackification for different classes of raw
materials. Unfortunately higher shear does not lead to better converting
properties for CSBR/EVAc.

Aging Resistance
For common PSAs generally, time and temperature stability of the
adhesive/laminate at room temperature is required. There are special end-
uses where high temperature stability or low temperature performance is
necessary. Table 6.42 summarizes data concerning UV and temperature
stability of different PSAs.
Because of their bonding ability (accentuated flow) common acrylic
PSAs show only a limited temperature stability in PSA labels. High
molecular weight or partially crosslinked acrylic PSAs display an improved
thermal stability; partial crosslinking may be obtained with special built-in
monomers (e.g., acrylonitrile). Therefore some common ethyl hexyl-based
acrylic PSAs with acrylonitrile as comonomer will display higher than
normal shear values. Because of the temperature resistance of these nitrile-
based crosslinks, especially high hot shear values are obtained.

Laminate Properties
Some aspects of the laminate properties are dealt with in Chaps. 2 and 8.
The adhesive/solid state component interactions influence the stability of
the laminate properties. First, these interactions may be restricted to mutual
interactions between adhesive and paper-based face stock material. The
most important mutual interaction takes place between the porous, rough
Table 6.40 Shear Values as a Function of the Physical State and Chemical Nature of the PSA

Chemical Characteristics

Chemical basis Crosslinked Physical State Shear Resistance

AC RR EVAc Yes No SB WB HM Value Unit Temperature (oC) Method Ref.

x x x 0.7–2 hr RT 0.500  0.500 , 500 g 2


x x x 5–10 min RT — 267
x x x >200 min RT — 267
x x x 40–90 min RT 0.500  0.500 , 500 g —
Adhesive Performance Characteristics

x x x 1 min 70 — 269
x x x 5000 min 70 — 269
x x x >10,000 min RT 0.500  0.500 , 500 g 46
x x x 260 hr RT — 264
x x x >200 min RT — 267
x x 500–4500 min RT — 46
x 3 hr RT 0.500  0.500 , 500 g 273
x 4.6 d RT — 265
x 50–400 min RT — 159
x x 300 min 95 — 20
x x 100 hr 88 — 274
x x 1 hr RT — 275
x x 1 hr RT — 276

AC: acrylic; RR: rubber-resin; EVAc: ethylene-vinyl acetate; SB: solvent-based; WB: water-based; HM: hot-melt; RT: room temperature.
333
334 Chapter 6

Table 6.41 Influence of the Tackification on the Shear of Water-Based PSA on


Different Chemical Bases

Tackifier Level

PSA Supplier Nature 20 25 30 35 40 45 50

V205 BASF AC 41 38 36 33 38 — 25
80D BASF AC 39 37 34 32 22 — 25
Hot shear 9 7 — 7 4 — —
RT shear 100 80 55 39 27 — 55
3703 Polysar CSBR 40 — — — 30 — 25
V205 BASF AC 20 — — — 22.5 — 19
80D BASF AC 20 — — — 7.5 — 8
Hot shear — — 19 22 13 — —
RT shear 250 — — — 25 — 14
3703 Polysar CSBR 40 37.5 35 — 30 22.5 25
80D BASF AC 40 37.5 35 — 30 22.5 25
Hot shear > 1500 > 1200 240 — 2 2 2
RT shear
3703 Polysar CSBR — — 70 — — 55 50
Hot shear
RT shear 1300 500 368

Shear data given in minutes. Hot shear was tested at 50 C/min. An acid rosin was used as
tackifier.
RT: room temperature; AC: acrylic.

Table 6.42 Aging Stability of PSA on Different Chemical Bases

Adhesive Properties

After aging

Adhesive Initial Thermal treatment

Peel Tack Peel Tack Light


Code Supplier Nature (N/25 mm) (cm) (N/25 mm) (cm) exposure

2382 Syntho AC 15.0 2.5 14.0 4.5 1


2343 Syntho VAc-AC 14.6 2.5 24.0 PT 2.5 —
1338 Syntho CSBR 20.4 5.5 15.0 8.0 5
10066 Syntho S-AC 22.0 7.5 20.0 10.0 5
1247 Syntho CSBR — — — — 5
EAF60 Wacker EVAc 25.0 PT 4.0 25.0 PT 6.0 1
V205 BASF AC 25.0 PT 2.5 25.0 PT 4.0 1
1 ¼ good, 5 ¼ poor; AC, acrylic; VAc, vinyl acetate; CSBR, carboxylated styrene-butadiene
rubber; S-AC, styrene acrylic; EVAc, ethylene vinyl acetate.
Adhesive Performance Characteristics 335

face stock material and the fluid adhesive (i.e., a physical interaction
between paper and the PSA); this is known as migration.

Migration Resistance. The migration resistance is a complex charac-


teristic. It depends on the adhesive, on the carrier, and on the manufacturing
method. As discussed in detail in [277] the parameters of the migration
include the nature, molecular weight, and gel content of the base elastomer,
the nature and softening point of the tackifier used, and the nature and
boiling point of the plasticizer. For dispersed systems the solids content, the
surfactant, and the thickener affect the migration too. The migration resis-
tance of untackified acrylic PSAs remains superior to other formulations,
with solvent-based ones better than water-based dispersions. Tackified acrylic
PSAs show better migration resistance than competitive raw materials:

Migration : AC ðTackifiedÞ  EVAc ðTackifiedÞ  CSBR ðTackifiedÞ


ð6:79Þ

There are some acrylic PSAs with lower migration resistance (e.g.,
removable and deep-freeze PSAs). A high emulsifier loading decreases the
migration resistance of water-based PSAs.

Anchorage. Anchorage is the phenomenon of adhesion between face


stock material and adhesive. For fiber forming, textured supports (e.g.,
paper) this is a physical phenomenon also. Physical interaction and flow of
the adhesive are the main components for the anchorage of PSAs onto rough,
textured substrates. Common, tacky, unformulated acrylic PSAs display
good anchorage on paper, even if transfer coated. Hard, low tack acrylic
PSAs do not exhibit enough anchorage on paper. Removable, unformulated
acrylic PSAs give less anchorage on unprimed surfaces than tackified
formulations. Anchorage also is a function of the emulsifier level; therefore
solvent-based acrylic PSAs display better anchorage than water-based ones.
Untackified EVAc or CSBR exhibit poorer anchorage than acrylic PSAs;
untackified acrylic PSAs give better anchorage than competitive materials.
Tackification improves the anchorage. Tackified acrylic PSAs and tackified
rubber display better anchorage than tackified EVAc. With respect to the
chemical interaction between acrylic PSAs and paper, untackified acrylic
PSAs do not interact chemically with normal paper. A similar behavior is
observed for competitive base polymers. Tackified formulations may interact
with special (e.g., thermal) papers, due to the tackifying resin.
Most plastic films are compounded materials (i.e., they contain
additives). Therefore there are possibilities for an interaction between
micromolecular components of the film (e.g., PVC or polyolefins) and the
336 Chapter 6

adhesive. Plasticizer and emulsifier from the film, slip agents, or antifogs
could migrate into the adhesive. On the other hand, monomers, oligomers,
and surface agents from the adhesive can migrate into the face material,
causing stiffening of the face stock (or its shrinkage) as well as the loss of the
adhesive properties. Because of their low UV stability rubber-resin adhesives
cannot be used for films. EVAc copolymers may interact with some film
materials and cause shrinkage. Thus only acrylic PSAs are recommended for
film coating. In a similar manner the anchorage of unformulated acrylic
PSAs on nonpolar, untreated film is better than the anchorage of its
competitors (except for ethylene-maleinate multipolymers and Nacor 90).
Acrylic-based PSAs exhibit varying shrinkage on soft PVC. There
are some leading acrylic PSA dispersions that show shrinkage values of
0.3–0.9% (Table 6.43). A detailed discussion of these characteristics is
presented in Chap. 10.
It is evident that adhesives interact with the release liner as well.
Acrylic PSAs with different formulations exhibit different release levels and
release level stability. Ethylene vinyl acetate-maleinate copolymers display
less sensitivity towards the nature and age of the release liner. Unformulated
acrylic PSAs display the best die-cuttability. In the range of formulated
products, crosslinked, tackified, solvent-based adhesives show the best die-
cuttability (see Chapters 8 and 9). Generally, acrylic formulations require
less surface active agents for stabilizing and wet-out, thus the humidity
content of the laminate is lower. Therefore acrylic-based laminates are
superior compared to CSBR/EVAc-based laminates with respect to lay-flat,
die-cuttability, and water resistance.

Special Features of Water-Based Acrylic PSAs as


Compared to Other Water-Based PSAs
Water-based PSAs show different characteristics according to their chemical
nature and manufacturing technology. The properties of the aqueous
dispersions as well as those of the coated layer and of the laminate vary
widely.
In general, water-based acrylic PSAs possess a higher solids content and
a higher solids content/viscosity ratio than common CSBR/EVAc disper-
sions; they also respond better to diluting. Water-based acrylic PSAs exhibit
a lower surface tension than common EVAc or CSBR dispersions. Current
surface tension values range from 34–40 mN/m, whereas EVAc dispersions
show values above 45 mN/m. Rubber-based dispersions have a surface
tension between 45–60 mN/m. Furthermore water-based acrylic PSAs show
higher coagulum (grit) than CSBR dispersions, although they display a better
mechanical stability than CSBR. Water-based acrylic PSAs generate less
Table 6.43 Shrinkage Values for PSA-coated PVC (Different PSAs)

PSA Test Conditions Shrinkage (%, Face Stock)

Nature White PVC Transparent PVC Test method


Temperature Time
Code Status Chemical basis ( C) (days) Longitudinal Transverse Longitudinal Transverse Mounted Unmounted Ref.

01 WB AC 70 7 0.20–0.30 0.15 — — x — —
02 WB AC 70 7 0.50 0.40 0.85 0.60 — x —
Adhesive Performance Characteristics

03 WB AC 70 7 — 0.40 0.85 0.50 — x —


04 WB AC 70 7 — — 1.70 1.20 — x —
05 WB AC 67 7 — — 6.00 — — x 278
05 WB AC 67 1 — — 17.0 — x — 278
06 SB AC 67 7 — — 8.0 — — x 278
06 SB AC 67 1 — — 11.0 — x — 278
07 WB EVAc 70 7 — — 0.50 0.40 x — 279
07 WB EVAc 70 7 — — 2.60 1.40 — x 279
08 WB AC 67 7 — — 0.40 0.40 — x 280
09 WB EVAc 70 7 — — 2.0–3.0 — — x 275
10 WB AC 70 7 — — 0.5–1.0 — — x 275

WB: water-based; SB: solvent-based; AC: acrylic; EVAc: ethylene vinyl acetate.
337
338 Chapter 6

foam formation than CSBR and EVAc. Finally, water-based acrylic PSAs
show better wet-out properties than EVAc- and CSBR-based PSAs, and can
be converted at higher speeds than CSBR and EVAc dispersions.

Special Features of Solvent-Based Acrylic PSAs


as Compared to Solvent-Based Rubber PSAs
Solvent-based acrylic PSAs have some special characteristics compared
to classical rubber-resin PSAs. First, a higher solids content (more than
20–30%) is possible and the solids content/viscosity ratio can be adjusted
exactly. Usually special solvents are used and both solutions or dispersions
can be produced. In general fewer or no plasticizers and antioxidants are
used; finally crosslinking is possible.

Modification of Adhesive Properties


The main factors influencing the adhesion are: the properties of the
adhesive, the properties of the carrier (face stock, liner), the construction
and geometry of the laminate, the coating weight, the coating technology,
and the age of the laminate. The adhesive properties of the PSAs can be
influenced through polymer synthesis, formulation, and converting.
Polymer Synthesis. There exists a broad range of different functional
acrylic/methacrylic monomers which can be used for PSA synthesis. Acrylic
PSAs have the broadest monomer basis. Some of the monomers impart
adhesive properties, others are used as emulsifier or stabilizer monomers.
A butyl acrylate (BuAc) base provides a higher cohesion level than ethyl
hexyl acrylate (EHA); other nonacrylic comonomers also can be used. For
example acrylonitrile imparts hardness and solvent resistance, styrene and
alpha-methyl styrene impart firmness and improve peel adhesion, t-butyl
styrene improves tack, methylmethacrylate makes the adhesive firmer, vinyl
acetate improves adhesion to certain plastic surfaces, etc. Compared to
acrylic-based PSAs, rubber- and EVAc-based copolymers possess a
narrower raw material basis.
EVAc-based PSAs must incorporate more than 18% ethylene, while
SBR-based PSAs contain less than 25–35% styrene. Thus, their formulating
freedom remains limited. Most vinyl acetate copolymers need protective
colloids as their particle size and particle size distribution are limited;
they have a predetermined viscosity and solids content/viscosity ratio, and
a built-in water resistivity. EVAc-based PSAs do not have crosslinked
structures, whereas acrylic-based PSAs may have them. Styrene-butadiene
rubber-based PSAs must always have a gel/sol ratio as their properties
(peel, zshear, migration) depend on this.
Adhesive Performance Characteristics 339

A pressureless well-known polymerization technology is available for


acrylic PSAs. Styrene-butadiene rubber PSAs need medium pressure and
EVAc PSAs require high pressure reactors. For solvent-based acrylic PSAs
a crosslinking technology was developed; for water-based acrylic PSAs
branching and grafting procedures were established.
Polymer Formulations. A general compatibility exists between
different acrylic PSAs, as well as between acrylic PSAs and other polymers.
However, no similar compounding freedom exists for rubber- or EVAc-
based polymers. Plasticizers and tackifiers can be used for acrylic PSAs and
other polymers. Crosslinking by built-in comonomers, metal complexes, or
PUR is possible for acrylic PSAs; EVAc PSAs have no similar crosslinking
possibilities. Attempts were made to crosslink CSBR using polyisocyanates.
Converting. Theoretically there are three converting technologies for
acrylic PSAs: hot-melt, solvent-based, and water-based. Practically, only a
few experimental hot-melt acrylic PSAs exist (like experimental hot-melt
EVAc-based PSAs). Rubber-based PSAs also can be converted using all of
the above technologies.
The coating technology is a function of the adhesive, face stock,
release nature, and of converting, processing, and end-use requirements.
Acrylic-, rubber-, and EVAc-based PSAs allow direct or transfer coating.
Acrylic solutions are coated directly on paper and special films, and
indirectly on soft PVC. Solvent-based rubber PSAs are directly or transfer
coated on paper. Water-based acrylic PSAs are directly or transfer coated
on films. Through direct/transfer coating, anchorage, die-cutting, water
resistance, and release can be adjusted.
High viscosity (and high solids) water-based and solvent-based acrylic
PSAs are available, as well as low viscosity (ready-to-use) ones. Primers
provide better anchorage and less migration, and increase the removability.
Rubber-based PSAs need primers; acrylic PSAs can be converted with/
without primer. Common acrylic dispersions need more wetting agents
for transfer coating than for direct coating. They need more wetting agents
on solventless silicones than on solvent-based silicones. Wet-out depends on
the siliconized paper. The age of the siliconized liner influences the wetting
also. Anchorage of the coated adhesive also depends on the age of the
laminate.
Special Features of Acrylic PSAs. Next the main attributes of acrylic
PSAs are highlighted. From the point of view of the end-user acrylic PSAs
provide the best balance between adhesive and cohesive properties, the best
independence from the face stock and substrate, the best aging properties
(e.g., oxidative stability, heat resistance, UV stability), and the best tack
340 Chapter 6

at varying temperatures. In addition, they exhibit the best machinery


properties (adhesion/release balance), as well as FDA and BGA approval.
The advantages of acrylic PSAs for the converter include a broad range
of raw materials, a broad compatibility with other polymers, and different
tackifying possibilities (e.g., high polymers, resins, plasticizers). They accept
a broad range of tackifying resins, and display at low tackifier level a more
pronounced tackifier response. They are generally suitable for different face
stocks and for permanent and removable applications in rolls and sheets.
Acrylic PSAs possess excellent converting properties (e.g., wetting out,
drying speed, low foam, high mechanical stability) and allow many different
ways for adjusting peel adhesion and shear.
The disadvantages of acrylic PSAs for the end-user are relatively
lowadhesion on untreated, nonpolar surfaces, relatively low die-cuttability if
formulated for high tack and coated on lightweight laminate components,
and relatively low water resistance for water-based PSAs. Disadvantages of
acrylic PSAs for the converter include the lack of available acrylic hot-melt
PSAs and the lack of or limited compatibility with rubber-based adhesives.
Solvent-based acrylic PSAs are superior to water-based PSAs.
Although acrylic PSAs are the most expensive PSAs, they are used
with increasing dynamism. There are only a few manufacturers formulating
monomers for acrylic PSAs (mainly for EHA and BuAc). The most
important component for acrylic PSAs are the tackifier resins. In this field
there is an abundant supply of new and classical products.
Formulating Ease and Costs. Acrylic PSAs are multipolymers, so
their properties can be varied using different monomer feeds and
polymerization technologies. Solvent-based technology is more expensive
but has the advantage of crosslinking (higher quality at a higher price level).
Water-based technology presents the advantage of high solids, seed and
branching, and particle size distribution control. Depending on the end-use
requirements, it is possible to synthesize or buy acrylic PSAs for
compounding, resulting in more or less expensive products. Acrylic PSAs
can be tackified using a medium level of tackifier loading only; they have
a reserve of adhesive properties and theoretically they can be modified
with fillers.
The price of the laminate depends on the following components: the
face stock, the release liner, the adhesive, and the choice of the converting
(laminating) technology.
Paper. Tackified acrylic PSAs show good anchorage and low
migration (solvent-based or water-based), thus they also can be used with
inexpensive non-surface-treated paper. On the other hand, water-based
acrylic PSAs need special (primered) face stock in order to obtain a good
Adhesive Performance Characteristics 341

die-cuttability. Heavy, voluminous papers increase the die-cuttability also.


Removable water-based acrylic PSAs also need primered papers.
Film. Water-based and solvent-based acrylic PSAs can be used with-
out restrictions for different film materials. The same acrylic can be used for
paper as well. Direct coating on nonpolar films requires a surface treatment.
Acrylic PSAs can be coated on paper-based release liners as well as on
film-based release materials. Solvent-based or solventless siliconizing can
be used. Direct/transfer coating provides another possibility for release level
modification. Transfer coating of water-based acrylic PSAs on solventless-
silicones remains the cheapest way.
Solvent-based acrylic PSAs are the most expensive adhesives. Hence:
The use of filled solvent-based acrylic PSAs decreases the adhesive
costs.
Compounding solvent-based acrylic PSAs also decreases the
adhesive cost.
Tackifying solvent-based/water-based adhesives decreases the
costs.
Common water-based acrylic PSAs are less expensive than solvent-
based ones (price ratio 2/1).
Tackified water-based acrylic PSAs are the most cost-effective
materials within the range of acrylic PSA adhesives.

Converting Costs. Converting acrylic PSAs is possible in solution


or as an aqueous dispersion. Converting solvent-based acrylic PSAs is
more expensive than water-based acrylic PSAs, whereas converting solvent-
based rubber-resin PSAs is less expensive than acrylic PSAs. Converting
low viscosity water-based acrylic PSAs costs less than coating highly
viscous ones. As pointed out before, acrylic PSAs can be converted directly
or by transfer, and transfer coating remains less expensive than direct
coating.

REFERENCES

1. A. Zosel, Colloid Polymer Sci., 263, 541 (1985).


2. C.M. Chum, M.C. Ling and R.R. Vargas (Avery Int. Co, USA), EP 12257927/
18.08.1987.
3. J. Johnston, Adhes. Age, (11) 30 (1983).
4. Glossary of Terms Used in Pressure Sensitive Tapes Industry, PSTC, Glenview,
IL, 1974.
5. ASTM, D-1878–6 IT; Pressure-Sensitive Tack of Adhesives, American Society
for Testing and Materials, Philadelphia.
6. M. Barquins and D. Maugis, J. Adhesion, 13, 53 (1981).
342 Chapter 6

7. D. Ahn and K.R. Shull, Macromolecules, 29, 4381 (1996).


8. K.L. Johnson, K. Kendall and D. Roberts, Proc. Roy. Soc. Lond., A 324, 301
(1971).
9. K.R. Shull, D. Ahn, W.L. Chwen, C.M. Flanigan and A.J. Crosby, Macromol.
Chem. Phys., 199, 489 (1988).
10. H. Mizumachi, J. Appl. Polymer Sci., 30, 2675 (1985).
11. H. Barquins, Kautschuk, Gummi, Kunststoffe, (5) 419 (1987).
12. J.A. Greenwood, ‘‘Viscoelastic Fracture and Healing,’’ Proc. 24th Annual
Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.14.
13. H. Barquins, R. Courtel and D. Maugis, Wear, 38, 385 (1976).
14. K.A. Grosch, Proc. Roy. Soc., A274, 21 (1963).
15. Y.Y. Lin and C.Y. Hui, ‘‘Modeling the Failure of an Adhesive Layer in a Peel
Test,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg,
VA, p.230.
16. J. Wiedemeyer, VDI Berichte, 600 (3) 72 (1987).
17. G.R. Hamed and C.H. Hsieh, J. Polymer Physics, 21, 1415 (1983).
18. A. Roos and C. Creton, ‘‘Adhesion of PSA Based on Styrenic Block
Copolymers’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001,
Williamsburg, VA, p.371.
19. P. Tordjeman, E. Papon and J.J. Villenave, ‘‘The Mechano-Optical Tack
Tester. A New Apparatus to Study the Tack Properties of PSA,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg, VA, p.424.
20. A.N. Gent and A. J. Kinloch, J. Polymer Sci., A-2 (9) 659 (1971).
21. A.N. Gent and J. Schultz, J. Adhesion, 3, 281 (1972).
22. C. H. Andrews and A. J. Kinloch, Proc. Roy. Soc., A332, 385 (1973).
23. A. Aymonier, E. Papon, J.J. Villenave and P. Tordjeman, ‘‘Direct Relation
Between Copolymerization Process and Tack Properties of Model PSA’s,’’
Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg,
VA, p.280.
24. P. Tordjeman, E. Papon and J.J. Villenave, J. Polym. Sci., Polymer Physics,
Part B, 38, 1201(2000).
25. S.C. Misra, C. Pichot, M.S. El-Aasser and J. W Vanderhoff, J. Polymer Sci.,
Polym. Lett., 127, 567 (1979).
26. C. Pichot, M.F. Llauro and O.T. Pham, J. Polym. Sci. Polym Chem., 19,
2619 (1981).
27. M.S. El-Aasser, T. Makgawinata and J.W. Vanderhoff, J. Polymer Sci.,
Polymer Chem., 21, 2363 (1983).
28. S.C. Misra, C. Pichot, M.S. El-Aasser and J.W. Vanderhof, J. Polymer Sci.
Polym. Chem., 21, 2383 (1983).
29. M. Okubo, Makromol. Chem., Makromol. Symp., 35/36, 307 (1990).
30. A.J. Crosby and K.R. Shull, ‘‘Debonding mechanisms of PSAs,’’ Proc. of the
22nd Annual Meeting of the Adhesion Society, Febr. 21, 1999, Panama City
Beach, FL, p.320.
31. D. Maugis and M. Barquins, J. Phys. D; Appl. Phys., 11, 1989 (1978).
32. C. Gay and H. Leibler, Phys. Rev. Lett., 82, 936 (1999).
Adhesive Performance Characteristics 343

33. C. Creton and H. Leibler, J. Polym. Sci., Polymer Physics, Part B, 34, 545
(1996).
34. R.S. Chang, Rubber Chem. and Technol., (20) 847 (1957).
35. J. Johnston, Adhes. Age, (11) 30 (1983).
36. S.N. Can, D.R. Burfield and K. Sogh, Macromolecules, (18) 2684 (1985).
37. Adhes. Age, (9) 19 (1986).
38. Shell, Technical Bulletin, Polybutenes.
39. M.A. Krecenski and J.F. Johnson, Polym. Eng. Sci., 29 (1) 36 (1989).
40. C. Creton, in Processing of Polymers, H.E.H. Meijer (Ed.) VCH, Weinheim,
1997, Chapter 15, pp.707–741.
41. A. Midgley, Adhes. Age, (9) 17 (1986).
42. S. Tobing and D. Klein, J.Appl.Polym. Sci., 79, 2230 (2001).
43. N. Willenbacher and A.E. O’Connor, ‘‘Effect of Molecular Weight and
Temperature on the Tack of Model PSAs from Polyisobutene,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.378.
44. D.J. Harrison, J.F. Johnson and J.F. Yates, Polymer Eng. Sci., (14) 865 (1992).
45. Adhes. Age, (10) 24 (1977).
46. P.A. Mancinelli, ‘‘New Developments in Acrylic HMPSA Technology,’’ in
TECH12, Advances in Pressure Sensitive Tape Technology, Technical Seminar
Proceedings, Itasca, IL, May, 1989, p.165.
47. Polysar, Product and Properties Index, Arnhem, 02/1985.
48. Adhes. Age, (5) 32 (1987).
49. Adhes.Age, (9) 37 (1988).
50. E.G. Ewing and J.C. Erickson, Tappi J., (6) 158 (1988).
51. Dow Corning, Silicone PSA, Application Information, 1986.
52. Adhes. Age, (10) 16 (1977).
53. J. Ouyang, S. Jacobson, L. Shen, S. Reedell, ‘‘Characterization of Acrylic
PUR-Based Waterborne PSAs,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr.25, 2001, Williamsburg, VA. p.236.
54. S.A. Parrie and P.F. Ritchie, Adhesion, (5) 201 (1967).
55. Pack Report, (10) 15 (1986).
56. J.A. Greenwood and K.L. Johnson, Philosophical Magazine, A 43, 697 (1981).
57. J. Johnston, Adhes. Age, (12) 24 (1983).
58. A. Bates, Adhes. Age, (12) 26 (1983).
59. S.R. Aubuchon, R. Ulbrich and S.B. Wadud, ‘‘Characterization of Pressure-
Sensitive Adhesives and Thermosets by Controlled Stress Rheology and
Thermal Analysis,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001,
Williamsburg, VA, p.427.
60. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 5.
61. A. Kamagata, T. Saito and M. Toyama, J. Adhesion, (2) 279 (1972).
62. M. Sherrif, R.W. Knibbs and P.G. Langley, J. Appl. Polymer Sci., 17,
3423 (1973).
63. P.J. Counsell and R.S. Whitehouse, Development in Adhesives (W.C. Wake,
Ed.), Vol. 1, Applied Science Publishers, London, 1977, p. 99.
344 Chapter 6

64. W.R. Dougherty, 15th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1990, p.70.
65. D.H. Kaelble, Trans. Soc. Rheology, (2) 135 (1965).
66. A. Schultz, A. Siddiqui, E. Kleinfeld, L.R. Farwaha Phan and P. Hayes, US
Pat. 5928783 (2000); in A.K. Schultz and N. Kofira, ‘‘The Use of Non-
Migrating Surfactants in Pressure-Sensitive Adhesive Applications,’’ Proc. 24th
Annual Meeting of Adh. Soc., Feb.25, 2001, Williamsburg, VA, p.163.
67. H.W. Yang, ‘‘Effect of Polymer Structural Parameters on PSAs,’’ Proc. of the
22nd Annual Meeting of the Adhesion Society, FL, Febr.21, 1999, Panama City
Beach, p.63
68. I. Benedek, Adhäsion, (5) 16 (1986).
69. T.K. Sherwood, Ind. Eng. Chem., 21, 12 (1929).
70. J. Hansmann, Adhäsion, (4) 21 (1985).
71. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.
72. W.D. Freeston, Jr., Coated Fabrics Technology, Technonic Publ. Co.,
1973, p.25.
73. S.S. Hwang, Chem. Eng. Sci., 34, 181 (1979).
74. Backofen & Meier AG, Bulach, Reverse Gravure Coating of Emulsion PSA,
Private Communication.
75. F. Weyres, Coating, (4) 110 (1985).
76. H.D. Patermann, Coating, (6) 224 (1988).
77. E. Brada, Papier u. Kunststoffverarb., (5) 170 (1986).
78. Adhäsion, (10) 399 (1969).
79. Die Herstellung von Haftklebstoffen, Tl.2.2, 15d, Nov. 1979, BASF,
Ludwigshafen, Germany.
80. Adhäsion, (9) 349 (1965).
81. Coating, (11) 4 (1988).
82. Coating, (11) 380 (1986).
83. C.P. Iovine (National Starch Chem. Co., USA), EP 0212358
A2/04.03.1987, p. 3.
84. H. Müller, J. Türk and W. Druschke, BASF, Ludwigshafen, EP 0, 118,
726/11.02.1983.
85. J.A. Fries, Tappi J., (4) 129 (1988).
86. Hyvis, Prospect, British Petrol, (1985).
87. Tappi J., 67 (9) 104 (1983).
88. P. Caton, European Adhes. Seal., (12) 18 (1990).
89. J. Boutillier, llth Munich Adhesive and Finishing Seminar, Munich, Germany,
1983, p.3.
90. Lihua Li, C. Macosko, G.L. Corba, A. Pocius and M. Tirrell, ‘‘Interfacial
Energy and Adhesion Between Acrylic PSA and Release Coatings,’’ Proc. 24th
Annual Meeting of Adh. Soc., Feb.25, 2001, Williamsburg, VA, p.270.
91. C. Verdier and J.M. Piau, ‘‘Understanding Peeling of PSAs by Use of
Visualization,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society,
Myrtle Beach, SC, Febr.20, 2000, p.116.
Adhesive Performance Characteristics 345

92. A.A. Chalykh, A.E. Chalykh, V. Yu Stepanenko and M.M. Feldstein,


‘‘Viscoelastic Deformations and the Strength of PSA Joints Under Peeling,’’
Proc. of the 23rd Annual Meeting of the Adhesion Society, Myrtle Beach, SC,
Febr.20, 2000, p.252.
93. B. Zhao, E. Miasek and R. Pelto, ‘‘Peeling Tapes from paper,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.364.
94. G.R. Hamed and W. Preechatiwong, ‘‘Peel Adhesion of Uncrosslinked
Styrene Butadiene Rubber Bonded to PET,’’ Proc. of the 23rd Annual Meeting
of the Adhesion Society, Myrtle Beach, SC, Febr.20, 2000, p.511.
95. D.J. Yarusso, J. Ma and R.J. Rivard, ‘‘Properties of Polyisoprene Based
PSAs Crosslinked by Electron Beam Radiation,’’ Proc. of the 22nd Annual
Meeting of the Adhesion Society, Panama City Beach, FL, February 21,
1999, p.72.
96. Johnston, ‘‘Alternate Methods for the Basic Physical Testing of PSA Tapes
and Proposals for the Future,’’ Proc. of the 23rd Annual Meeting of the
Adhesion Society, Myrtle Beach, SC, Febr.20, 2000, p.327.
97. S.F. Christensen, ‘‘Rheology of Adhesion,’’ Proc. of the 22nd Annual Meeting
of the Adhesion Society, Panama City Beach, FL, Febr.21, 1999, p.66.
98. K. Kendall, ‘‘Molecular Adhesion and Elastic Deformations,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.11.
99. J.W. Obreimoff, Proc. R. Soc., A127, 290 (1930).
100. K.L. Johnson, Brit. J.Appl. Phys., 9, 199(1958).
101. M.B. Taub and R.H. Dauskardt, ‘‘The Effects of Crosslink Density and
Environment on the Adhesion of Pressure Sensitive Adhesives with
Application in Transdermal Drug Delivery,’’ Proc. 24th Annual Meeting of
Adh.Soc., Febr.25, 2001, Williamsburg, VA, p.141.
102. J. Kinloch, H. Hadavinia, B.R.K. Blackman; M. Ring-Groth, J.-G. Williams
and E.P. Busso, ‘‘The Peel Behaviour of Adhesive Joints,’’ Proc. of the 23th
Annual Meeting of the Adhesion Society, Myrtle Beach, SC, Febr.20, 2000, p.25
103. J.M. Piau, G. Ravilly and C. Verdier, ‘‘Experimental and Theoretical
Investigations of Model Adhesives Peeling,’’ Proc. 24th Annual Meeting
Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.287.
105. M. Mazurek (Minnesota Mining and Manuf.Co., St.Paul, MN, USA), EP
4.693.935/15.09.1987.
106. Adhäsion, (10) 401 (1968).
107. European Adhesives and Sealants, (9) 4 (1987).
108. R.P. Wool and I. Lee, ‘‘Polymer-Solid Interfaces; Role of the Sticker and
Receptor Groups,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001,
Williamsburg, VA, p.297.
109. L. Li, M. Tirrell, A.V. Pocius and G.L. Korba, ‘‘Adhesion Studies on Acrylic
PSAs: Temperature and Composition Effect,’’ Proc. of the 23rd Annual
Meeting of the Adhesion Society, Myrtle Beach, SC, Febr.20, 2000, p.31.
110. E.J. Chang, Adhesion, 34, 189 (1991).
111. J.B. Class and S.G. Chu, J. Appl. Polym Sci., 30, 815(1985).
112. H. Yang, J. Appl.Polym. Sci., 55, 645 (1995).
346 Chapter 6

113. R. Sweet and K. Ulman, CRS Symposium 1997.


114. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
‘‘Rheology in the Release of Silicone Coatings,’’ in Book of Abstracts, 215th
ACS National Meeting, Dallas, March, 29 April (1998).
115. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
Adhes. Age, (11) 35 (1998).
116. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
‘‘Release Force control: The Bottom line Õ .Synergism Between Adhesive and
Liner,’’ Proc. of the Pressure-Sensitive Tape Council, 1999, Technical Seminar,
Washington DC, May 5, 1999.
117. G.V. Gordon, M.J. Owen, M.S. Owens, S.V. Perz, J.L. Stasser
and J.S. Tonge: ‘‘Interfacial and Bulk Contribution to Adhesive Release,’’
in Proc. of the 22nd Annual Meeting of the Adhesion Society, Panama City,
FL, Febr.21, 1999.
118. J.J. Bikermann, Trans. Soc. Rheol., 2, 9 (1957).
119. K. Fukuzawa and T. Uekita, ‘‘The Mechanism of Peel Adhesion for PSA
Tape,’’ Proc. of the 22nd Annual Meeting of the Adhesion Society, Panama
City Beach, FL, Febr. 21, 1999, p.63.
120. Firestone, Technical Service Report, 6110, August, 1986.
121. G.R. Hamed and W. Preechatiwong, ‘‘Anomalous Time-Temperature-
Response of the Peel Adhesion of Crosslinked-Epoxidized Natural Rubber
to Polyester Film,’’ Proc. of the 22nd Annual Meeting of the Adhesion Society,
Panama City, FL, Febr.21, 1999, p.304.
122. S. Tobing et al., ‘‘Molecular Parameters and Their Relation to the Adhesive
Performance,’’ J. Appl. Polyner Sci., 2001 in press; in S.D. Tobing and
A. Klein, ‘‘Synthesis and Structure Property Studies in Acrylic Pressure-
Sensitive Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc., Feb.25, 2001,
Williamsburg, VA, p.131.
123. A.W. Aubrey, G.N. Welding and T. Wong, J. Appl. Chem., (10) 2193 (1969).
124. E. Gronewaldt, Coating, (5) 144 (1969).
125. M.E. Fowler, Polymer, (11) 2146 (1987).
126. H. Lakrout, C. Creton, D. Ahn and K. Shull, ‘‘Adhesion of Monodisperse
Acrylic Polymer Melts to Solid Surfaces,’’ Proc. of the 23rd Annual Meeting of
the Adhesion Society, Myrtle Beach, SC, Febr.20, 2000, p.46.
127. Adhes. Age, (12) 19 (1986).
128. M.B. Taub, and R.H. Dauskardt, ‘‘Adhesion and Debonding of PSA used in
Transdermal Drug Delivery Systems,’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr.,25, 2001, Williamsburg, VA, p.141.
129. D.W. Aubrey, G.N. Weldung and T. Wong, J. Appl. Polymer Sci., 13, 2193
(1969).
130. M.M. Feldstein, A.E. Chalykh, R. Sh. Vartapetian, S.V. Kotomin, D.F.
Baraimov, T.A. Borodulina, A.A. Chalykh and D. Geschke, ‘‘Molecular
insight into Rheological and Diffusion Determinants of Pressure-Sensitive
Adhesion,’’ Proc. of the 23th Annual Meeting of the Adhesion Society, Myrtle
Beach, SC, Febr.20, 2000, p.54.
Adhesive Performance Characteristics 347

131. M.M. Feldstein, A.E. Chalykh, A.A. Chalykh, G. Fleischer and R.A. Siegel,
Polym. Mater. Sci. Eng., 81, 467 (1999).
132. Die Herstellung von Haftklebstoffen, Tl-2, 2–14d, Dec. 1979, BASF,
Ludwigshafen, Germany.
133. M. Schlimmer, Adhäsion, (4) 8 (1987).
134. G. Henke, ‘‘Schmelzhaftklebstoffe fur non wovens,’’ 9th Munich Adhesive and
Finishing Seminar, Munich, Germany, 1985.
135. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 2.
136. P.D. Hyde and D.J. Yarusso, ‘‘ATR-IR Studies of Strain Induced
Crystallized Natural Rubber Pressure-Sensitive Adhesives,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg, VA, p.433.
137. D.E. Yarusso and P.D. Hyde (Minnesota Mining and Manuf. Co, St. Paul,
MN) US Pat., 5, 866, 249.
138. Adhes. Age, (6) 32 (1986).
139. Adhäsion, (6) 24 (1982).
140. J. Kirn, K.S. Kirn and Y.H. Kim, J. Adhesion Sci. Technol., (3) 175 (1989).
141. J. Skeist, Handbook of Adhesives, 2nd Edition, Van Nostrand-Rheinhold Co.,
New York, 1977.
142. E.P. O’Brien and T.C. Ward, ‘‘Characterization of thin films: Shaft Loaded
Blister Test,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society, Myrtle
Beach, SC, Febr.20, 2000, p.202
143. D.H. Kaelble, Trans. Soc. Rheol., 3, 161 (1960).
144. Kai Tak Wan and Yiu-Wing Mai, Internat. J. of Fracture, 74, 181 (1995).
145. A.J. Kinloch, B.R.K. Black, H. Hadavinia, M. Paraschi and J.G. Williams,
Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg,
VA, p.44.
146. K.S. Kim and N. Aravas, Int. J. Solid Structure, 24, 417 (1988).
147. I. Benedek, ‘‘Bond Failure in Pressure-Sensitive Removable Thin Plastic Film
Laminates,’’ Proc. of the 22nd Annual Meeting of the Adhesion Society,
Panama City, FL, Febr. 21, 1999, p.448.
148. J. Wilken, Allg. Papier Rundschau, (42) 1490 (1986).
149. E.O. Brien, T.C. Ward, S. Guo and D. Dillard, ‘‘Characterizing the
Adhesion of Pressure-Sensitive Tapes Using the Shaft Loaded Blister Test’’
Proc. 24th Annual Meeting of Adh. Soc., Feb. 25, 2001, Williamsburg,
VA, p.113.
150. Z. Miyagi, S. Zaghi and D. Hunston, ‘‘The Sandwich Bending Specimen for
Characterizing Adhesive Properties,’’ Proc. of the 22nd Annual meeting of the
Adhesion Society, Panama City, FL, Febr. 21, 1999, p.119.
151. Adhäsion, (5) 202 (1986).
152. Adhäsion, (8) 237 (1976).
153. Coating, (1) 35 (1989).
154. Coating, (10) 271 (1988).
155. Allg. Papier Rundschau., (14) 340 (1987).
156. W. Druschke, Adhäsion, (5) 30 (1987).
348 Chapter 6

157. J.J. Elmendorp, D.J. Anderson and H. De Koning, ‘‘Dynamic Wetting Effects
in Pressure Sensitive Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr. 25, 2001, Williamsburg, VA, p.267.
158. M. Nardin and I.M. Ward, Materials Science and Technology, 3, 814 (1987).
159. C.Y. Hui, Y.Y. Lin and E.J. Kramer, ‘‘The Mechanics of Contact and
Adhesion of Periodically Rough Surfaces,’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr. 25, 2001, Williamsburg, VA, p.340.
160. A.N. Gent, G.R. Hamed and W.J. Hung, ‘‘Adhesion of Elastomers: Dwell
Time Effects and Adhesion to Fillers,’’ Proc. of the 23th Annual Meeting of the
Adhesion Society, Myrtle Beach, SC, Febr., 20, 2000, p.6.
161. A. Paiva, M.D. Foster, A.J. Crosby and K. Shull, ‘‘Studying Changes in
Surface Adhesion of PSAs with AFM and Spherical Indenter Test,’’ Proc. of
the 23th Annual Meeting of the Adhesion Society, Myrtle Beach, SC, Febr.20,
2000, p.43.
162. A.J. Kinloch, Adhesion and Adhesives, Chapman and Hall, London, 1987,
Chapter 2.
163. A.E. Chalykh, ‘‘Diffusion in Polymer Systems,’’ Chemistry, Moscow, 198; in
A.E. Chalykh, A.A. Chalykh, D.F. Barainov, T.A. Borodulina, and M.M.
Feldstein, ‘‘Contribution of PSA Polymer Creep Recovery to the Kinetics of
Adhesive Bonding,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society,
Myrtle Beach, SC, Febr.20, 2000, p.127.
164. H.P. Schreiber, ‘‘Polymer Diffusion Processes and the Evolution of Adhesive
Bond Strength,’’ Proc. 25th Annual Meeting of Adh.Soc., and the Second
World Congress on Adhesion and Related Phenomena, Febr. 10, 2002, Orlando,
FL, p.435.
165. A.A. Chalykh, A.E. Chalykh and M.M. Feldstein, Polym. Mater. Sci. Eng.,
81, 456 (1999).
166. J.C. Hooker, C. Creton, P. Torjemann and K.R. Shull, ‘‘Surface Effects on
the Microscopic Adhesion of Styrene-Isoprene-Styrene-Resin PSAs,’’ Proc. of
the 22nd Annual Meeting of the Adhesion Society, Panama City Beach, FL,
Febr., 21, 1999, p.415.
167. L. Benyahia, C. Verdier and J.M. Piau, J. Adhesion, 62, 45 (1997).
168. G. Bonneau and M. Baumassy, ‘‘New Tackified Dispersions for Water-Based
PSAs for Labels,’’ 19th Munich Adhesive and Finishing Seminar, 1994,
Munich, Germany, p.82.
169. G. Meinel, Papier u.Kunststoff Verarbeiter, (5) 177 (1989).
170. M. Cicotti, B. Giorgini and M. Barquins, Internat. J. of Adhesion and
Adhesives, (1) 35 (1998).
171. A. Carré and J. Schultz, J. Adhesion, 17, 135 (1984).
172. H.D. Brooks, J.Y. Kelly, P.H. Madison, C.D. Thacher and T.E. Long,
‘‘Synthesis and Characterization of Acrylamide Containing Polymers for
Adhesive Applications,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.25,
2001, Williamsburg, VA, p.150.
173. J.C. Conti, E.R. Strope and E. Jones, Proceedings of the 20th Annual Meeting
of the Adhesion Society, 425 (1997).
Adhesive Performance Characteristics 349

174. J.C. Conti, E.R. Strope, E. Jones and D. Rohde, Proceedings of the 21st
Annual Meeting of the Adhesion Society, 418 (1998).
175. J.C. Conti, E.R. Strope, R.D. Gregory and P.A. Mills, ‘‘Cyclic Peel
Evaluation of Sterilized Medical Packaging,’’ Proc. of the 22nd Annual
Meeting of the Adhesion Society, Panama City, FL, Febr. 21, 1999, p.116.
176. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 8.
177. G. Hombergsmeier, Papier u. Kunststoffverarb., (11) 31 (1985).
178. L.M. Prytikin and W.M. Wacula, Adhäsion, (12) 44 (1983).
179. Adhes. Age, (7) 36 (1986).
180. A. Dobmann and H. Braun, ‘‘Neues Engineering Konzept für
Etikettenindustrie,’’ 9th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1985.
181. B. Zhao, E. Miasek and R. Pelto, ‘‘Peeling Tapes from Paper,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg, VA, p.364.
182. J.J. Bikermann and W. Whitney, Tappi, 46 (7) 420 (1963).
183. T. Yamauchi, T. Cho, R. Imamura and K. Murakami, Nordic Pulp Paper J.,
4, 128 (1988).
184. R. Pelton, W. Chen, H. Li and M. Engel, ‘‘The Link Between Paper
Properties and PSA Adhesion,’’ Proc. of the 23rd Annual Meeting of the
Adhesion Society,’’ Myrtle Beach, SC, Febr., 20, 2000, p.127.
185. M.S. Kafkalidis, M.D. Touless and S.M. Ward, ‘‘Toughness of Adhesive
Joints with Elastic and Plastic Substrates,’’ Proc. of the 22nd Annual Meeting
of the Adhesion Society, Panama City, FL, Febr. 21, 1999, p.214.
186. ASTM D 1876–61T
187. European Standard NMP 458, No. 16–94D.
188. G. Josse, O. Poizat and C. Creton, ‘‘Probe Tack Tests of PSAs on Silicone
Release Coatings,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society,
Myrtle Beach, SC, Febr., 20, 2000, p.36.
189. I. Benedek, Pressure Sensitive Formulation, VSP, Utrecht, 2000, Chapter 2.4.1.
190. US. Pat., 3,691,140/12.09.1972; in H. Miyasaka, Y. Kitazaki, T. Matsuda and
J. Kobayashi (Nichiban Co. Ltd., Tokyo, Japan), Offenlegungsschrift, DE
3544868A1/18.12.1985.
191. Bevitack, Technical Bulletin, Bergvik-Arizona, 1990.
192. Z. Czech, European Adhes.Seal., (6) 4 (1955).
193. W. Druschke, Adhäsion, (6) 27 (1987).
194. P. Holl, Verpackungs-Rundschau, (3) 220 (1986).
195. Adhes. Age, (5) 24 (1987).
196. J.C. Fitch and A.M. Snow Jr., Adhes. Age, (10) 23 (1977).
197. Novamelt Adhesive Technology, Technical Bulletin, (1994).
198. A.C. Makati, Tappi J., (6) 147 (1988).
199. M.H. Tanashi, K. Yasuaki, M. Tetsuaki and K. Yunichi (Nichiban Co., Ltd.,
Tokyo, Japan), OS/DE 3,544,868 Al/15.05.1985.
200. G.L. Schneeberger, Adhes. Age, (4) 21 (1974).
201. J. Lin, W. Wen and B. Sun, Adhäsion, (12) 21 (1985).
350 Chapter 6

202. Coating, (4) 123 (1988).


203. Coating, (1) 13 (1970).
204. H. Reip, Adhes. Age, (3) 17 (1972). V104
205. H.K. Porter Co., USA, US Pat., 3.149.997; in Adhäsion, (2) 79 (1966).
206. Offset Technik, (8) 8 (1986).
207. Offset Praxis, (11) 98 (1986).
208. J. Merretig, Coating, (3) 50 (1974).
209. Kjin Co., Jap. Pat., 28.520/70; in Coating, (2) 38 (1972).
210. I.J. Davis (National Starch Chem. Co., Bridgewater, USA), US Pat., 4, 728,
572/01.03.1988.
211. D.J. Bebbington (Wiggins Tape Group Ltd.) EP251672/07.01.1988; in CAS
Silicones, 14, 4 (1988).
212. T. Tatsuno, K. Matsui, M. Takashi and M. Wakimoto (Kansai Paint Co.
Ltd.), Japan Pat., 22285927/11.12.1987; in CAS Adhesives, 14, 3 (1987).
213. I. Benedek, Pressure Sensitive Formulation, VSP, Utrecht, 2000, pp. 25–28.
214. Tappi J., (5) 25 (1988).
215. G. Schweizer, Das Papier, 10A, 1(1985).
216. Coating, (6) 184 (1969).
217. Coating, (11) 89 (1984).
218. T.J. Bonk (Minnesota Mining and Manuf. Co., St.Paul, MN, USA), EP 12,
070, 831/03.10.1984.
219. R. Schuman and B. Josephs (Dennison Manuf. Co., USA), PCT/US86/
02304/28.10.1986.
220. K. Göller, Adhäsion, (4) 101 (1974).
221. Coating, (4) 23 (1969).
222. J.C. Pasquali, EP 0122847/24.10.1984.
223. German Pat., 2,110,491, in Coating, (12) 363 (1973).
224. U.E. Krause (Hensel Textil Co.), EP 0,238,014 A2/23.09.1987.
225. Coating, (3) 42 (1969).
226. M. Hasegawa, US Pat., 4,460,634/17.07.1984; in Adhes. Age, (3) 22 (1987).
227. Eastman Kodak, US Pat., 359, 943; in Adhäsion, (5) 328 (1967).
228. Vitta Co., US Pat., 3,598,073, in Adhäsion, (5) 328 (1967).
229. Japan Pat., 2.736/75; in US Pat., 3.691.140/18.09.72.
230. E. Pagendarm, OS/DE 3.628.784 Al/25.08.1986.
231. 3M Kokkai, Tokyo Koho, Jap. Pat., 63.175.091, in CAS, 25 (7) (1988).
232. I. Benedek and L. Heymans, Pressure-Sensitive Adhesives Technology,
Marcel Dekker, New York, 1999, Chapter 6.
233. I. Benedek, Pressure Sensitive Formulation, VSP, Utrecht, 2000, Chapter 5.
234. P. Tkaczuk, Adhes. Age, (8) 19 (1988).
235. Minnesota Mining and Manuf. Co. St. Paul, MN, USA, EP 4,693,935/18.09
1987.
236. C. Donker, R. Luth and K. Van Rijn, 19th Munich Adhesive and Finishing
Seminar, Munich, Germany, 1994, p.87.
237. L. Jacob, ‘‘New Development of Tackifiers for SBS Block Copolymers,’’
19th Munich Adhesive and Finishing Seminar, Munich, Germany, 1994, p.87.
Adhesive Performance Characteristics 351

238. Novamelt Research GmbH, Technical Bulletin, 1999.


239. Poli-Film Kunststofftechnik GmbH, Wipperfürth, Germany, Technical
Bulletin, 1994.
240. Novacel GmbH, FRG, Technical Bulletin, 1994.
241. Adhes. Age, (12) 43 (1981).
242. Tappi J., (9) 104 (1983).
243. K.N. Nakao, J. Adh. Soc. Japan, (6) 6 (1970).
244. K. Fukuzawa and T. Uekita, ‘‘New Methods of Evaluation for Pressure-
Sensitive Adhesives,’’ Proc. of the 22nd Annual Meeting of the Adhesion
Society, Panama City Beach, Florida, Febr. 21, 1999, p.78.
245. J. Andres, Allg. Papier Rundschau, (16) 444 (1986).
246. US Pat., 4,563,389; in S. Krampe and C.L. Moore (Minnesota Mining and
Manuf. Co., St.Paul, MN, USA), EP 0,202,831 A2/26.11.1986.
247. A.W. Norman, Adhes. Age, (4) 36 (1974).
248. S.D. Tobing and A. Klein, ‘‘Synthesis and Structure Property Studies in
Acrylic Pressure-Sensitive Adhesives,’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr.25, 2001, Williamsburg, VA, p.131.
249. R.D. Adams, R. Thomas, F.J. Guild and L.F. Vaughan, ‘‘Thick Adherend
Shear Tests,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr., 25, 2001,
Williamsburg, VA, p.59.
250. R.A. Fletscher, ‘‘The Temperature Factor in Compounding Thermoplastic
Rubber-based HMA,’’ 11th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1986, p.36.
251. A. Sustici and B. Below, Adhes. Age, (11) 17 (1991).
252. A. Bell, Shell Bulletin, TB, RBX/73/8/6.
253. J.A. Schlademan, Coating, (1) 12 (1986).
254. L.M. Schrijver, Coating, (3) 70 (1981).
255. J.A. Schlademan and J.G. Bryson, Proc.of the 21st Annual Meeting of the
Adhesion Society, Savannah, GA, Febr., 1998.
256. A. Pocius and C. Dahlquist, Adhesion and Adhesives, ACS Audio Courses
(1986); in H.W. Yang, ‘‘Effect of Polymer Structural Parameters on PSAs,’’
Proc. of the 22nd Annual Meeting of the Adhesion Society, Panama City Beach,
FL, Febr. 21, 1999, p.63.
257. J.M. Gordon, G.B. Rouse, J.H. Gibbs and W.M. Risen Jr., J. Chem. Phys.,
(66) 4971 (1977).
258. G. Meinel, Papier u. Kunststoffverarb., (10) 26 (1985).
259. H.J. Boob and H.M. Ulrich, Kautschuk, Gummi, Kunststoffe, (3) 209 (1984).
260. L.A. Sobieski and T.J. Tangney, Adhes. Age, (12) 23 (1988).
261. M. Ulrich, US Pat., 24. 906; in Handbook of Pressure Sensitive
Technology (D. Satas, Ed.), Van Nostrand-Rheinhold Co., New York,
1982, p.203.
262. Coating, (2) 43 (1987).
263. W.J. Whitsitt, Tappi J., (12) 163 (1988).
264. Tappi J., (5) 182 (1988).
265. L. Krutzel, Adhes. Age, (9) 21 (1987).
352 Chapter 6

266. Coating, (7) 186 (1984).


267. Adhes. Age, (11) 27 (1983).
268. C.K. Otto, Allg. Papier Rundschau, (16) 438 (1986).
269. J.P. Keally and R.E. Zenk (Minnesota Mining and Manuf. Co., St. Paul, MN,
USA), Canad. Pat., 1, 224, 678/19.07.1982 (USP. 399, 350).
270. Adhäsion, (2) 43 (1977).
271. I. Benedek, Pressure Sensitive Formulation, VSP, Utrecht, 2000, Chapter 3.
272. R.P. Mudge (National Starch Chem. Co., Bridgewater, USA), US Pat., 4,
753, 846/28.06.1988.
273. Tappi J., (9) 105 (1984).
274. Allg.Papier Rundschau, (5) 228 (1987).
275. R.P. Mudge (National Starch Chem. Co., Bridgewater, USA), EP 0, 225, 541/
11.12.1985.
276. J.A. Fries, ‘‘New Developments in PSA,’’ Tappi HM Seminar, 2 June, 1990,
Toronto, Canada.
277. I. Benedek, Pressure Sensitive Formulation, VSP, Utrecht, 2000, p.55.
278. Paper, Film and Foil Converter, (3) 76 (1988).
279. Vinamul GmbH, Nacor 90, Data sheet, 1988.
280. D.G. Pierson and J.J. Wilcynski, Adhes. Age, (8) 53 (1990).
7
Converting Properties of PSAs

Convertability is the ability of the adhesive to be used for coating and


laminating, and the property of the laminate to be used for label (or tape)
manufacturing. The most important factors which contribute towards good
convertability of a self-adhesive laminate include the nature of the face
paper, the type of the adhesive and silicone, the release force, the properties
of the siliconized release liner, and the die-cutting properties of the adhesive
and of the face paper. There are two different aspects of the convertability of
an adhesive, namely the convertability of the fluid adhesive (solvent-based,
water-based, or hot-melt), and the convertability of the coated adhesive
(processability); good convertability implies the suitability of the adhesive
fluid to be coated without any problem, and the capability of the laminate
to be processed (cutting, die-cutting, printing, labeling, etc.) without any
problem. As discussed in detail in [1] and in Chap. 3, Section 1.4,
convertability in practice is the sum of the convertability of the adhesive and
that of the laminate.

1 CONVERTABILITY OF THE ADHESIVE

Because of the different materials used as face stock materials and the dif-
ferent technologies and adhesives used, the coatability of the liquid adhesive
is a function of many different parameters (e.g., the physical state of the
adhesive, the nature of the face stock, the coating technology, the laminate
structure, the end-use properties, the cost, and the nature of the release).

1.1 Convertability of Adhesive as a Function


of the Physical State
The convertability (coatability) of the adhesive is characterized by the
parameters such as the wet-out, the coating rheology, coating speed, and

353
354 Chapter 7

versatility. All of these parameters depend on the physical state of the


adhesive, whether it concerns a diluted liquid system, yielding an adhesive
layer through the evaporation of the carrier liquid (water or solvent) or a
100%-solid system, giving a solid adhesive layer by a physical (hot-melt) or
chemical (UV or electron beam) process. Hence the coatability of diluted
and undiluted systems needs to be discussed separately.

Coatability of Diluted Systems


Diluted adhesive systems are defined as solutions or dispersions of
adhesives. There are aqueous or solvent-based adhesive solutions, and
aqueous or solvent-based dispersions. The most important ones are organic
solvent solutions of adhesives (solvent-based adhesives) and water-based
dispersions of adhesives (water-based adhesives). In fact most of the solvent-
based adhesives also contain (at least partially) dispersions (like rubber-resin
adhesives), and most of the dispersion-based aqueous adhesives contain
some dissolved materials (e.g., thickeners, tackifying polymers, etc.). In
Chap. 2, the special features of the wetting out of solvent-based adhesives as
well as their coating rheology were discussed. For coating practice more
information is given by the versatility of the PSAs.

Coatability of Solvent-Based Adhesives


Factors influencing the coatability of solvent-based systems include the
influence of the solvent content on the coating thickeners, coating appea-
rance, and the mechanical-hydraulic forces in the coating nip [2]. Solvent-
based adhesives are mostly solutions of rubber-resin mixtures or acrylic
copolymers. Their physical properties depend on both components, namely
on polymer and solvent, and on the composition of the adhesive solution.
Solvent-based adhesives are generally highly viscous systems with low
surface tension solvents as the carrier; they are (at least partially) directly
coated onto high surface energy webs. The wet-out of solvent-based PSAs is
less difficult than with water-based PSAs. The coating rheology of solvent-
based PSAs concerns the rheology of a true solution, thus is mainly viscosity
driven and more simple than that of water-based systems. Coating versatility
is the capability of an adhesive to be coated on different machines, using
different coating technologies without changes in its composition or perform-
ance characteristics. Because of the high viscosity of the solutions, there are
only a limited number of coating devices able to use solvent-based PSAs.
Coating speed is the most important coatability-related property of
solvent-based PSAs. The key issues include:
different coating devices for different viscosities;
base viscosity as a function of the polymerization recipe;
Converting Properties of PSAs 355

thickening/diluting response of the adhesive;


different technologies (transfer/direct) for different adhesives;
different technologies for different end-uses.
In general the most important parameters of the liquid solvent-based
adhesive (solids content and viscosity) can be easily and precisely adjusted
because the (mechanical) stability of the system is independent of viscosity
(or solids content). On the other hand, there is no need for an exact
adjustment of these properties because of the higher viscosity. Thus solvent-
based PSAs are suitable to be used on different coating devices. The
(possible) different nature of the used solvents only acts as a limiting factor.
Therefore,

Versatility : Solvent-based PSAs > Water-based PSAs ð7:1Þ

The coating speed remains the most important converting property of


solvent-based adhesives. In contrast to water-based PSAs, there is no limit
(minimum speed) for the coating speed because of the unlimited wet-out of
these adhesives. The limits of the running speed for a given machine are
dictated only by the explosion limits of the solvents. Theoretically, a high
solids content and low boiling point solvent allow high drying speeds; thus
an increase in solids content and the appropriate choice of adequate solvents
ensure better running speeds. The newest experimental products generally
contain a higher solids content. Theoretically, a 100%-solids content (at very
low molecular weight) is possible; however, the most concentrated aqueous
dispersions have 67–69% solids content (theoretically 75% is possible).
Thus ways to achieve higher coating speeds appear to be quite different for
solvent-based or water-based PSAs.

Coatability of Water-Based Systems


The coatability of water-based adhesives also is a function of the wet-out,
coating versatility, coating rheology, and coating speed. As stated by [3] the
driving force for dynamic wetting is in fact the surface tension of the
adhesive rather than the work of adhesion. Thus, the wet-out of water-based
systems constitutes a difficult problem because of the high surface tension of
the carrier (water), the low viscosity of the most important ready-to-use
formulations, the low density, and the transfer technology used for coating.
The use of co-solvents is limited, while higher viscosities need different
coating devices. In contrast to solvent-based PSAs, an improvement of the
wet-out characteristics (with surface active agents) induces irreversible
changes to the adhesive properties. Thus the coatability of water-based
formulations is more sensitive than that of solvent-based PSAs.
356 Chapter 7

For water-based PSAs, the coating rheology concerns a dispersion, not


a solution, and is thus time dependent. No changes in carrier material are
allowed and the dispersed polymer is of high molecular weight; coalescence
yields a composite, anisotropic layer. Coating devices are very sensitive
towards viscosity and thus their coating versatility is reduced; the coating
speed depends on solids content as well as on dispersion characteristics.
A minimum level of surface tension is required for dynamic wetting, the
maximum speed for a machine is dictated by rest water content (humidity)
of the dried layer. Rest humidity content is an equilibrium value which
depends on the wet-out additives; thus wetting out is the most important
convertability parameter. The coating speed is more dependent on the
coating weight for water-based PSAs than for solvent-based PSAs. In
addition, the formulating freedom with coating speed remains very limited,
unlike with solvent-based PSAs. The special features of the wetting out for
aqueous dispersions were discussed in Chap. 2 concerning the rheology of
water-based dispersions. Summarizing, the most important key issues
concerning the wet-out are:
wet-out as the combination of wetting/dewetting;
dynamic versus static wetting;
factors acting against dewetting (shear, coalescence);
parameters influencing wetting/dewetting (surface tension, contact
angle, viscosity, density);
surface tension and parameters influencing it;
contact angle, and the problems of measuring this parameter;
the influence of the viscosity (versatility, wet-out, coating weight).
The theoretical aspects of the coating rheology of water-based adhesives
were discussed in Chap. 2.
There is a qualitative difference between the coating rheology of the
liquid and that of the semisolid (dried) adhesive layer. The former is like a
solvent-based (solute) rheology, whereas the latter is more that of a
reinforced matrix. For the first step of the coating, temporary and reversible
shear sensitivity of the water-based PSAs determines their behavior on and
after the metering roll. In the second stage, the diffusivity of the coated layer
and its flow determine the drying speed and the aspect of the dried,
coalesced adhesive film. Coating rheology of solvent-based adhesives is a
function of polymer and solvent quality, whereas coating rheology of water-
based adhesives is a function of the dispersion structure. On the other hand,
the influence of the metering device, and of the drying ‘‘regime’’ is more
pronounced than for solvent-based PSAs. Slight shear rate changes may
avoid ‘‘stripe’’ structures, air and heating adjustment may avoid adhesive
blow-away, viscosity control may help adjust the smoothness of the final
Converting Properties of PSAs 357

coating. Unlike solvent-based adhesives, formulating only has a limited


influence on the coating rheology of the adhesive. Key parameters include:
adhesive build-up on the machine, mechanical stability, due to
insufficient flow;
coagulum build-up on the machine and during formulation
(mechanical stability);
coating uniformity (striped, textured structures, insufficient flow);
stability of the viscosity (thixotropy and decrease of the viscosity,
either reversible or irreversible);
foam formation.
Coating versatility refers to the suitability of the water-based PSAs to be
coated by different coating technologies. Today in Europe two coating
methods for emulsion PSAs dominate their respective fields [4], namely:
metering bar systems for tapes and protective films, and reverse gravure for
labels. The coating technology will be discussed in more detail in Chap. 9. The
mechanical stability of the water-based adhesive on the coating machine is
characterized by the following parameters: coagulum build-up on the ma-
chine, grit build-up in the dispersion, stability of the viscosity, and foaming.
Several latex adhesives have a tendency for coagulum build-up on the
metering roll of the reverse roll coater. Such coagulum may arise either from
poor mechanical stability under shear loading or from imperfect doctoring
leading to a thin coating on the roll, which dries and cannot be redispersed
so that these particles eventually transfer to the applicator roll [5].
Dispersions with a high solids content and a high surface tension tend to
build up dry, tacky residues. Frequently changing the rheology (diluting and
thickening) leads to deposit-free water-based adhesives. Often the adhesive
layer build-up on the machine is associated with insufficient fluidity of the
‘‘almost dried’’ adhesive, giving a coating appearance that is textured with
stripes. Diluting/thickening also helps to avoid this phenomenon.
Unlike adhesive build-up, grit (coagulum, discrete deposits) build-up
is associated with insufficient mechanical stability of the dispersion (i.e., an
insufficient stabilizing agent level in the formulation). According to [6]
tackified acrylic emulsions are more susceptible to high shear induced
coagulation. A higher acrylic acid level in the copolymer reduces the coagu-
lum. Generally, tackified formulations are more sensitive (the IR spectrum
of the coagulum showed that it is 3 times higher in tackifier than is the
formulation); rosin-ester tackified PSAs are more sensitive than acid-resin
tackified formulations (e.g., an acid-rosin with 60 ppm standard coagulum
content does not influence negatively current paper coating formulations;
rosin-ester dispersions must possess less than 20 ppm coagulum).
358 Chapter 7

Theoretically, high shear stirring tests show the grit build-up and the
lack of mechanical stability; generally, most PSA formulations withstand
this test. Real data can be obtained with a two-cylinder laboratory device, or
practically, a 3-day running test on the production line.
Most unthickened PSA dispersions exhibit only slight thixotropy (i.e.,
changes in the viscosity after formulation that are generally irreversible). It
is to be noted that 0.5–1.5 s (Ford Cup sec) are normal absolute values for
observed changes in the viscosity during 8–24-hr production runs.
Foaming does not really constitute a problem as most machines do not
attain critical speeds. Most dispersions possess viscosities below 500 or
above 5000 mPa s. Also there are a variety of very efficient defoamers.
Water-based PSAs require longer drying times (and hence longer
dryers) than solvent-based adhesives, and 15 to 150 m dryers are currently in
use. The coating speed of the adhesive not only has economical importance,
but ease of adjustment also influences the tolerances of the coating weight.
For gravure roll coating a short contact time between web and pressure roll
means a low coating weight, whereas a long contact time (low speed) leads
to a high coating weight. With a given gravure size it is possible to change
the coating weight within a range of approximately 10% [8].
Aqueous systems require more drying heat. The heat of evaporation of
water amounts to 540 cal/g, while that of typical coating solvents, such as
methyl ethyl ketone (MEK) and toluene, is 100–200 cal/g. On the other hand
water-based systems are dried more efficiently with large volumes of air
rather than with air of higher temperature [7]. This difference translates into
longer ovens, lower line speeds, and higher utility costs [8] for water-based
systems. Therefore the running speed has its economical importance,
possibly leading to lower costs and higher productivity (Fig. 7.1).

Figure 7.1 Parameters of the PSA laminate influenced by the coating speed.
Converting Properties of PSAs 359

Originally several technical parameters were evaluated in order to


improve the drying speed of the diluted systems, including an increase of the
solids content and the transfer coating technology. With transfer coating a
high coating speed of the non-water-sensitive web was possible; an increase
of the solids content allows (at least theoretically) a reduction of the volatiles
and a higher drying speed. The speed limiting steps are the foaming, the
appearance of the coated layer, and the residual water content of the dried
adhesive. It is evident that for a water-based PSA exhibiting an adequate
rheology the upper limit of the running speed is:

Running speed ¼ Maximum drying speed for normal dispersions


ð7:2Þ

For high solid dispersions:

Running speed ¼ Maximum speed leading to a smooth coated layer


(7:3Þ

For medium viscosity dispersions:

Running speed ¼ Maximum speed which limits foam formation


ð7:4Þ

The physical state of the PSA influences its drying properties [9].
Solvent-based and water-based PSAs display a quite different evaporation
behavior (i.e., they require a quite different temperature gradient during
drying) (Fig. 7.2). The coating and drying speed will depend on the proper
adjustment of the drying (oven) temperature. Key aspects include the
economical/technical importance of the running speed, the built-in
characteristics of the water-based PSAs influencing the running speed,
and the technical limits of the running speed.

Coatability of Hot-Melt PSAs


According to their physical state the coatability of hot-melt PSAs is limited.
Their viscosity depends on the temperature and on their chemical
composition/aging. Because of the restricted range of raw materials
available for their formulation, there is a narrow temperature interval
suitable for them to be coated. On the one hand, this interval is situated at
relatively high viscosities. Thus special metering devices only (see Section 2.2
of Chap. 9) may be used; on the other hand, the high viscosity in the
360 Chapter 7

Figure 7.2 Temperature gradient required for drying of: 1) solvent-based PSA
coating; 2) water-based PSA coating.

molten state allows a good wet-out. Contactless coating methods for


HMPSAs have been developed to allow the use of hot spray technology in
the nonwovens market.
Coating Versatility. The coating versatility of hot-melt PSAs remains
limited. Generally, special agents (e.g., waxes, oils) are used in the
formulation (see Section 2.9 of Chap. 8) in order to extend the coatability
range of the adhesive.
Coating Rheology. Hot-melt PSAs based on Kraton G are coated at
a viscosity of 25,000–80,000 mPa s (at 177 C) [10]. Acrylic hot-melt PSAs
possess a viscosity range of 8000–30,000 mPa s [11]. As illustrated by these
examples the viscosity of current hot-melt PSAs is of an order of magnitude
higher than that of solvent-based PSAs, and even higher than water-based
PSAs. It is not possible to change this with dilution agents or shear forces
(machine).
Coating Speed. Because of the lack of carrier or solvent, hot-melt
PSAs are ideal for the application of thick films onto webs at speeds of
hundreds of ft/min [7] versus acrylic PSAs coated at 170 m/min. Other data
Converting Properties of PSAs 361

show that in comparison with water-based PSAs (200–250 m/min) hot-melt


PSAs can be coated at higher speeds (300 m/min) [12] (see Chap. 9).
Warm-Melts. The development of new, low cohesion acrylic warm-
melts requires the use of a postcrosslinking agent [13–15]. The large scale
application of UV postcuring is enhanced by this raw material development.
Before crosslinking, the adhesive to be coated is a highly viscous fluid (at
room temperature) and can be processed at 120–140 C (its viscosity is about
10–20 Pa s). A special feature of this radiation crosslinking is its anisotropy.
The radiation is partially absorbed, partially transmitted, and partially
reflected. The maximum radiation is given at the top of the adhesive layer.
Therefore direct and transfer coating give different adhesive characteristics.
Such UV-curable polymers are suitable for different end-uses (e.g., deep
freeze or removable adhesive) as a function of the UV dose [16]. The
crosslinking level is high enough to cut and stamp without edge bleeding.

1.2 Convertability of Adhesive as a Function of


Adhesive Properties
The adhesive properties influence the coatability of PSAs. A certain level of
initial flow (tack and peel adhesion) is needed to achieve the anchorage of
the adhesive onto the face stock or release liner. In some cases a primer
coating is required. A pronounced influence of the tack and shear may be
observed during the converting of the laminate.

1.3 Convertability of Adhesive as a Function of the


Solid State Components of the Laminate
The coatability of the adhesive depends on the face stock and release liner.
Wetting out, drying parameters, and running speed depend on the surface
characteristics of the face stock and release liner. Machine parameters also are
a function of the bulk properties (mechanical characteristics) of the web and
influence the coatability of the PSAs. Other bulk properties such as porosity,
chemical affinity, aging resistance, etc., also influence the coatability.

1.4 Convertability of Adhesive as a Function of


Coating Technology
The convertability of the adhesive depends on the coating technology. For
all types of PSAs and independent of their physical state, it is theoretically
possible to use direct or transfer coating. Practically, the selected coating
technology (direct or transfer) supposes quite different rheological behavior.
As an example, without thickening, a low viscosity conventional latex will
362 Chapter 7

not completely wet a silicone-coated release liner. In this case the applica-
tion, convertability of the adhesive depends on the coating technology [5].

1.5 Convertability of Adhesive as a Function


of End-Use Properties
Transfer coating is preferred for heat-sensitive or porous face stocks;
transfer-coated PSAs have to be formulated for better wet-out. Hence the
convertability of the adhesive is a function of its end-use properties. On the
other hand, a better anchorage (see Section 2.2 of Chap. 9) is obtained
through direct coating. Certain adherend require a special adhesive
smoothness or texture, which is a function of the metering device. As
described in detail in [17] there are some special pressure-sensitive products
containing an adhesive imbedded in the carrier material. Carrierless tapes
were developed also. In such cases special manufacturing technologies and
converting characteristics are required.

2 CONVERTING PROPERTIES OF THE LAMINATE

In the present context laminating means the production of a layered material


from two or more webs using a bonding medium. In the case of PSA
laminates, the face stock and release liner are connected together
temporarily by means of a PSA. The PSA-coated material will be processed
(converted) and used in the form of the laminate. Generally, the face stock
will be printed either in-line or off-line, but after coating with a PSA; the
liner also may be printed. The coating properties of the solid state
components of the laminate (with adhesive, release, or inks) are discussed in
detail in our book concerning the manufacture of pressure-sensitive
products [1]. The product class (e.g., label, tape, protective film etc.) related
special features are described too. A coated, laminated material usually
starts as a continuous roll which is then cut in discrete pieces. Cut material
(printed before or after cutting) will be used for labeling purposes. Labeling
(separation of the liner from the adhesive and affixing the adhesive-coated
label on the final substrate) can be done manually or mechanically, this
process being influenced by the face material-adhesive-release properties.
Hence the convertability of the laminate will be influenced by the
printability, confectionability (ability to be slit, cut, die-cut, perforated,
embossed etc.), and labeling properties. As discussed in detail in [1] the
winding performances affect such characteristics also. All of these param-
eters depend on the solid state components of the pressure-sensitive
laminate and on the adhesive. Therefore a detailed examination of these
Converting Properties of PSAs 363

parameters requires a short description of the pressure-sensitive laminate, its


components, and its construction.

2.1 Definition and Construction of the


Pressure-Sensitive Laminate
Generally, pressure-sensitive laminates are sandwich structures where two
(different or identical) solid state components (generally flexible sheets) are
temporarily bonded together with the aid of a PSA. The pressure-sensitive
composite is a temporary structure where one of the solid state components is
acting as a temporary shield only, in order to avoid the deterioration of the
PSA layer. This part, known as the release liner, is in some cases (mainly for
tapes) identical with the face stock material. Such a construction is possible
and necessary for pressure-sensitive tapes used in continuous reels. Pressure-
sensitive labels are discontinuous materials where the release layer has to be
built in as a separate solid state component. Therefore, it may be assumed
that pressure-sensitive labels contain the following main components: a face
stock material, a pressure-sensitive adhesive, and a release liner. All these
components are relatively thin; paper and soft or hard plastic films are used as
solid components. Soft plastic films can be bonded only by using PSAs where
the use of rigid adhesives is not possible because of stress concentration.

Main Laminates
The components of the PSA laminate and its build-up depend on the
end-use, on the application technology of the labels, and on the PSA-coated
laminates.

Sheet Material. Pressure-sensitive labels are temporarily built-up


composites. The end-use (application) of the label assumes peeling off the
label from the release liner and the subsequent bonding of the free adhesive
surface onto a substrate. This end-use called labeling is carried out by hand
or automatically, with the aid of labeling guns. The main criterion concern-
ing the feasibility of hand applied or automated labeling is the dimension of
the label. It is evident that large sized labels have to be considered as sheet-
like materials. On the other hand, economical considerations also influence
the labeling. Unicates or small series of labels have to be manufactured by
technologies using sheet-like materials. Big series of relatively small and low
cost labels have to be applied automatically with labeling guns. Printing,
cutting, and debonding conditions for sheet and reel materials are different
and therefore the quality requirements for both these label categories also
differ. Sheet materials were used with priority some years ago; films as face
364 Chapter 7

stock material were tested for sheet-like labels first. Because of the relatively
big surface and the low delaminating and labeling speed of these labels, the
face stock quality and the cohesion of the adhesive play an important role
for sheet-like pressure-sensitive labels.
Roll Material. Labels applied from a reel have to be tacky in order
to ensure a suitable adhesion onto the adherend after a very short dwell
time; at the same time they have to be very slightly bonded to the release
liner in order to display a low debonding (delaminating) resistance.
Laminate Build-Up. Theoretically, a PSA laminate consists of at least
two solid state components (face stock and release liner) and a liquid phase
(PSA) present as a continuous layer. The face stock or release liner itself
may have a composite structure, the components of the laminate may show
a discontinuous nature, and the laminate may possess a multilayer structure.
Therefore the construction of the laminate exerts a strong influence on its
properties. A detailed description of the pressure-sensitive product classes
and their build-up is given in [18].

Laminate Components
The pressure-sensitive layer (a liquid adhesive layer) is enclosed between two
solid state sheet (forming) materials. Generally, there is only one liquid state
component in the laminate, namely the pressure-sensitive adhesive layer
between the face stock material and release liner. In some cases an additional
primer coating is applied between the face stock material and the PSAs.
Because of its reduced and limited flow properties, the primer coating may be
considered (like the release coating) as a solid state component of the
laminate.
The main solid state components of a pressure-sensitive label are the
face stock material and the release liner. There are also sophisticated
sandwich structures which possess a lot of face stock and release layers,
where in some cases one or more face stock materials should also act as the
release liner. Multiweb constructions can include solid-state components and
adhesives that differ in their chemical nature and build-up. There are
multilayer self-adhesive labels which comprise a carrier layer containing a
silicone layer, an adhesive layer, a printed message, a carrier layer (e.g.,
polyester), a release layer (e.g., silicone), a carrier layer (e.g., polyethylene),
and a printed message. These labels are useful when a printed message is to be
attached to two different objects [19]. Forms are labels that have a multiweb,
multilaminate structure, where continuous and discontinuous carrier
materials and PSAs with different adhesive properties are laminated together,
to allow stress and time dependent controlled delamination [18,20].
Converting Properties of PSAs 365

Face Stock Materials. The face stock material is the solid state
component of the pressure-sensitive label with a permanent character (i.e.,
the adhesion of the PSAs on the face stock material—the anchorage—
should be higher than its bonding to the release liner). Generally, the
adhesion of the PSAs to the face stock material (i.e., anchorage) should be
superior than its bonding to the substrate. The most important face stock
materials are paper, cloth, plastic film, metallized film, metal foils,
elastomeric foams, and nonwoven materials. Face materials include
specially modified papers, as well as plastic films such as PVC, cellulose
acetate, polyolefins, polyester, and polystyrene, and aluminum foil as well as
metallized papers and films. Market requirements impose certain character-
istics of the label material [21]. Labels for oil containers need to display
durability, visual impact, and flexibility. Toiletries and cosmetics require a
‘‘no label’’ look, a high consumer image, and stain resistance; the electrical
market needs include visibility, durability (e.g., vinyls, PET), and heat
resistance. Toys need nontoxic labels with long life and visual impact. The
chemical industry requires resistance to chemicals, visibility, and durability.
Thus quality requirements for the label indicate the need for adequate
face stock materials. Laminate manufacturers must make an adequate
choice of the face stock material to be coated. In most cases this choice is
limited, either imposed by the customers or, in some cases, determined by
the manufacturers’ own capabilities (i.e., coating machine or adhesive
system).
The most important decision is the choice between a paper or film-
based face stock material and in each of these domains between permanent
or removable labels. Different mechanical performance characteristics are
needed for tear-debonded (peeled) permanent and easily removable
products. Tamper evident labels require carrier materials having low
mechanical stability. A different material stiffness is required for squeezable
or ‘‘rigid’’ labels or for conformable face materials. For special uses (e.g.,
medical products) preferred face stocks are those which permit transpiration
of moisture; for other applications (e.g., pharmaceutical packaging)
migration-resistant barrier layers are needed. On the other hand, storage
or weather conditions may point towards quite different materials.
Processing of the laminate (as sheet or roll material) and its printing
determine the machining ability and surface properties. Various printing
methods or writeability require quite different film carrier materials.
Economical considerations may interfere with technical requirements. The
following parameters influence the choice of face stock materials:
conformability, grain direction, appearance, printing method, mechanical
properties, chemical and environmental resistance, and primer or barrier
coating [22,23].
366 Chapter 7

Edge lifting of labels can occur if a label stock is too stiff to conform to a
curved surface. For foil laminates and whenever possible, the layout should
be in the long-grain direction to provide maximum conformability. Many
plastic containers are labeled with 60-lb high-gloss paper or foil-laminated
papers selected for their attractive appearance. Special inks and printing
techniques may be used because of the label end-use. Cosmetic bottles, for
example, are commonly labeled with a PVC-coated foil laminate which resists
alcohol [22]. The required tear and tensile strength of the face material of the
label depend upon the label converting method and the method of applying
the finished labels. Additional factors to be considered include UV light
resistance and stability, cold temperature application and resistance, water
immersion or high humidity resistance, and resistance to various chemicals.
A primer coating which improves the anchorage of the adhesive to the label
stock may also provide a barrier to prevent components in the adhesive from
penetrating the label stock. Figure 7.3 summarizes the main criteria for the
choice of the face stock material for PSA laminates.
Through the choice of the face stock materials new trends may be
observed [24]. Some years ago, price and productivity factors were
predominant. Today environmental, esthetical, chemical resistance, flex-
ibility, and recycling considerations appear more important. The most
important face stock materials used are fibrous materials and films.
Fibrous materials are mostly paper-like products (paper and synthetic
paper), textiles, and nonwoven materials. Films may be made of plastic,
metal, or a combination thereof (metallized plastic) [25,26]. The most

Figure 7.3 The main criteria for the choice of the face stock material.
Converting Properties of PSAs 367

commonly employed film face stock materials in PSAs are polyester,


polypropylene, polyethylene, polyvinyl chloride and, of course, reinforced
paper [27]. Paper was the traditional face material for PSA labels for most of
the almost fifty years of the existence of the industry. This is changing, the
estimate for nonpaper materials running as high as 40% by 1990 [28], and
actually covering more than 50%. There are several reasons for this trend;
pulp prices have risen dramatically while petroleum prices have declined.
Changing packaging concepts such as the squeezable plastic bottles
dictate the use of flexible face materials. Eye-catching labels compete more
effectively. Durability of the label and retention of its appearance also are
factors, with the ‘‘last look’’ prior to discarding the package forming a
mental image which can influence future buying decisions [29]. Previously,
the only opportunity for film-based primary labels was in situations where
the package might be exposed to harsh environments or where superior
graphics were necessary [30]. Actually the packaging industry was looking
for striking graphic effects and chose relatively expensive plastic face stocks
because of their smoother, more uniform printing surface [31].

Paper Used as Face Stock. Paper is the main face stock and release
liner material actually used. Its good wettability, dimensional stability
during coating, printability, and cuttability as well as its well-known
technical history as an information carrier made it possible to use it from the
beginning of the label industry; an ideal face stock paper possesses a high
gloss, a high surface strength, a good wet-out, and good anchorage for
printing inks [26]. A high surface strength (tear) may be achieved with a high
binder level. Gloss and wetting out are influenced by the pigments used. The
most important mechanical properties of paper type as a function of its
weight are listed in Table 7.1.
Noncoated papers, clay-coated papers, colored papers, paper-film
laminates, and special papers also are used as face stock material [28]. The
most important factors for base papers used for pressure-sensitive face
materials are the required quality level, quality consistency, and efficiency
[31]. Special paper grades can be selected which exhibit barrier properties
(e.g., to prevent the paper surface being marred by bleedthrough of adhesive
components). The pressure-sensitive label papers must be suitable for reel-
to-reel conversion, reel-to-reel printing, electronic data processing (EDP)
label printing, and for sheet printing. Demands on label paper from the
converter’s point of view are printability, die-cutting, lay-flat, and matrix
stripping properties. Suitable pressure-sensitive label papers must be
printable in all commercial printing systems: letterpress, flexo, rotogravure,
offset, silk screen printing. Nonimpact printing methods were developed
also and are used specially for narrow web printing. Important growth in
368 Chapter 7

Table 7.1 Quality Criteria for Face Stock Paper

Properties

Tensile Elongation Bekk smoothness


strength (N/cm) at break (%) (sec) Bursting
pressure
2

Weight Thickness MD CD MD CD Top Bottom (N/cm )

50 0.050 40 14 4 6 2000 200 15


60 0.050 40 25 4 6 1500 400 15
70 0.075 40 18 4 6 300 30 18
80 0.080 60 25 4 6 1500 500 25
80 0.085 45 20 4 6 200 200 18
110 0.150 100 50 20 20 — — —
125 0.180 100 50 20 20 — — —
150 0.150 80 60 7 18 — — —
330 0.430 250 100 4 6 50 10 —

the label industry has come from blank labels or nearly blank labels, for
which more printing is done at the point of use. Therefore combined
printing methods have been among the main processes used.
Years ago, either simple, uncoated grades or cast-coated papers were
used for pressure-sensitive labels. The properties of different paper qualities
have been studied [32]. The sensitivity of paper towards humidity should
also be taken into account (see Chap. 3) since paper is made up of both
crystalline and amorphous cellulose, with the latter undergoing softening as
the moisture content increases. The elastic modulus of dry paper is usually
considered to be temperature dependent. The decrease in relative modulus
with increasing moisture content is different for papers of different
crystallinity [33] (see Chap. 3, Section 2.1 too). There are very different
paper qualities used as face stock, and there are many quality criteria
recommended for selecting adequate face stock materials. In a first step, the
paper weight is used as a simple quality index. Almost without exception,
standard labels are printed in color on 85–100-g paper. A high stiffness can
be obtained by using a high weight or a more voluminous paper grade. The
strong influence of the calendering should also be noted. Papers used as face
stock material have a weight ranging from 30 to 250 g/m2 and weigh
generally about 80 g/m2 [25,28].
Paper for data systems is primarily uncoated, engineered for strength,
and has a high absorbency to capture ink from a computer terminal
[34]. Coated papers constitute the larger market segment, ranging from
cast-coated sheets to enamel and matte finish grades. These include thermal
Converting Properties of PSAs 369

sensitive and other specially treated coated products. Specialty structures are
another major market that includes fluorescent-coated papers, latex
impregnated papers, polyethylene and paper laminations, and metallized
laminates [35]. Uncoated and coated papers are used in impact printing:
flexography, gravure, dot matrix, ultraviolet light-cured letterpress, and
offset. Both types of papers also are used in nonimpact printing, laser,
inkjet, ion deposition, magnetography, and direct thermal and thermal
transfer printing. In thermal transfer two contradictory properties are
required: a high degree of smoothness and a surface that at the same time
will absorb ink. For special uses there are different kinds of paper; humidity
resistance, fat resistance, and a flameproof character are characteristics of
these papers. Laminated papers are manufactured in order to achieve chemi-
cal resistance. Metallized papers are used to get a better lay-flat; aluminum/
paper laminates (7 m aluminum, 50-g paper) also are used. For dispersion
coating, paper heavier than 80 g/m2 should be used in order to ensure
enough dimensional stability [35]. On the other hand, the porosity of paper
is more critical when direct coating. For labels for high speed printing
applications, the absorbency and porosity of the paper should be improved
[36]. The technical requirements for thermal papers were described by
Park [37]. Paper types used include calendered Kraft for paper roll labels
with the following advantages: good surface hold-out, smoothness,
dimensional stability, and an average weight of 65 g/m2 [38]. The effect
of super-calendering on paper properties is better gloss, smoothness, and
ink hold-out; super-calendering increases the stiffness of the paper [39].
Clay-coated paper is used for sheet labels with the following advantages:
smoothness, dimensional stability, lay-flat during printing and converting,
with an average weight of 85 g/m2 [28]. The stiffness of machine-finished,
machine-glossed papers is higher than that of colored clay-coated and
calendered paper [40]. Polymer-coated papers for transfer coating of
aqueous adhesives [38] are suitable for very low coating weights, possess a
waterproof layer under the silicone, and exhibit good solvent resistance, as
well as good dimensional stability with a double-sided coating. Clay-coated
papers make up more than 40% of the market for paper face stock materials
[41]. The overlayer of coated papers may contain chemically very different
components such as kaolin, casein, acrylics, polystyrene-butadiene latex,
clay, starch, etc. [42]. The general requirements for paper as carrier material
and the special requirements for paper used as carrier material for different
pressure-sensitive products are discussed in detail in [43].

Films Used as Face Stock. In addition to the traditional base


materials within the range of high quality papers, films made of
thermoplastic materials are becoming more and more important. The
370 Chapter 7

volume of PSA labels made of paper face stocks remains significantly larger
than the volume of labels made with film face stocks. This is primarily due to
the relative cost differences between the face stocks. Hence the benefits of
film substrates for PSA labels are centered on appearance and performance,
and the dominance of paper in the PSA label market is being challenged;
this trend is more pronounced in Japan and Europe [44]. The advantages
of face stock materials other than paper include weather resistance,
durability, resistance against packaged fluids, squeezability, esthetics, and
transparency [45].
The following materials are used as film face stock: polyamide,
polyamide/aluminum, polyvinyl chloride, oriented polypropylene, poly-
ethylene, polyethylene terephthalate, and metallized plastics [45,46]. In
Europe, in the 1980s, 12% of the labels were film labels, on the bases of
PVC, PET, polystyrene, polypropylene, and extended polyethylene [21].
Other films used as face stock materials include cellulose acetate, cellulose
triacetate, cellulose acetobutyrate, and cellulose hydrate [47]. According to
Thorne, at the end of the 1980s, nonpaper products would possess a market
share of 12% by volume and 25% by value in Western Europe [48]. Actually
more than 20% of the face stock materials (in volume) are film based. It was
estimated that the cost of using synthetics is 4–5 times that of paper.
Therefore the first synthetics were only used when the properties of strength,
fine print acceptance, and moisture resistance were essential. In the last
decade the premises of the use of films were substantially changed. Less
expensive films were developed, and the paper price increased. Coatability
and printability of films were strongly improved. Films achieved a
well-defined standard quality together with other plastics-based web-like
materials, e.g., foams, nonwovens, etc. Thus for PSA labels with synthetic
face materials, the nature, caliper, gloss, opacity, and corona treatment are
detailed in the specification. Generally, films must display good coatability
and convertability in a laminate, using the same manufacturing conditions
as paper-based laminates. The wetting out of the film materials is more
difficult because of the smoothness, the lack of porosity, the low surface
tension, and additive migration from the film. Coatability is influenced by
the high deformability (elongation) and low temperature resistance of most
films; convertability of the laminate is influenced by the plasticity/elasticity
balance of the material.
The films most commonly used as face stock materials are PVC,
polypropylene, and PET [49]. Special polyethylene face stock materials were
developed also. Such carrier materials are top-coated or pretreated to allow
smooth transport through the printer without jamming and damaging
the fusing drum in the case of laser printers. Packaging tapes are dominated
by oriented polypropylene (OPP) and PVC [50]; 150 m PVC is also used
Converting Properties of PSAs 371

for labels [51]. Calendered PVC and biaxially oriented polyester face stocks
have dominated film applications in primary labels over the last ten years.
Table 7.2 lists the plastic materials used (alone or as laminate) as face stock
materials for PSAs. Proprietary synthetic papers such as DuPont’s Tyvek,
Arjo Wiggins’ Synteape, and Polyart from Arjobex also find wide usage.
Generally, both soft PVC (20–50% plasticizer) and hard PVC are used
as face stock materials. For tapes 25–40 m PVC, polypropylene, and PET
were used [49, 52]. PVC (polyvinyl chloride), or ‘‘vinyl’’ was the first major
plastic to be used as face stock material. It was first used in the 1950s and
remains a significant factor in the market place [44]. Its primary
disadvantage results from the need to add plasticizers to achieve flexibility
and stabilizers to prevent degradation. Plasticizers and stabilizers can migrate
and affect the adhesive performance, so careful selection and testing are
required. The problem of changes in peel value by plasticizer migration is a
common phenomenon for removable PSA-coated soft PVC. The properties
of hard PVC depend on the manufacturing procedure. Suspension-made

Table 7.2 Plastics used as Face Stock Materials

Primed Commercial Products

Supplier
Material Yes No Trade name converter Alternative to Ref.

Cellulose acetate — — — — PVC [44]


BOPP X Rayoface UCB PVC —
OPP X Kimdura Kimberley-Clark PVC [44]
OPP x Oppalyte Mobil — [44]
PE X Compucal Flexcon — [44]
PE X Fle Flexcon — [44]
PET X — Decora — [44]
PET X Polylaser Raflatac Paper [53]
Polyolefin x Prymalin 3M — [44]
Polyolefin X Opticlear Mactac — [44]
PE/PS X Polyart BXL — —
Polyolefin x Maclitho Mactac Paper [44]
Polyolefin X IMAGin Mactac PVC [54]
Polyolefin X Megarex x-film PVC [55]
Polyolefin x Teslin Impact [53]
PS x Opticite Dow Paper [44]
PVC x — — — [44]
PVC x Flexmark- Flexcon — [56]
W-400-FW
372 Chapter 7

PVC has superior electrical properties to emulsion-based PVC. Hard PVC


possesses a good resistance against chemical oils and greases with good
mechanical properties and inflammability resistance [43]. The drying
temperature of soft PVC should not exceed a maximum of 45–50 C, for
hard PVC 75 C. Printability depends on the nature of plasticizers used in
the PVC recipe [57]. Generally, monomeric plasticized PVC is used as face
stock for labels [58].
Lay-flat for paper requires the control of the humidity with a precision
of 0.3–0.6%. Lay-flat of printed PVC film (coated with PSAs) depends on
the thickness of the film, the thickness of the ink layer, the bonding forces
between label and release, and the elasticity of the film and ink [59].
Film face stock materials may be coated directly or by transfer.
Although new, high productivity machines use the transfer process, there
remain some older special machines using the direct process. The surface
tension of the web and the surface tension of the water-based PSAs influence
the wet-out. In the most extreme case, when the surface tension of the
dispersion differs greatly from the critical surface tension of the film, fish
eyes and cratering will result (when the film surface tension is less than the
dispersion’s surface tension). Fortunately PVC displays an increased surface
tension (40 mN/m) [60] as compared to the surface tension of the common
water-based acrylic PSAs (28–38 mN/m). Soft and hard (plasticized) PVC
exhibit different levels of critical surface tension. The critical surface tension
of rigid PVC is 39 mN/m whereas flexible soft PVC has a critical surface
tension with values of 33–38 mN/m [61]. The most important disadvantages
of soft PVC films are the shrinkage and the flagging of the edges. Both
phenomena are most apparent for extruded PVC film. Shrinkage concerns
the change of the original dimensions during storage at room or higher
temperatures due to chemical or physical (environmental) influences.
Shrinkage is caused by the residual stresses introduced in the material
(i.e., by relaxation). In the manufacture of a PSA laminate and during PSA
coating or laminating, the PVC film is stretched and strained; the strained
film will be laminated with the dimensionally more stable paper-based
release liners. During storage relaxation occurs and the strained material
returns to its original length (i.e., it shrinks) (Fig. 7.4).
The shrinkage of soft PVC does not depend on the tension of the film.
It is caused by the laminate manufacture. Another part of the shrinkage is
induced during the manufacture of the film. This part is more pronounced
for extruded films. Usually during the manufacturing of the base film, heat
is applied to convert dry powder blends or granules into a molten state for
calendering or extrusion into a plastic film [62]. Whilst still warm the
unsupported film is often subject to tension which can cause stretching in
the longitudinal direction. This manifests itself as shrinkage (recovery of
Converting Properties of PSAs 373

Figure 7.4 Schematical representation of the shrinkage of a PSA laminate: A)


nonprinted material; B) printed material; 1) release paper; 2) release layer; 3)
adhesive layer; 4) face stock material; 5) printing ink.

strain) when the product is reheated to the temperature at which the tension
was applied. The laminator must act to eliminate this hazard with film made
under controlled conditions. Additionally, product shrinkage can be
induced by residual solvents after printing, or through plasticizer migration
out of the film into the adhesive or substrate.
Flagging or wing-up of the labels on curved surfaces due to insufficient
adhesion of the PSAs to the substrate is different from flagging after coating
(i.e., the loss of the lay-flat). This phenomenon is due to the chemical influ-
ence of the adhesive or printing inks [59]. Wing-up depends on the following
factors: the adhesive/printing ink composition, the thickness of the film, the
adhesive properties of the PSAs, the thickness of the adhesive/ink film, the
differences in the modulus of coating/coated material, and the coating width
(full or limited). Figure 7.5 illustrates flagging after printing of a PSA film.
Another disadvantage of PVC used as face stock material is the
requirement for special coating machines to avoid tensile forces during
coating and laminating [63]. Plasticizer migration from the PVC may
influence not only the face stock shrinkage and thickness, but also the
adhesive properties. Peel adhesion values change when removable PSAs are
used on plasticized face stock [64]. An additional disadvantage of the use of
PVC as face stock is its sensitivity towards adhesive rheology. This is a
general phenomenon also characteristic for other film face stock materials.
374 Chapter 7

Figure 7.5 Flagging after printing of a PSA film: 1) nonprinted; 2) printed; 3)


applied.

On the other hand, this phenomenon is induced by and related to the


adhesive. Practically, the poor coating rheology typical of conventional latex
adhesives manifests itself in several ways: first, there is a tendency in roll
applications to form ridges parallel to the direction of coating. Ridges
adversely affect adhesive performance by reducing the effective area of
contact with the secondary substrate, but more importantly they substan-
tially detract from the appearance of a lamination when a clear film is
used (see Section 2.9 of Chap. 8). PVC has a temperature resistance of about
70 C, which is low compared to cellulose acetate (120 C). Processing of PVC
needs special machines because of its high elongation. Peel values on
plasticized PVC may vary [65]. Stabilizers of PVC (e.g., Pb, Sn) cause
oxidation of the tackifier resin and thus loss of the tack. Thickness variations
of vinyl labels often cause problems in the applications [66]. Generally,
thickness variations of 10% are observed during the manufacture of PVC
film. Other data show that 8% profile tolerances are common [67].
The environment and product resistant properties of film labels are
significant advantages [68]. These advantages are obvious in high humidity
environments and they also provide a significant marketing advantage. The
polar surface of PVC provides good printability (i.e., PVC displays good
processability and printability). PVC films used as face stock materials
may be cured via electron beam [69]. They may be coated with silicones using
electron beam-curing. The back side of the film may be coated via screen
printing with a PSA; thus a ‘‘mono-web’’ material may be manufactured.
Converting Properties of PSAs 375

In the range of plastic films PVC displays the best balance between
coating, converting, and end-use properties (e.g., good wettability, good
anchorage, high drying temperature resistance, good printability, good
weatherability, and nonflammability). Its replacement with polyolefins is
due mainly to environmental considerations. There are some differences
between the properties of clear and opaque PVC:
Soft, clear PVC contains more plasticizer; consequently, shrinkage
values are higher for clear PVC than for opaque film.
Soft, clear PVC is more flexible than opaque film, thus bonding
wet adhesion (BWA) values are better.
Soft, clear PVC is more ‘‘plastic’’ than opaque film, hence cutting
properties are poorer than those for opaque film.
Soft, clear PVC has lower surface tension values, therefore wet-out
is better on opaque film.
The range of polyolefins used as face stock materials is shown in Table 7.3,
with each kind of plastic film displaying its special advantages; in parallel to
PVC other PSA film materials were developed [70].
The most dynamic segment of other face stock materials is that of the
polyolefins, mainly polyethylene and polypropylene. The properties of face
stock materials used for tapes were summarized by Meinel [71]. A detailed
description of the polyolefin films used as packaging and face stock material
is given by Placzek [72–74] and their printability is discussed by Verseau [47].
Polyolefin carrier materials for different PSPs and their manufacture are
presented in [17]. Polyethylene face stocks offer excellent chemical, solvent,
moisture, and shrinkage resistance. Because of its flexibility, polyethylene
was used successfully as squeezable label material on plastic packages [30];
there are different kinds of polyethylene, the most important materials being

Table 7.3 Performance Characteristics of Different Plastic Films Used as Face


Stock Materials

Performance Characteristics of Plastic Films

Criteria LDPE HDPE OPP Hard PVC PET

Stiffness Stiffness Low Medium Good Very good Very good


Die-cuttability Very low Medium Good Very good Very good
Thermal resistance Very low Low Good Good Very good
Ductility High Medium Low Low Low
Tear resistance High Medium Very low Low Very low
Coatability Low Low Low Very good Medium
376 Chapter 7

LDPE and HDPE. LDPE is clear and HDPE has a slight white color
(opacity) [75]. HDPE is generally used for labels dispensed with labeling
guns; it is used for plastic bottle labeling (oils, chemicals, cosmetics). LDPE
possesses a good reversible deformability and is adequate for squeeze
bottles. Polyethylene is sensitive towards solutions of surface active agents
and environmental stress cracking [57]. It is temperature sensitive, therefore
the drying temperature should be less than 50 C during printing. Poly-
ethylene is used as face stock material alone or as a composite (co-extrudate
or co-laminate). Polyethylene-coated paper (cardboard) or nonwoven
coated paper are used for special labels. The use of polyethylene as release
liner will be discussed later (see Section 1.7 of Chap. 9). It should be
mentioned that top-coated polyethylene may be used also.
Polypropylene films offer high clarity and good resistance properties.
Polypropylene labels offer potential as lower cost alternatives to polyester
labels in numerous applications [30]. Polypropylene is stiffer, cleaner, and
has better temperature resistance than polyethylene. Special polypropylenes
may be subjected to radiation sterilization. The drying temperature when
printing on polypropylene may attain 60 C [57] oriented polypropylene may
withstand 110 C. Oriented polypropylene face stock materials are manu-
factured with a resulting thickness range of 19–60 m [76]. Generally, OPP
face stock materials for labels display the following benefits [63]: strength
and durability, versatility, choice of finishes, printability, cost effectiveness,
compatibility of polypropylene label and container, environmental accept-
ability, a noncrosslinked character, and excellent die-cutting properties.
There are matte and glossy, transparent, white, and pearlized, corona-
treated (on both sides) films, with a gauge of 55–80 m [77].
A range of oriented polypropylene (OPP) labeling films for in-mold
and self-adhesive applications has been introduced. These films offer a
choice of finishes. Printable with all usual methods, they are expected to be
suitable for end-uses ranging from toiletries to food, car care, and gardening
products [78]. Mobil is manufacturing clear and opaque OPP films as face
stock material (Labe-LyteÕ ), with a gauge of 19–60 m, used in pressure-
sensitive label applications. They are said to be adequate for label converters
using different printing, and cutting techniques, to ensure sensitive roll stock
laminates of superior resistance to tearing and moisture, and breakage
reduction during stripping [79]. There are different polypropylene film
qualities used as face stock material, each with its advantages. A comparison
of the main properties of polyethylene and cast and oriented polypropylene
used as face stock material is given in Table 7.4.
Polypropylene films manufactured via the chill roll procedure show
fewer thickness variations than blown film (Table 7.4). A surface treatment
on blown polypropylene films lasts longer than on cast polypropylene
Converting Properties of PSAs 377

Table 7.4 CPP/OPP/PE, Used as Face Stock


Materials: Comparison of the Main Properties

Face Stock Material

Properties CPP OPP PE

Thickness þþ —— þþ
Profile —þ þþ ——
Transparency þ— þþ ——
Shrinkage þ— þþ ——
Storage dimensional stability þ— þþ ——
Strain resistance þ— þþ ——
Price þþ —— þþ

Table 7.5 The Choice of the Film Face Stock Material as a Function of the End-
Use of the Laminate

Performance Characteristics, Flm Material

End-Use Requirements PE PS PP Soft PVC

Chemical resistance Good Medium Good Low


Moisture resistance Very good Very good Very good Medium
Shrinkage resistance Good Very good Medium-low Low
Flexibility Good Low Medium-low Very good
Clarity Low Very good Very good Low

films [47]. Taking into account its mechanical and thermal characteristics
(stiffness and temperature resistance) as well as the optical ones, poly-
propylene seems more appropriate as a replacement for PVC than
polyethylene. A detailed discussion of PVC versus polyolefins as face
stock material was given by Hammerschmidt [80]. For large surface labels
cast PVC should be selected; a lower price alternative is calendered PVC.
Polyester displays better temperature resistance, cellulose acetate better
clarity. In any case, the choice of an adequate material also is influenced by
its stiffness (E modulus) (Table 7.5). In the same range of materials, the
required thickness may influence the formulation. As an example, for a
thickness of up to 60 m hard PVC is used; above 80–120 m soft PVC is
suggested.

Low Surface Energy Face Stock Materials. Materials with low


surface energy such as polyethylene and polypropylene are difficult to be
378 Chapter 7

coated. Wetting out on these materials is problematic and they exhibit a low
bond strength (anchorage) unless the surface is pretreated to improve
wettability and adhesion (see Chap. 2, Section 2 also). The surface tension of
water is 72 mN/m, while a typical solvent has a surface tension of 25–
30 mN/m. With many substrates having surface free energies in the range of
30–40 mN/m wetting proves to be problematic. In the case of water-based
PSA technology, the surface energy of the web must be raised 7–10 units
above the dyne level of the coating [81]. Polyethylene for tapes and labels
has to be treated [82]. Numerous pretreatment methods in order to improve
the wettability and adhesion on the surface of nonpolar face stock materials
are known [83]. Generally, they improve the polarity of the material through
physical or chemical effects [84].
The most common classical treatment methods are chemical treatment
[85], flame (thermal) treatment [86,87], and corona treatment [88,89].
Chemicals or flames produce a dehydrogenation/oxidation of the surface.
No universal theory exists for corona treatment, but the number of CO
groups increases as a function of the electrical energy (power) used [90]; for
example, about 2–3  104 carboxyl groups per cm2 are formed for an energy
level of 500 J/m2.
Corona treatment works by setting up an electrical field around the
film to be treated. This field ionizes the air molecules directly in its vicinity.
These ionized air molecules, which are usually in the form of ozone and/or
nitrous oxides, actually oxidize the surface of the film. The untreated surface
is composed entirely of carbon/hydrogen repeating units. After the corona
treatment, the carbon/hydrogen units are oxidized to include hydroxyls,
carboxylic acid, and occasionally amine, or nitrous groups. The chemical
change of the surface accounts for the increase of the critical surface tension.
The effect of this oxidation only penetrates 5–10 Å beneath the surface of the
material; this may not yield a permanent change of the film surface tension as
plastic films are in a constant state of change and untreated polymers can in
time migrate to the surface and cover the treated polymers. Generally,
different frequencies between 10–80 kHz are used; higher frequencies yield
a better treatment. The final surface tension depends on the material treated.
Polypropylene is more difficult to be treated than polyethylene.
A disadvantage of the corona treatment is its short duration. The
surface tension of a corona-treated blown film decreases from 56 to 36 mN/m;
this effect is due to the migration of slip additives [91].
Untreated polyethylene and other polyolefins are difficult face
materials for acrylic adhesives, and a large amount of effort has gone into
the search for a technique which will easily and reliably increase anchorage
without a significant change in the bulk properties. The most extensively
studied techniques for enhancing adhesive bonding are treatment with helium
Converting Properties of PSAs 379

gas plasma, oxygen gas plasma, or chromic acid. These and other surface
modification procedures suffer from the same common shortcoming as
corona treatment: the short life time of the treatment [92]. Corona treatment
is possible for different face stock materials, including polymers [93–98],
aluminum [99], or other similar materials [100,101]. During corona
treatment (in a 12–20 kV, 20 kHz electrical field) chain scission and
oxidation occur. Energy-rich electrons and ions break carbon-carbon
(C-C) and carbon-hydrogen (C-H) levels (3.7 and 4.3 eV, respectively).
The distance of the electrodes from the web influences the efficiency of the
processes; the best results are obtained with a gap of about 1 mm [102].
Although corona treatment is a complicated subject, one can conclude from
current data that corona treatment generates polar sites and then increases
the polar surface component of the free energy [103]. The possible functional
groups that may be formed are shown in Table 7.6.
The surface tension of polypropylene amounts to 28 mN/m; corona
treatment raises the surface tension up to 50 mN/m [102]. For printing
and adhesion the surface tension should be 40 mN/m [107]. HDPE is
more difficult to treat than LDPE. After a day the concentration of oxygen

Table 7.6 The Influence of the Corona Treatment on the


Polyolefinic Surface. Possible Polar Functional Groups
Formed [104–106]

Oxygen Nitrogen Halogen


Cl
R–OH RNH RC
H
O Cl
RC RCNHCH RC
OH O
R
C¼O
R
RCðNO2 Þ Cl
RC¼C RCH2C
RO OR H
380 Chapter 7

in the top (corona-treated) layer decreases by 13% [102]; after a year this
decrease is 50%. Metallized plastic films lose their treatment more rapidly.
Generally, the loss of the treatment effect (i.e., the decrease of the surface
tension as a function of the time) depends on the material nature
(formulation) and environmental conditions. Table 7.7 shows the decrease
of the surface tension of corona-treated polyolefin films as a function of the
time. Due to the slight degree of crosslinking in the modified surface, the
mobility of the modifying chemical groups is high in the case of polymer
materials, and therefore a decrease of the pretreatment effect occurs during
storage.
More stable surface energies can be obtained with semiorganic or
organic layers, only a few nanometers thick, which may be deposited by
plasma polymerization. Compared to conventional ‘‘wet’’ processes, the
plasma process offers new facilities for adhesion pretreatment [108], namely
fats, oils, and waxes can be removed without residues; polymer surfaces can
be roughened gently and without the use of aggressive chemical baths; the
wettability of nonpolar plastics can be improved by gently oxidizing the
surface; undefined plastics can be coated with a thin primer film. The films
are treated for several seconds or minutes in a ‘‘cold’’ plasma generated in a
common device, applying a high voltage to the process gas. Special
advantages of the plasma pretreatment are: easy control and automation,

Table 7.7 Shelf Life of the Corona


Treatment. Changes in the Surface Tension
During Storage of a Blown LDPE Film

Storage Time Surface Tension


(weeks) (mN/m)

0 49
1 47
2 46
3 46
4 46
5 50
6 49
7 49
8 48
9 49
10 48
11 50
12 50
Converting Properties of PSAs 381

difficult geometries can be treated, and nontoxic gases generally are used in
small quantities. Adhesive anchorage may be improved by grafting of the
polyolefin film. Monomers suitable for graft polymerization onto the
polyolefin film to promote the anchorage of normally tacky acrylic PSAs to
the film include acrylic acid, methacrylic acid (and esters thereof),
acrylamide, methacrylamide, sterically nonhindered tertiary alkyl acrylates
and methacrylamides, secondary alkyl acrylamides and methacrylamides
having three or fewer carbon atoms in the alkyl group, and N-vinyl
pyrrolidone. Crosslinking agents may be added to enhance the resistance of
the product [92]. For plasma treatment a high frequency electrical field, at
0.2–2 mbar is applied [109,110]. The open air plasma technology has a jet
tube where a gas (in most cases air) is blown through an electrical discharge
zone generating a highly active plasma which is then projected outside the
tube like a torch [111]. Corona treatment is oxidation, without monomers,
and without crosslinking reactions; totally differently, the plasma treatment
occurs as a sum of oxidation, reduction, functionalizing, crosslinking, and
deposition, with the aid of monomers. Figure 7.6 illustrates the change of
the surface energy and wetting angle after plasma treatment. Figure 7.7

Figure 7.6 The change of the 1) wetting angle and 2) surface energy after plasma
treatment.
382 Chapter 7

Figure 7.7 The change of the wetting angle after: 1) plasma, 2) chemical, and 3)
corona treatment.

illustrates the change of the wetting angle after plasma and corona
treatment in comparison to chemical treatment [112].
Plasma treatment yields homogeneous and stable surface tension
values superior to flame treatment [70]. Another possibility is sulfonation;
this procedure makes it possible to increase the surface tension from
21–23 mN/m to 41 mN/m. Technical data about low pressure plasma
technology are given by Liebel [113]. All the pretreatments induce differing
functional groups into the surface and cause molecular modification to
varying depths. The pretreatments have also been found to modify the
surface topography of the PP in some cases by roughening the surface and
causing the creation of new features, such as created by flame [114]. For
polymeric surfaces a considerable degree of oxygen is implanted onto the
surface, leading to a high amount of ketonic and hydroxyl groups. Up to
30% oxygen has been measured after treatment with plasma compared to
10% with corona treatment [115]. Surface analysis data of functional groups
and the O:C ratio of pretreated PP (by corona, flame, fluorination, vacuum
plasma, and air plasma) showed that vacuum plasma induced oxygen into
the surface in higher concentrations and closer to the surface than with
nitrogen. Laser was suggested for surface treatment also [116]. In order to
combine the advantages of controlled surface activation via oxidation
Converting Properties of PSAs 383

and of grafting, ‘‘corona deposition,’’ a treatment under controlled


atmosphere (of polymerizable or activatable compounds) was developed
also. Irradiation of polyolefin substrates, such as with an electron beam, to
improve the adhesion of various coatings also has been used [92, 117–122].
Sources to modify the HDPE surface include flame, laser, UV radiation as
well as discharge from electrical corona or plasma. These physical methods
produce a combination of chemical and morphological modifications
including crosslinking, oxidation, grafting of active/polar groups, chain
scission, ablation, and roughening at the polymer surface [123].
The flame treatment (flame process of Dr Kreidl) mainly is used on
hollow packaging items [124,125] and for packaging tapes. Flame treatment
can offer advantages where moisture must be driven out or a longer shelf life
of the pretreatment is required. Gas flame treatment is known as the
Kritshever process [126]. Polyethylene for tapes also may be chemically
pretreated [82,127]. Recipes for chemical treatment of polyethylene films
used as face stock material have been described [82,128]. A patented
pretreatment solution includes sodium or calcium hypochlorite, acetic
anhydride or acetic acid or succinic acid [129]. Milker and Koch [130] give a
detailed description of the chemical treatment of face stock materials. The
Lohmann/Ahlbrand procedure works with a fluor/inert gas mixture contain-
ing a 5–10% fluor volume. For polyolefins, the surface tension values have
attained 50 mN/m (Table 7.8) using about 2.5 kg fluor/100,000 m2 web [130].
The improvements in storage stability (shelf life) through the treat-
ment with fluor are better than for other procedures, generally lasting longer
than eight months. As an example, the immediate value of the surface
tension for polypropylene is higher than 54 mN/m; after 2 months it is 50
and it is 45–50 after eight months [130]. The various surface modification
methods for synthetic carrier films are described in detail in [17] and [131].

Table 7.8 Surface Treatment with Fluor: Comparison


of the Surface Tension Values Versus Corona-Treated
and Untreated Polyolefin Films

Surface Tension (mN/m)

Face Stock Without Fluor-


Material Treatment Corona-Treated Treated [130]

LDPE 30–32 35–44 > 54


HDPE 32 32–36 > 54
BOPP 32 35–38 > 54
384 Chapter 7

Polyolefins contain slip agents, antistatic agents, low molecular weight


oligomers, and waxes. Erucaamide or slip agent is commonly used in
polyethylene films to reduce the film’s coefficient of friction. A slip agent is
an effective modifier of the polyethylene film because of its low solubility
in the polyethylene matrix. This low solubility forces the slip agent to
migrate to the film surface, causing the usually tacky film surface to become
slippery and easily handled manually and by machine [133]. The migration
of slip agents does not allow the anchorage of the adhesive, thus
polyethylene-based face stock material should contain little or no amount
of slip agent.
Table 7.9 shows a comparison of the versatility of PVC and polypro-
pylene as face stock material for labels and tapes. The polarity and stiffness
of PVC allows its use in both application fields. The opportunities for
polypropylene as face stock material are superior for tapes. Table 7.10

Table 7.9 Comparison of the Versatility of PVC and PP as Face Stock Materials

Face Stock Material, Application Field

Tapes Labels

Criteria of Evaluation PVC PP PVC PP Ref.

Machining costs — Higher — — [132]


Primer for printing — Yes — Yes/No [50,80,132]
Release on the printed surface — Yes No No [80,132]
Low noise — Yes No No [50,66]

Table 7.10 The Main Performance Characteristics of PP and PE as Label Face


Stock Material

PE PP (nonoriented)
Performance
Characteristics LDPE HDPE Homopolymer Copolymer

Stiffness Low-medium Medium-high High Medium-low


Flexibility High Medium-low Low Medium
Dimensional stability Medium High Low Low
Cuttability Medium High Low Low
Printability Low Very low Medium Medium
Temperature resistance Low Medium High Low-medium
Clarity Low Very low High Medium
Weatherability Low Medium High Medium
Converting Properties of PSAs 385

displays the main characteristics of polyethylene and polypropylene used as


face stock material.
There are opportunities for other plastics in the development of
security, computer, battery, and other specialty labels. Several different film
materials, such as cellulose derivatives, polyester, polystyrene, etc., were
evaluated.
The following cellulose derivatives are used as face stock material:
cellulose acetate, cellulose triacetate, cellulose acetobutyrate, and cellulose
hydrate. Cellulose acetate possesses good dimensional stability, printability,
and oil resistance, and it is waterproof, but sensitive towards solvents.
Because of its tendency to be electrostatically loaded it remains difficult to
process [47,57]. Based on a 50-m cellulose acetate film which was modified
to produce a brittle film with a low tear strength, the film can be used as an
over-laminate seal on labels which are actually impossible to remove without
cleavage [134]. Tamper-proof, fragile films guarantee that the printed label
will simply be destroyed if it is attempted to remove it. Cellulose hydrate
(Zellglas) is generally plasticized and over-coated in order to improve its
printability or chemical resistance [71]. As plasticizers cyclohexyl phthalate,
and as lacquers nitrocellulose, maleinate resins, sulfonamide resins, or
dammar are used. Waxes for reducing the water sensitivity also are used;
these may make the printing difficult. MS-quality (moisture proof and
scalable) films need a higher drying temperature than normal Zellglas (P).
Polystyrene and rubber-modified polystyrene films are considered
semi-squeezable because of their stiffness [30]. These films are mostly used as
front and back labels on plastic containers. Polystyrene does not show enough
stability against fats and oils but possesses a good resistance against water,
alkaline and acid materials. Polystyrene is sensitive towards high temperature
and environmental stress cracking. Polystyrene films have to be corona char-
ged to a level of 55 mN/m, which is maintained for a year; good dimensional
stability, lower thickness, and good dispensing characteristics are obtained.
The properties of polyester (PET) films have been described [135,136].
Polyester exhibits very good mechanical properties and chemical resistance,
but is difficult to print. PET labels are used on polycarbonate (PC) bottles.
Squeezable bottles need squeezable labels. Hot-stamp PVC- or PET-based
films are manufactured (36–60 m); metallized PVC-, PET-, and OPP-based
films also are used for UV protection purposes. The coated and uncoated PET
are used as face stock material. For untreated polyester the adhesive should
have improved humidity resistance, good film clarity before and after humid
aging, no adhesive transfer failure mode, and a good cohesive strength [137].
Polyamide, with a good abrasion resistance, resistance against scratch-
ing, mechanical resistance, good resistance against oils, and very good
printability, also may be used as face stock material [138]. Polycarbonate
386 Chapter 7

may be used for touch sensitive electrical connections as a 250-m PSA-


coated film.
Polymer films, used as face stock material, show different water
sensitivity. Polyethylene, polypropylene, and PVC possess water absorption
of about 0.1% after immersion for 24 hr; PET absorbs more (0.3–0.5%)
[139]. Laminates of white, opaque polypropylene films with PET are
suitable for soft drink labels. The film has a very low coefficient of friction
of film to metal, which allows trouble-free high speed application; the film
shrinks 4.5% in the machine direction [140].

Comparison Between Paper and Plastics. The nature of the face stock
will be determined by the intended end-use. Typically it will be made of
paper, polymeric films, foils, or similar materials. Usually, it will be a face
material (mass per unit area) of at least 50 g/m2, although for paper it will
typically be at least 70 g/m2 as lower weight materials will make it difficult to
provide the necessary strength to be peeled off without tearing. Face
materials may include other flexible and rigid (solid) natural and synthetic
materials and their combinations, such as plastics, elastomers, solid metals
and foils, ceramics, wood, cardboard, and leather; essentially, they may be
of any form including film, solid articles, woven and nonwoven textile
materials, etc. Representative porous materials include paper and nonwoven
fabrics, which may be composed of cured pulp, rayon, polyester, acetate
fibers, cotton fibers, and blends thereof. The thickness of the film materials
is determined by the end-use of the adhesive-coated product, for polyolefins
in the range of about 0.025–5.0 mm. The most used film sheet materials,
PVC and polyolefins, both contain other micro- or macromolecular
additives. The polyolefin sheet material may contain additives such as
carbon black, calcium carbonate, silica, titanium dioxide, crosslinking
agents, dispersants, and extrusion aids; PVC contains plasticizers and
stabilizers. Both paper and plastic-based face stock material may have a
multilayer structure like top-coated paper, co-extruded or laminated films.
The most important difference between paper and plastic film face stock
materials lies in their porosity, polarity, and mechanical and chemical/
thermal resistance.
Paper is porous; its porosity allows the penetration of the adhesive and
improves its anchorage or the anchorage of top-coated layers. Therefore it is
easier to manufacture removable labels with paper as a face stock material.
On the other hand, the penetration of the adhesive may cause bleedthrough
of the adhesive (staining) or of the water (from water-based adhesives),
changing the dimensional and mechanical characteristics of the paper. The
porosity of the paper and its polarity due to its cellulose nature allow a good
anchorage of adhesives of quite different chemical nature. The most
Converting Properties of PSAs 387

important mechanical characteristics of the paper are its stiffness and its low
creep compared to plastic films. The modulus of elasticity for paper is about
2929–3750 mPa s, its creep rate is 4.42–1.07 Pa1 [141]. On the other hand,
the modulus of the main polyolefins is about 200–1400 mPa s [142], which is
much lower. The modulus of paper is strongly influenced by its water
content. Through full saturation paper loses 95% of its mechanical stability
(Table 7.11).
On the other hand, the modulus of plastic films may be improved
through the choice of the chemical basis. The toughness properties of
polymers increase with increasing molecular weight. The modulus of plastics
also may be improved using fibers. Another possibility is to use oriented
plastic films, where mechanical characteristics are superior, but depend on
the degree of orientation and direction (Table 7.12).
Another way to improve the stiffness is to build a sandwich structure
of plastic films as co-extruded or adhesive bonded laminates. Top coating
improves the stiffness also. Quite different film qualities are manufactured
via blown extrusion and casting. The chill roll process is generally superior,
with better transparency, dimensional tolerances, and dimensional stability,
but remains more expensive. The low creep resistance of plastics (cold flow)

Table 7.11 The Influence of the Humidity on the


Modulus of Paper at 65 C

Moisture Specific Elastic


Content (%) Modulus (kNm/g)

4 9.1
6 8.2
8 6.5
10 5.5
12 4.5
14 4.0
16 3.3

Table 7.12 The Influence of the Orientation of Plastic


Films on Their Modulus

E Modulus (N/mm2)

Material PVC PET MOPP BOPP

Value 4000 4500 200 2000


388 Chapter 7

causes high and partially uncontrolled elongation, and deformation of the


web during film extrusion or coating; therefore curling and shrinkage may
occur. These defects may be at least partially avoided using laminates or
films with a special profile (i.e., a higher thickness at the ‘‘edge’’ of the film)
[143]. The low stiffness of the films makes their high speed application (with
labeling guns) difficult.
Advances in label dispensing have also influenced materials used
for high speed applications. Conventional dispensers for paper-based PSAs
do not require special application systems because of the inherent stiffness
of the paper. With today’s conformable and squeezable PSA film products
sophisticated application systems are equipped with conventional dispensing
heads. As seen in Table 7.13, the stiffness, the dimensional stability, and
the die-cutting properties of paper as face stock material are generally
superior to those of different films [144]. Therefore pressure-sensitive
laminates based on thin plastic films usually possess release liners based on
voluminous paper.
On the other hand, plastic films bring a new level of transparency and
humidity resistance. Generally, the choice of an adequate plastic material
depends on the end-use of the label. Cast PVC is used for large labels,
calendered PVC for cheaper ones; PET is suggested for heat resistant labels,
cellulose acetate for clear ones. HDPE is recommended for dimensionally
stable items, pearlescent polypropylene for toiletries, extended polyethylene
in sleeving applications, etc. With regard to their stiffness the different face
stock materials may be listed in the following ranking [145]:

Stiffness: OPP > PVC > Paper > Polystyrene > Polyethylene ð7:5Þ

The mechanical and thermal properties of different carrier materials are


discussed in detail in [146]. As far as their cost effectiveness is concerned,

Table 7.13 Stiffness, Dimensional Stability and Die-Cutting


Properties: Common Films Versus Paper

Dimensional Die-Cutting
Material Stiffness Stability Properties

LDPE Low Fair Low


HDPE Fair Good Fair
OPP Good Good Good
H-PVC Very good Fair Good
PET Very good Very good Very good
Paper Very good Fair Very good
Converting Properties of PSAs 389

there is a quite different ranking (which may differ as future trends in plastic
prices vary):

Cost effectiveness:
Paper > OPP > Polyethylene > PVC > Polystyrene ð7:6Þ

As alluded to earlier porosity of the face stock material allows pene-


tration of the adhesive. Smoothness and the nonpolar character of the
face stock surface impart slight anchorage of the adhesive. In the case of
film-based face stock materials, plasticizers or slip additives may migrate to
the surface of the film. Emulsions or suspensions make polymers exude
surfactants on the film surface.
Fibrous materials may suffer fiber tear during rapid debonding. In all
these cases a top coat (primer) on the face stock surface is required. Table 7.14
summarizes the application and mechanisms of the primer coating.
It is not the aim of this book to discuss the manufacture and perform-
ances of various pressure-sensitive products, which are described in detail in
[1]. However, the understanding of the mechanism of how pressure-sensitive
adhesive-coated products work, requires a short presentation of the auxiliary
components (e.g., release liner) also.

Release Liners. The release liner does not play any role in the
bonding of the label, but it influences the coating, converting, and labeling
properties. Like the face stock material, the release liner is a sheet-like, thin,

Table 7.14 Application and Mechanism of the Primer Coating

Primer Effect as a Function of the Application Method of the PSA

Direct coating Transfer coating


1. Improves the anchorage of PSA: Improves the anchorage of PSA:
— no smoothening effect required — smoothens the paper face stock surface
— increases the chemical affinity of the — increases the chemical affinity of
the contact surface the contact surface
2. Strengthens the face stock: Strengthens the face stock:
— reinforcement of the fiber structure — reinforcement of the fiber structure
— strengthening of the surface layer — strengthening of the surface layer
3. Absorbs stresses Absorbs stresses
4. Stiffens the face stock material; Stiffens the face stock material;
changes the peel and cuttability changes the peel and cuttability
5. Limits adhesive bleedthrough Less important
6. Influences drying No importance
390 Chapter 7

solid state component, and is usually paper or film based. Its release,
dehesive nature is given by a special polymer coating (see Chap. 5). The
most important factors affecting the release liner include the nature of
the adhesive (chemical type, thickness, modulus, diluents), the nature of the
silicone coating (chemical composition, coating weight, film continuity,
degree of cure, crosslink density), the face material variations (roughness,
porosity), the laminate characteristics (paper age, laminate age, thickness
and modulus, adhesive coating method), and the matrix stripping operation
(speed, angle, physical dimensions) [147].
The most widely used material for release liners remains paper [26].
More than 50% of liner material is based on calendered paper, more
than 20% on top-coated paper, and only 2% on films. The liner is the
‘‘packaging material’’ for the adhesive, the carrier material for die-cutting,
and the transport material for the labels. The role of the paper as transport
(end-use) material for the label is replaced by plastic films in some
specific application fields such as medical labels, where polyethylene with
adequate stiffness and good die-cuttability is used [70]. Another application
field concerns metallic labels or PSA-coated materials, where the quite
different toughness of the metal and plastic substrate brings about low
peel adhesion or where the plastic-based release liner displays a better
weatherability.
Release papers are specially coated materials, normally made from
30–60 g/m2 high density papers (glassine types) which have a special coating
that repels the solidified PSAs. A silicone coating is generally used, but
release papers can carry a wax or another type of nonsticky coating. The
nature of the paper influences in a decisive manner the dehesive properties of
silicone liners. Low profile tolerances, chemical inertness, high barrier pro-
perties, good wetting out, smoothness, no yellowing, and high dimensional
stability are necessary [148]. The type of pore formation and surface structure
of the base paper exert a determining influence on the consumption of
silicone and the degree of film forming during the coating process. By means
of various measuring techniques for paper it is now possible to record the
influence of the base paper on the silicone coating. These include the mea-
surement of rigidity, porosity, and above all, penetration. The presence of
migrating species in a release coating can be disastrous to adhesive perfor-
mance. Very small quantities of silicone fluids such as the ones included in
antifoam agents can detackify rubber-based adhesives.
Release liners must fulfill a variety of demands. They must offer the
correct physical properties for the envisaged end-use, as well as allowing
the converter to use existing equipment and chosen silicone systems to
obtain the required release properties. Glassines, treated Krafts, polyolefin-
coated papers, and pigment-coated grades are available. A typical base
Converting Properties of PSAs 391

paper comprises clay, cellulose pulp fibers for strength, and short fibers for
smoothness and uniformity. In order to achieve a closed silicone film, with
less than 1 g/m2 of silicone, a very smooth and even surface is necessary. The
pigment coating must be free of pits and flaws if subsequent release
problems are to be avoided.
For several years PVC films have been used as release liners for rubber
and silicone PSAs. Better results were obtained using stiffer or siliconized
films. The main films used as liners for siliconizing include polyester, OPP,
LDPE, HDPE, and polystyrene. Films used as release liners make labeling
easier (photosensitive regulation), faster (the ratio stiffness/thickness is
better for films than paper), and different colors may be produced; there
is no humidity influence on the film, no rehumidification after drying is
necessary. Film release liners are smooth and impart smoothness to the
PSAs. No blocking through die-cutting occurs. As they can be molten,
recycling as polymer is possible.

2.2 Printability of the Laminate


The printing methods used for labels and the most important problems
occuring during printing PSA materials are summarized by Fust [149,150].
There are close correlations between printability and other technical
requirements of the material used for the production of labels [151]. If
a large metallized high-gloss surface such as gold, silver, or copper is
required, this can be best achieved with thermoplastic foils like PVC, PET,
or OPP. While PVC films can be easily printed, OPP foils may require
considerable development efforts in order to fulfill the special requirements
made on the quality of the printed surface. Furthermore, the mechanical
resistance which plays a significant role for the printing, die-cutting, or
matrix stripping, and for the application of PSA labels, must be included
into these basic considerations. The caliper of the substrate also is
important.
There are some differences between the printability of polyethylene
and polypropylene [47]. Polyethylene is printable by gravure, with a mini-
mum film thickness of 0.025 mm polypropylene of 0.012 mm. Polypropylene
is superior to polyethylene concerning the resistance towards fatty materials,
has better gas barrier properties, and higher elasticity. On the other
hand, the anchorage of printing inks on polypropylene improves only
after 12–24 hr. For paper the printability depends on the lay-flat properties.
Lay-flat depends on waving which is a function of the flexibility; the latter
depends on the paper weight. Paper with a 40–80 g/m2 base weight shows
waving with short wavelength, heavy papers (100 g/m2) show waving with
longer wavelengths [152].
392 Chapter 7

The use of screen printing for PSA laminates was addressed by Perner
[153]; screens with 21–43 mesh are used. Flex printing uses up to 48 gravure
lines (Raster). Gravure printing uses 60 line numbers; offset printing also
is carried out [140]. The most explosive growth area is label stock for
nonimpact printing applications [34]. Direct thermal printing is the largest
segment, followed by toner-based printing (laser, ion deposition, magneto-
graphy, etc). The nonimpact techniques of direct thermal and thermal
transfer printing are the most reliable and cost effective means of bar coding.
Printability depends mainly on the quality of the face material. For
laminates the nature of the face is a function of the adhesive also; adhesive/
face interactions change the face quality. Generally, the following pheno-
mena influence the printability, namely the lay-flat (flatness) of the laminate
(curl control), the dimensional stability of the laminate (and uniform
caliper), and the surface quality of the laminate (gloss, smoothness, opacity).
The lay-flat of the laminate generally depends on the face/adhesive
equilibrium and the moisture content, and is especially important for paper-
based labels. The dimensional stability of the laminate is a function of the face
material and its environment/adhesive resistance. For paper labels it depends
more on the laminate/moisture interaction; for film labels it is a function
of the liner/adhesive interaction, or ink/film, face stock/adhesive interaction.
Temperature sensitive cold flow of the adhesive (edge bleeding or
oozing) influences the printability (especially for heat generated laser
printing, or xerox printing). The surface quality of the laminate can be
influenced through adhesive migration in the face, or migration of the face
components in the adhesive layer (soft PVC). For paper labels adhesive
penetration changes the visual appearance of the face; special additives from
the adhesive (e.g., antioxidants) can make the anchorage of the ink difficult.
In general the mechanical stability of the laminate has its importance also.
Labels are now printed at speeds of 175–200 m/min and higher [34].

Lay-Flat
Pressure-sensitive adhesive laminates may lose their lay-flat properties
(i.e., flagging, curling, wing-up of the laminate occurs). The curling of PSA
films depends on the thickness and on the stiffness (elasticity) of the film
[154]. For paper or other fiber-like materials the construction of the web
(fiber direction) and fiber sensitivity towards water may have an influence on
the lay-flat properties [155]. On the other hand, shrinkage of the paper-
forming fibers as a function of the chemical composition, respectively
manufacturing technology, determines the wet elongation of the fibers. For
multilayer papers, wet elongation, elasticity, and thickness of the paper are
the most important factors influencing the lay-flat [155]. Wet elongation also
Converting Properties of PSAs 393

depends on the humidity of the medium, thus generally in industrial practice


webs (especially paper) are rehumidified in order to avoid curling due
to changes in the water content [12]. In general, sheet or roll materials
have different lay-flat problems. The ‘‘rolling effect’’ (concavity) is mainly
known for paper rolls; the ‘‘tunnel effect’’ (convexity) is displayed by sheets
(or films).
Eliminating preheating and reducing the humidity content of the
release liner (to 4.2–4.5%) may improve lay-flat. These phenomena may be
explained on the basis of the theoretical knowledge concerning the paper
structure. By water absorbtion or desorbtion, changes in the volume of
cellulose fibers occur [156], they swell or shrink (20% changes in the fiber
diameter). These changes occur across the original machine direction: at
30% relative humidity (RH) the water content of the paper increases to 4%;
at 50% RH the humidity of the paper reaches 8% at 70% RH a 9%
absolute value of the humidity is observed. Practically, lay-flat may be
improved using special release liners, with higher thickness, and with
controlled humidity (corresponding to the equilibrium content at 55% RH
and 20 C). ‘‘Hygrorest’’ [157] is such a product. Single cut labels after curing
do not lie flat in the delivery boxes. Differences in moisture desorption of
face papers cause edge wrinkling upon conversion [158]; lay-flat depends on
the wet elongation of the paper [159]. For a multilayer construction the
deformation depends on the tension s; this tension is a function of the
modulus E, elongation ", and wet elongation e, or

 ¼ Eð2 eÞ ð7:7Þ

Taking into account the distance between medium and bottom layer (Z ), the
elongation is given by the following equation:

e ¼ e þ Zk ð7:8Þ

where k is the inverse of the radius of curvature.

Dimensional Stability
Dimensional stability is exemplified by shrinkage. Shrinkage of soft PVC
films is affected by the stability of the film and its anchorage (if mounted) on
the substrate (Figs. 7.8 and 7.9). Prestretched (e.g., calendered or cast)
plastic films tend to shrink if subjected to elevated temperatures. Mounted
films shrink back less than unmounted ones, because of the fixing effect
of the adhesive. In the case of labels, cohesive strength prevents shrinkage
and subsequent dirt accumulation on the exposed adhesive. Cohesive
394 Chapter 7

Figure 7.8 Parameters influencing the shrinkage of a PVC film.

Figure 7.9 Parameters influencing the dimensional stability of a PVC film.

strength depends on the built-in cohesive strength of the adhesive and on the
time-temperature stability of the adhesive (Fig. 7.10).
Cellulose acetate used in tapes exhibits a shrinkage of 0.6% (24 hr at
71 C) [160]. Thermo-fixed OPP is shrink resistant up to 110 C [161]. On the
other hand, humidity changes can cause a shrinkage of 1% in cellulose films.
Drechsler [162] demonstrated the influence of the surface layer on the base
film, and on the shrinkage of the base film. He states that a 30 m biaxially
oriented polypropylene (BOPP) film shrinks at 70 C. A polymer lacquer
on both sides increases the temperature resistance to 75–78 C; a coating
of acrylic PSAs increases the shrinkage resistance to 75–80 C. The shrinkage
of biaxially oriented polypropylene (8% MD, 4% TD) is more pronounced
than for oriented polyester (3% MD, 2% TD). Paper also undergoes
shrinkage.
Converting Properties of PSAs 395

Figure 7.10 Dependence of the shrinkage on the cohesive strength of the adhesive
and factors influencing it.

Drying of the paper during laminating causes a shrinkage of 0.8–1%


TD (corresponding to a decrease of the relative humidity from 50 to 30%).
Dimensional changes of the label paper produced by environmental
humidity may occur during printing or cutting too [163]. The test methods
to quantify the shrinkage will be discussed in Section 3.1 of Chap. 10.

Surface Quality of Labels


The surface appearance of the label is influenced by the adhesive (i.e., its
migration, penetration, bleeding, and staining—see Chap. 2, Section 3.2;
Chap. 3, Section 2; Chap. 5, Section 2.4; and Chap. 6 also). Migration may
also produce coating and/or printing defects [164]. Penetration is a function
of the face stock material, of the adhesive, and of the manufacturing
parameters, and depends on the porosity of the paper, its humidity content,
and pH [165]. Dispersion-based PSAs display a higher tendency to migration
than solvent-based ones. Therefore migration tests are carried out for water-
based PSAs at a lower temperature (i.e., 60 C), than for solvent-based
adhesives (i.e., 71 C) [166]. The activation temperature of the migration
depends on the formulation recipe [167]. For tackified PSAs the activation
temperature of the migration decreases with increasing resin content. There
are two separate kinds of staining: pressure induced (roll products) and heat
induced (see Chaps. 2, 8, and 10). The converting properties of the laminate
also are influenced by its aging. The change of the laminate properties due to
aging is discussed in Chaps. 8 and 10.

Cuttability of the Pressure-Sensitive Laminate


There are only a few data available in the literature about the cuttability of
self-adhesive laminates. Die-cutting properties of the PSA label depend on
396 Chapter 7

the strength and type (failure) of the face paper (matrix removal), the type of
adhesive and silicone, the adhesive-silicone release force, the die-cutting
properties of the adhesive and of the face paper, the properties of the
siliconized paper, and the cutting process parameters (see Chap. 2,
Sections 3.1 and 3.2 and Chap. 4 also). The evaluation of the viscosity is
an index for holding power, telescoping, oozing, and die-cutting properties
[168]. The influence of cutting rate and temperature was studied by
Matschke [169]. The manufacturing parameters of the film and its thickness
also influence the cuttability. The adhesive affects cuttability via its recipe
and aging. The converting (i.e., unwinding and slitting of tapes) is strongly
influenced by the machine parameters. For tapes the unwinding rate and the
temperature of the film influence the adhesion. The unwinding (peeling)
force/is a function of the adhesion, modulus, and thickness (coating weight)
of the PSA, and the thickness, modulus, and width of the tape. The modulus
of elasticity of the adhesive depends on the speed of unwinding and
temperature. A similar behavior is observed when cutting/die-cutting PSA
label materials. Different cutting angles are necessary for cutting different
labels: gummed paper (22 ), plastics (19–24 ), soft PVC (22 ), and
polyethylene (25 ) [170]. On the other hand, low coating weights provide
‘‘clean’’ converting conditions (no oozing) [171]. For good cuttability the
PSAs should exhibit moderate cohesion; too high values of the cohesion
results in ‘‘pull out,’’ while too little strength allows smearing [172]. The
cuttability of paper can be improved using lubricants [173].
Cuttability of the laminate is a cold flow related characteristic; hence
there are two distinct aspects of this property: static and dynamic cuttability.
Static cuttability is the time dependent behavior of the cut material,
displaying more or less cold flow of the adhesive, and smearing of the side of
the roll/sheet material. It is very important to remember that low level forces
(due to the material’s own weight or to the winding tension) may cause cold
flow of the adhesive. On the other hand, dynamic cuttability related to the
cold flow of the adhesive during cutting includes two different aspects:
cuttability and die-cuttability. Cuttability is necessary for the initial cutting
(dimensioning) of the roll/sheet materials, and it means cutting through the
whole laminate. Die-cuttability refers to cutting-out the labels, while leaving
the carrier release sheet intact. Adhesive cold flow influences in the same
manner both cuttability and die-cuttability; the influence of the liner and of
the nature of the release may be different.

Cuttability of Sheet Materials: General Aspects


Paper label products have to display the cohesion and heat resistance needed
for sheet stock subjected to flying knife (cutting of the rolls in TD, i.e.,
Converting Properties of PSAs 397

Figure 7.11 The main cutting methods influencing cuttability.

slitting) and guillotine blade converting (cutting of the sheets) (Fig. 7.11).
Cutting, guillotining, and die-cutting are mechanical operations carried out
at high frequency. The less viscous and more rigid the response of the
polymer the cleaner the process tends to be. If viscous flow within the
polymer is significant during the converting operation, poor die-cutting
(characterized by unsatisfactory matrix stripping) or poor guillotining (knife
fouling) can result. Formulating approaches which increase the high
frequency modulus will enhance the overall convertability.
Evaluation of Cuttability. There are no laboratory methods allowing
the evaluation of cuttability properties. Different laboratory tests only help
to optimize an adhesive formulation, its final performance characteristics
being dependent on the face stock, liner, and their interaction with the
adhesive. Final conclusions about cuttability can only be drawn on the basis
of real cutting tests, made on an industrial guillotine. Generally, every
converter has established specific evaluation methods for laminates. As
general criteria the smearing of the knife (front side and back side) and the
smearing of the cut material should be considered. The level of the adhesive
deposits on the knife, the smearing of the cut strips by the adhesive layer,
and blocking of the cut sheets (no moving capability) characterize the
cuttability of the laminate. Table 7.15 summarizes the cuttability results for
different laminates illustrating how to quantify the different aspects of the
cuttability. Steps of the cutting process (cutting through, guillotining) and
the phenomena associated with the cutting were shown in Fig. 2.16.
The following phenomena can be observed during the cutting operation:
smearing of the knife;
smearing of the cut material (sheets and strips);
blocking of the cut material;
adhesion of the cut material to the knife.
398

Table 7.15 Cuttability for Different Laminates (Evaluated Using an Internal Cuttability Index)

Laminate Characteristics Cutting Characteristics

Aspect of the cut strips Aspect of the cut surface

Comments Index Comments Index

Coating Coating Coating Number of cuts Number of cuts Number of cuts Number of cuts
Adhesive weight weight weight
type (g/m2) (g/m2) (g/m2) 10 35 10 35 10 35 10 35

WB AC 24 80 100 DM DM 1 1 ac, NM PM, ac 4 5


WB AC 24 80 100 DM DM 1 1 ac, NM ac, NM 4 5
WB AC 20 80 85 DM DM 1 1 ac, PM ac, PM 3 4.5
WB AC/CSBR 26 80 85 DM sac, PM 1 2 ac, NM ac, NM 5 5
WB AC/CSBR 28 80 primed 85 DM DM 1 1 DM DM 1 1
SB AC 19 70 88 DM DM 1 1 DM DM 1 2
WB AC 20 78 95 DM DM 1 1 sac, PM sac, PM 1 3
Crosslinked HM 22 79 100 DM DM 1 1 DM DM 1 1
Chapter 7
WB AC 2 3 30 4 Blocked 1 16
WB AC 2 3 30 4 Pulled up 5 23
WB AC 2 3 25 3.5 Blocked 1 17
WB AC/CSBR 2 3 40 5 Blocked 1 22
WB AC/CSBR 2 3.5 25 3.5 Blocked 1 12
SB AC 0 1 20 2 Blocked 1 9
WB AC 0.5 2 30 4 Blocked 1 13
Crosslinked HM 0 1 0.5 1 Blocked 1 7
The lower the overall index, the better the cuttability.
Converting Properties of PSAs

WB AC, water-based acrylic; CSBR, carboxylated styrene-butadiene rubber; SB AC, solvent-based acrylic; HM, hot-melt; DM, dry and removable;
PM, partially removable; NM, not removable; sac, slightly adhesive coated; ac, adhesive coated.
399
400 Chapter 7

Blocking and pulling up of the cut material are secondary phenomena,


produced by the smearing of the cutting and cut surfaces. In order to obtain
a good cuttability (C ) one requires no smearing of the surfaces and low
adhesion of the smeared surface. There is a need for no adhesive flow out
(Fa), absorption of the smeared adhesive by the cut surfaces (Aa), and the
lack of adhesion of the smeared adhesive (Da). Hence,

C ¼ f ½ðAa  Da Þ, ðFa Þ1  ð7:9Þ

Adhesive Flow Out. Important parameters include smearing of the


edges (oozing), edge flow, face bleed, and gumming. Adhesive flow out is a
function of the cold flow or creep of the adhesive. Creep of the pure (bulky)
adhesive is different from that of the adhesive in the laminate and Fa
depends on the adhesive and on the laminate:

Fa ¼ f (Adhesive, Laminate) (7 .10)

The adhesive flow out is a function of the nature and amount of the adhesive,
the laminate components (nature and dimensions), and laminate structure
(see Chap. 2, Section 3.1; Chap. 3, Section 2; and Chap. 4, Section 1.1 also).
On the other hand, creep also depends on the cutting conditions (nature and
level of the mechanical stress); the rate and distribution of the stress as well as
the working temperature influence the cold flow. Finally, Fa is a function of
the adhesive, the laminate, and working conditions:

Fa ¼ f (Adhesive, Laminate, Working conditions) (7 .11)

Like thermoplastics PSAs are viscous fluids exhibiting cold flow.


A PSA in a laminate is a viscous fluid, flowing under hydrodynamically
limited conditions (i.e., like a thin fluid layer between parallel surfaces). The
flow velocity depends on its distance from the top surface of the fluid layer
(i.e., from the coating weight):

Fa ¼ f (Adhesive nature, Adhesive amount) (7:12Þ

The adhesive nature influences the creep through the viscoelastic properties
and the creep from the laminate through its interaction with the laminate
components, which impart a resistance against the flow (Fig. 7.12).
The fluidity of the PSA depends on its viscoelastic properties
characterized by its viscosity and modulus and the dependence of these
properties on the stress rate and temperature (see Chap. 2). The cohesion of
the adhesive and the cohesion/adhesion balance characterize the viscous
behavior. The modulus and its dependence on the temperature must be
Converting Properties of PSAs 401

Figure 7.12 The influence of the adhesive on the creep.

known in order to make a first approximation about the fluidity of the


adhesive. For a better understanding the frequency dependence (tempera-
ture dependence) of the modulus (storage and loss modulus) must be
known. The value of the modulus (viscosity) and parameters influencing it
at ambient temperature, and the temperature/speed sensitivity of the
modulus (viscosity) must be discussed separately. The viscoelastic properties
of the adhesive (a) depend on the viscoelastic properties of the bulk
adhesive (ab), and on those of the composite adhesive layer (ac), or

a ¼ f ðab, ac Þ ð7:13Þ

The viscoelastic properties of the bulk adhesive are different from


those of the composite material. Generally, the viscoelastic properties of
composite solvent-based adhesives are superior to those of the bulk
material. The viscoelastic properties of water-based (composite) PSAs are
inferior to those of the bulk adhesive. This behavior can be explained by the
different nature of the additives. The viscoelastic properties of the bulk
material depend on its chemical nature, molecular weight, MWD, and
structure (crosslinked, branched, or linear) (see Chap. 2); hence,

ab ¼ f (Nature, Molecular weight, MWD, Structure) ð7:14Þ


402 Chapter 7

The tackified adhesive should also be considered as the bulk adhesive


although tackification (with plasticizer or resin) decreases the cohesion level.
Generally, low cold flow (good cuttability) implies more elasticity amid a
less viscous character of the viscoelastic fluid (i.e., higher cohesion). From
knowledge of the inversely proportional relation between cohesion and
adhesion, in a first approximation, good cuttability means high cohesion
and low adhesion:

C ¼ f ½ðCohesionÞ, ðAdhesionÞ1  ð7:15Þ

The chemical/physical structure of the adhesive determines its viscoelastic


characteristics and the dependence of the viscoelastic behavior on the
temperature/stress rate. During cutting, the temperature of the laminate
increases and, consequently, the viscosity of the material decreases.
Thus the value of viscosity at higher temperature (cohesion at higher
temperature) is very important for cuttability purposes. The most simple
method of measuring the temperature dependence of the viscosity is the test
of the shear at elevated temperatures, and the test of the hot shear (HS)
at different temperatures gives the hot shear gradient (GHS) (see Chaps. 6
and 10). Figure 7.13 illustrates the dependence of the hot shear on the
temperature for different tackified formulations.

Figure 7.13 The dependence of the hot shear for different tackified formulations.
1), 2), 3), 4), and 5) are different water-based acrylic PSA formulations.
Converting Properties of PSAs 403

Different formulations display a different temperature sensitivity of


the cohesion (Fig. 7.13). On the other hand, cutting tests carried out
with different formulations demonstrate that formulations with a limited
temperature sensitivity of the shear display better cuttability. Room tempe-
rature shear values are more important for the cohesion/adhesion balance,
influencing mainly the peel on polyethylene.
Cuttability and Adhesive Properties of Resin. The adhesive properties
of the resin influence the cuttability of the laminate. Unfortunately, there is
no theoretical or empirical equation clarifying the dependence of the
cuttability (C ) of the laminate on the adhesive properties (shear, peel, and
tack) of the adhesive;
C ¼ f ðAdhesive propertiesÞ ð7:16Þ
The best documented interdependence is the correlation of cuttability and
shear. Unfortunately the best known shear values are room temperature
shear values, and there are only a few hot shear data available; no hot shear
gradient data are known from the literature. Cuttability improvement is a
function of the shear increase (order of magnitude) and of the hot shear
gradient (GHS). Thus, the cuttability improvement (Ic) depends on the shear
as follows:

Ic ¼ f ðlog HSÞ, ðHSnþi  HSn Þ1 ð7:17Þ


where HSn denotes the hot shear at a given temperature, HSnþi is the hot
shear at another temperature with T at least 10 C, HS is the
improvement of the hot shear (HS ¼ HSi – HSo, where HSo is the hot
shear of a known product with an average cuttability). Table 7.16 illustrates
the dependence of the hot shear and GHS on the formulation (resin nature).

Table 7.16 Dependence of the Hot Shear and Hot Shear Gradient on the
Formulation.The Influence of the Resin Nature for Water-Based Acrylic
Formulations

Tackifier Hot Shear Values (min) ( C)

Code Nature Supplier 30 40 50 60 70

CF 52 Acid rosin A&W 90 35 12 10 11


DEG G Rosin ester DRT 390 170 72 45 21
MBG 64 Rosin ester Hercules 330 90 72 38 23
CF 8/21 Rosin ester A&W >1500 192 160 55 26
E-9241 HC-resin Exxon 230 144 90 45 40
Acrylic ¼ a 40/60 (wet/wet) weight blend of V205/80D (BASF).
404 Chapter 7

Although changes in the room temperature shear and slight changes in


hot shear are possible, no cuttability improvement can be observed (except
for CF 52). On the other hand, the study of the GHS and the dependence of
the cuttability (and peel) on GHS denote the necessity of a minimum GHS for
good cuttability (Fig. 7.14). It is evident that slight changes in the
formulation (using resins with a higher melting point) do not necessarily
impart an improved cuttability.
For shear tests the modulus of the adhesive is important also. The
contact area A between the adhesive and the face stock/release liner may be
estimated according [99] as a function of the modulus E, the roughness of
the surface 1/ , the limited contact area Al, and the density of the adhesive:

A ¼ Al ð1  eð Þ=E Þ ð7:18Þ

The viscous flow, creep, or cold flow of the adhesive may be charac-
terized by its cohesion (see Chaps. 2 and 3). Smearing by cutting will be
described by the elongation viscosity and relaxation of the fluid adhesive.
The coefficient of viscous traction o(t) as a ratio of tensile stress s(t)
and elongation rate 0 is [174]:

o ðtÞ ¼ ðtÞ=0 ð7:19Þ

Figure 7.14 Interdependence of 1) hot shear, 2) peel, and 3) cuttability.


Converting Properties of PSAs 405

On the other hand, the section A(t) during elongation also depends on the
elongation rate:

AðtÞ ¼ Ao e2 ð7:20Þ

The coefficient of viscous traction (elongation viscosity) includes an


ideal elastic and a viscoelastic component, both dependent on the modulus
E. The viscoelastic part depends on the relaxation time of the material. If
elongation occurs rapidly, the tension is given by the following equation:

ðtÞ ¼ ½EðtÞ þ ðtÞo ð7:21Þ

where E(t) is the tension relaxation modulus (a ratio of tension and


elongation) and c(t) is the relaxation function:

EðtÞ ¼ ½ðtÞo   ðtÞ ð7:22Þ

c(t) is approximated for rubber according to the following equation:

ðtÞ  f ð1:0Þt ð7:23Þ

For tests carried out at a constant rate:

ðtÞ ¼ Kðt  to Þ ð7:24Þ

where K is a constant and t is the time. The relaxation modulus for natural
rubber is not zero at temperatures as high as 100 C (i.e., rubber alone does
not behave like a fluid). Tackified rubber compositions (i.e., PSAs) display a
quite different behavior. These considerations show that the modulus and
generally the complex time/temperature dependent rheology of PSAs
influence the cold flow (shear/cuttability behavior) (i.e., it is difficult to
find a simple relation between shear resistance and cuttability).
Concerning the peel dependence on the cuttability, a 10–30% cutta-
bility improvement needs a 1000% peel decrease and/or a 50–100% tack
decrease (Figs. 7.15 and 7.16). Changes in the tack are common, but changes
in the peel are rarely possible (e.g., sheet/roll material). On the other hand,
changes in the hot shear of an order of magnitude are generally possible
only through crosslinking. In conclusion the improvement of the cuttability
of a given formulation from the point of view of the bulk adhesive remains
limited. Actually cuttability improvement by means of the composite
structure of the adhesive (laminate) seems to be more important.
406 Chapter 7

Figure 7.15 Interdependence of peel/cuttability. The dependence of the guillotine


cuttability on the peel of PSA film laminates that have a different peel level: 1), 2), 3),
4, and 5) are different water-based formulations.

Figure 7.16 The dependence of the cuttability on the tack. Cuttability (guillotine)
as a function of the rolling ball tack for different PSA formulations.
Converting Properties of PSAs 407

Influence of Adhesive Nature on Creep in the Laminate. Creep in the


laminate differs from the creep of the bulk material because of
the mechanical/chemical interaction between adhesive and solid laminate
components. For this mechanical/chemical anchorage a minimum level of
fluidity and tack is necessary. That is the reason why hard low-tack
formulations display poorer cuttability (showing fiber tear) than low-
tackified medium-tacky formulations. Thus, in a first approximation the
creep in the laminate is inversely proportional to the anchorage:

Creep ¼ f ð1=AnchorageÞ ð7:25Þ

C ¼ f ð1=AnchorageÞ ¼ f ðAdhesionÞ ð7:26Þ

Creep also depends on the cutting process conditions. The amplitude


of the deformation increases as a function of the nature of the applied
stresses as follows: 1, 1.3, and 1.5 for compressive, shear, and tensile forces
respectively [175]. Thus laminates with elastomers have to be designed to
resist tensile stresses. On the other hand, it is to be taken into account that
the temperature increase of the elastomer through hysteresis increases with
the frequency and the square of the amplitude of the deformation.
Fillers increase and softeners decrease the wall slip behavior of
polymer blends [99]. For rubber slippage through a high pressure capillary
viscometer, the flow volume possesses three different components, a wall slip
component Vg, a shear component Vs, and a middle layer component Vk so
that:

Vg ¼   R2  wg ðw Þ ð7:27Þ

Z tw
Vs ¼ ð  R3 Þ=ðw3 Þ ½r3 =ðÞdr ð7:28Þ
to

Vk ¼   r  ðw Þ2  wðrÞ ð7:29Þ

where wi is the flow rate,  is the viscosity,  is the stress, R is the radius of
the capillary, and r is the radius of the middle (shear and slip free) layer [99].
The flow volume generally depends on the stress applied, on the geometry of
the layer, and (for the middle layer only) on the viscosity. Therefore it can
be suggested that the stress distribution in the layer (i.e., anchorage of the
adhesive and stiffness of the solid state components) and geometry of the
408 Chapter 7

layer (coating weight) are more important for cold flow than the viscosity of
the adhesive (see Section 3.2 of Chap. 2). In a similar manner, plasticizers
and temperature promote wall slip, while fillers decrease it. The same
interdependence may be supposed for the cold flow (bleeding) of the
adhesive.
Dependence of Cold Flow on Adhesive Coating Weight. Cold flow (Fc)
of the adhesive in a laminate is hindered by marginal interaction of
the adhesive with the ‘‘walls’’ of the laminate. Because of the local character
of this interaction the middle layer displays free flow. It is evident that the
width of the free flowing layer depends on the coating weight and on the
width of the anchored layer. The anchored layer depends on the coating
technology and the nature of the solid laminate components; thus the free
flowing layer depends on the coating weight and coating technology (Tc).

Fc ¼ f ðFree flowing layerÞ ¼ f ðCw ,Tc Þ ð7:30Þ

In a first approximation the cuttability (C ) depends on the coating weight as


follows:

C ¼ f ½ðTc Þ, ðCw Þ1  ð7:31Þ

Dependence of Cold Flow on the Laminate. Cold flow in the laminate


depends on the width of the free flowing region, this being a function of the
anchorage. The anchorage of the adhesive depends on the face material (i.e.,
the anchorage in the laminate depends on the adhesive bonding on the face
and liner). On the other hand, the uniformity of the forces acting on the
laminate, their distribution, and the compression stress character in the first
cutting step are determined by the rigidity of the face. Paper face materials
are hygroscopic and there is a continuous water transfer between adhesive
and face, thereby influencing the composite character of the adhesive. The
solid components of the laminate influence the cold flow in the laminate
through the anchorage of the adhesive, the nature of the stress on the
adhesive, and the change in composite structure.
The anchorage of the adhesive on the face material (F) and liner (L)
depends on the smoothness (Sm) and chemical character (Ach) of both.
As to the smoothness, it is well known that papers of different quality are
used for face materials and release liners. The chemical affinity of the surface
may be changed using a dry coating (primer). The primer is used mainly
for removable coatings (with a softer formulation) and improves the
cuttability (see Chap. 2). Figure 7.17 summarizes the factors influencing the
anchorage.
Converting Properties of PSAs 409

Figure 7.17 Parameters influencing the anchorage of a PSA on the face stock
material.

The water content of the paper influences the anchorage. ‘‘Wet’’


anchorage is weaker than dry anchorage; thus:

Fc ¼ f [(Smoothness of F& L), (Chemical affinity of F& L)1 ]


(7:32Þ

C ¼ f ½ðAc Þ,ðSmÞ1  ð7:33Þ

C ¼ f ½ðAc , Primer of F& LÞ, ðSm, Humidity of F& LÞ1  ð7:34Þ

The chemical affinity of the face/release liner also depends on the


adhesive nature. Different adhesives display different adhesion (release force)
on different liners (Fig. 7.18). The release peel force influences the composite
behavior of the laminate (cuttability) and the labeling properties. A high
release force increases the yield stress of the liner resulting in unsatisfactory
matrix stripping (die-cuttability) possibly causing liner breakage.
As incompressible liquids PSAs support compression forces. In the
first step of the cutting operation the force acts perpendicularly on the whole
laminate (or laminate assembly). This situation changes only by the local
stress concentration and weakness of the solid laminate components, giving
410 Chapter 7

Figure 7.18 Dependence of the release force on the chemical nature of the PSA.

rise to the flexure of the face stock and shearing of the adhesive. The
stiffness of the solid laminate components influences the stress distribution
and the creep of the adhesive. Generally, the stiffness of the paper depends
on its composite structure and its humidity content (Fig. 7.19); both cold
flow (Fc) and cuttability (C) depend on the composite structure and
humidity:

Fc ¼ f [(Humidity of the F& L), (Composite Structure of F& L) 1 


ð7:35Þ

C ¼ f [(Composite Structure of F& L), (Humidity of F& L)1 ]


ð7:36Þ

Another parameter influencing the cuttability is the stiffness of the


laminate components. The flexural resistance of a laminate depends on the
siliconizing, on the adhesive nature and state (crosslinking, humidity), on
the laminate age, on the coating technology (direct/transfer), on the adhesive
amount (coating weight), and on the face stock/release nature. Generally,
Converting Properties of PSAs 411

Figure 7.19 Factors influencing the stiffness of a PSA laminate.

PSAs lower the flexural resistance (stiffness) of paper. A systematic study of


these parameters identifies the following factors influencing the stiffness
of the label: the materials (nature of face stock, release, and adhesive), the
thickness of the components (face stock, release liner thickness, and coating
weight), and the construction of the solid components. In this section,
the influence of these parameters on the stiffness of the label, and on the
properties (adhesive and end-use) of the PSA label will be illustrated. As can
be seen from Table 7.17, film laminates (plastic face stock/paper release liner
or filmic face stock/filmic release liner) are softer than paper-based ones. The
number of layers and their build-up into sandwich constructions also
influences the stiffness (Table 7.18). The nature and softness of the adhesive
also can change the stiffness of the label (Table 7.19) (see Chap. 2).
Siliconizing (one side or both) modifies the stiffness of the labels, eventually
increasing it (Table 7.20). The stiffness of the label also is a function of
the coating weight (Table 7.21). Engraved (embossed) face stock gives
(as supposed) better stiffness (Table 7.22). A primer coating (with a soft
primer) changes the stiffness only slightly, but improves the peel (Table 7.23).
The humidity modifies the modulus and stiffness of paper (Table 7.24). The
coefficient of friction of the paper (slip in the cutting machine) increases with
increasing humidity in the air [176].
The influence of the face stock material on the cuttability was dis-
cussed in Chap. 2. Here the dimensions and the mechanical characteristics
(mainly the modulus) of the face stock and their influence on the cuttability
are discussed. A high stiffness of paper face stock or liner can be obtained
by using a higher weight or a more voluminous paper grade (thickness). The
finish or type of fiber used also affects the stiffness; sulfite pulps are
generally stiffer and the difference in stiffness between papers made from
412 Chapter 7

Table 7.17 Stiffness of Film Labels Versus Paper Labels


with Different PSA Coatings

Face Stock Material

Stiffness
Adhesive Code Paper Film (mN/25 mm)

01 x 650
01 x 814
02 x 682
02 x 1014
03 x 632
03 x 961
04 x 148
04 x 331
05 x 648
05 x 833

Table 7.18 The Influence of the Number of Layers and Layer Build-Up on the
Stiffness of the PSA Laminate

Stiffness (mN/m)

Number of layers/build-up

2 3 4 5

Coating Weight (g/m2) FS-A FS-A-FS FS-A-FS-A FS-A-FS-A-FS

12.0 13 117 249 99


17.0 23 32 223 122
20.0 22 106 193 103
30.5 22.5 157 210 158
35.5 18 171 227 121
48.5 26 149 154 176
FS, Face stock; A, adhesive.

sulfite and sulfate pulps is about 20%. Mechanical strength and to a lesser
degree smoothness also are affected. However, this approach also reduces
ink hold-out (or water hold-out), thereby resulting in a higher stiffness
(at least theoretically). Generally, the thickness of the laminate is an index of
the stiffness of the laminate; laminates with a higher thickness exhibit a
better cuttability (Fig. 7.20).
Converting Properties of PSAs 413

Table 7.19 The Influence of the Nature and


Softness of the Adhesive on the Stiffness of the Label

Adhesive Nature
Stiffness
WB SB AC RR CSBR (mN/mm)

x x 698
x x 480
x x x 515
x x 483
x x 723

Table 7.20 The Influence of Siliconizing on the


Stiffness of the Paper

Number of Silicone Coating Stiffness


Layers on the Paper (mN/0 mm)

0 78.0
1 83.7
2 169.6

Table 7.21 The Influence of the Coating Weight on the


Stiffness of the Laminate

Coating Weight Laminate Stiffness


Laminate Code (g/m2) (mN/10 mm)

1 12.0 127
22.0 122
51.0 146
2 14.5 145
22.0 107
43.0 113

Generally, the stiffness of the liner is tested in order to evaluate its


machinability [177]. The stiffness of the liner depends on its specific weight;
the used specific weight (stiffness) also is a function of the face material. For
paper-based liners 55–100 g/m2 material, for film coatings 120–180 g/m2
papers are proposed [177]. The specific weight of the liner and its flexural
414 Chapter 7

Table 7.22 Dependence of the Stiffness on the


Embossing of the Face Stock Material

Stiffness (mN/m)

Face
Stock Quality Face stock Release liner PSA laminate

Normal 77 40 208
Embossed 155 40 470

Table 7.23 The Influence of the Primer on the Stiffness of the Laminate; the
Influence of the Dwell Time on the Stiffness

Stiffness of the Laminate and Laminate Components (mN/cm)

Dwell Laminate
Time (days) Face stock Adhesive Liner Laminate with primer

0 162 96 74 332 346


3 162 96 74 334 367
4 162 101 74 337 385

Table 7.24 The Influence of Solvents on the Stiffness


of Face Stock Paper; Immersion of Paper in Solvents

Stiffness of the Paper


Solvent (mN/10 mm)

None 166
Water 22
Toluene 153
Plasticizer 161
Oil 150

resistance (stiffness) MD and CD should be carefully controlled. In order to


evaluate the main parameters influencing the flexural resistance of the label,
the stiffness of a label can be estimated as follows. The stiffness S is a
function of the modulus E and thickness h of the material:

S ¼ f ðE,hÞ ð7:37Þ
Converting Properties of PSAs 415

Figure 7.20 The influence of the laminate thickness on the cuttability. 1) Through 6)
are different PSA laminates.

For a sandwich structure (label) the stiffness is the sum of the components,
where the final modulus and thickness are the sum of the components:

E ¼ Ei ð7:38Þ

h ¼ hi ð7:39Þ

where

hsandwich ¼ hfacestock þ hliner þ hadhesive ð7:40Þ

Assuming that:

hliner ¼ 1:5hfacestock ð7:41Þ

hadhesive ¼ 1:5hfacestock ð7:42Þ

then:

hsandwich ¼ 3hfacestock ð7:43Þ


416 Chapter 7

According to the dependence of the moment of inertia on the


thickness, the stiffness depends on h following the equation:

S ¼ f ðh3 Þ ð7:44Þ

Thus the increase of S with thickness changes for a label comparatively to


face stock material is given by:

S ¼ Fð3H 3 Þ ð7:45Þ

On the other hand, there is a decrease of the laminate modulus caused by the
adhesive; in a first approximation:

Es ¼ Efs =3 þ Efs =3 þ 0:1Efs =3 ¼ 2:1Efs =3 ¼ 0:7Efs ð7:46Þ

where Es is the modulus of the laminate and Efs is the modulus of the face
stock. Thus the change of the stiffness for a label, as compared to the pure
face stock material, will be given in a first evaluation, by the following
equation:

S ¼ f ½3h3 , 0:7E ð7:47Þ

The stiffness of the label is higher than that of the face stock material
and the most important factor influencing it is the thickness of the
components. An exponential dependence of 2.5 was found for cardboard
stiffness as a function of its thickness [178]; 4–5% thickness tolerances lead
to a 15% change in stiffness.
Influence of Other Parameters on Cuttability. Improved crossdirec-
tion tensile strength allows faster die cutting and stripping during converting
[179]. Laminates with a different humidity content display different
cuttability. The residual water content of water-based adhesives is less
than 5%, and as the equilibrium water content of the paper exceeds 5% the
paper acts as a continuous water reservoir for the adhesive; therefore
conditioning of the laminate before cutting is necessary.

Fc ¼ f (Humidity of F& L) ð7:48Þ

C ¼ f (Humidity of F& L1 Þ ð7:49Þ

There are two kinds of cold flow conditions: cold flow of PSAs under
high speed and high forces (cutting, die-cutting, slitting), and cold flow of
PSAs under low speed and low forces (storage). High speed stresses act like
Converting Properties of PSAs 417

normal stresses at low temperature. Low speed stresses act at normal


temperatures, thus their action is continuous and cannot be avoided; they
are characteristic for both sheet and roll materials. Cuttability depends on
the laminate and the cutting machine. ‘‘Engraved’’ knives lead to better
results than normal ones. Static cold flow depends mainly on the storage
conditions (temperature and humidity); for roll materials static cold flow is
also a function of the winding tension.
The scope of the formula is to design formulations exhibiting no
adhesive flow (oozing). The fluid character of the PSAs is necessary for its
bonding, thus it is not possible to completely avoid the flow of the adhesive.
Taking into account a certain level of adhesive flow, there is a need to
minimize its tack, in order to avoid smearing of the cutting and cut surface,
and the displacement of the cut sheets. Smearing of the cut and cutting
surface, and the instantaneous displacement of the cut strips are related to
the elevated tack level of the adhesive. Thus, in a first approximation:

Displacement ¼ f (Smearing) ¼ f ðTackÞ ð7:50Þ

Smearing and displacement are also functions of the peel, but in this
case one can assume that the instantaneous peel is not greater than the
instantaneous tack (the real debonding force is zero), thus the phenomenon
is regulated by tack;

C ¼ f (Tack1 ) (7 .51)

On this basis, the known (and potential) use of a tack decreasing agent
(silicones) for high tack solvent-based or hot-melt PSAs can be explained
(see Chap. 8).
Smearing of the surface and displacement of the strips are tack related,
thus they are coating weight-dependent phenomena. The coating weight—
the adhesive amount deposited as a consequence of the flow out and transfer
to the knife—depends on the total surface absorbtivity of the paper. Thick,
voluminous papers contain a smaller (relative) surface of adhesive and thus
display less tack:

Displacement ¼ f (Smearing) ¼ f (Cw , Surface1 Þ ð7:52Þ

C ¼ f (Smearing1 ) ¼ f (Front surface) ð7:53Þ

Thick, voluminous release liners should be used for water-based coatings


on soft PVC. The influence of the type of liner (stiffness) on the cuttability of
the laminate was confirmed in the field.
418 Chapter 7

Cuttability Differences Between Paper/Film Laminates. Generally,


paper laminates are based on tackified formulations, with a high tack and low
cohesion. Tack values of untackified film coatings are generally 0.1–0.5 times
smaller from those of tackified paper laminates. On the other hand, cohesion
values of the nontackified film coatings are 5–10 times those of the tackified
ones. Thus flow out of the film coatings is less prevalent than cold flow from
paper laminates; on the other hand, adhesion of the smeared surface is low.
The most important problem for film coatings is the absorption of the
smeared adhesive. In order to quantify the parameters influencing the
cuttability of the film/paper laminates, one should differentiate between
paper and film laminates. On the basis of practical experience the following
coefficient can be associated with the parameters used in Eq. (7.9):
Cfilm ¼ f ½ð1AA Þ, ð0:5D1 1
A Þ, ð0:8FA Þ ð7:54Þ

Cpoper ¼ f ½ð0:5AA Þ, ð0:8D1 1


A Þ, ð1FA Þ ð7:55Þ

where AA is the absorbtion of the smeared adhesives on the cut surfaces, DA


is the lack of adhesion of the smeared adhesive, and FA is adhesive flow out.
Quantitative Evaluation of Cuttability. On the basis of the above
equations the following interdependence between the parameters influencing
the cuttability (C ) can be defined:

C ¼ f ðAcoh ,Aanc ,A1 1 1 1


t ,Adim Þ, ðFdim ,Fq ,Fhum Þ, ðRdim ,Rq ,Rhum Þ ð7:56Þ

where:
Acoh ¼ cohesion of the adhesive
Aanc ¼ anchorage of the adhesive
At ¼ tack of the adhesive
Adim ¼ dimensions of the adhesive layer
Fdim ¼ dimensions of the face material
Fq ¼ quality of the face material
Fhum ¼ humidity content of the face stock
Rdim ¼ dimensions of the release liner
Rq ¼ quality of the release liner
Rhum ¼ humidity content of the release liner
The above relation becomes:

C ¼ f f½ðAcoh ,Actt ÞðAt Þ,ðAt ÞðAcw ,Ct Þ, ½ðFth ÞðFsm ,Fst ,Faf Þ=ðFhum Þ,
ð7:57Þ
½ðRth ÞðRsm ,Rst ,Raf Þ=ðRhum Þg
Converting Properties of PSAs 419

where:
Actt ¼ time/temperature dependence of the adhesion/cohesion
Acw ¼ adhesive coating weight
Ct ¼ coating technology
Faf ¼ face material affinity toward the adhesive
Fsm ¼ face material smoothness
Fst ¼ face material stiffness
Fth ¼ face material thickness
Raf ¼ release liner affinity toward the adhesive
Rsm ¼ release liner smoothness
Rst ¼ release liner stiffness
Rth ¼ release liner thickness
As demonstrated in [180], calculating the relative influence of the
adhesive A, face material F, and release liner R on the cuttability, one finds;

C ¼ f ½ð15:5%AÞ, ð42:0%FÞ, ð43:6%RÞ ð7:58Þ

Summarizing, the cuttability is more a function of the laminate structure


and coating technology than of the adhesive.

REFERENCES

1. I. Benedek, Development and Manufacture of Pressure Sensitive Products,


Marcel Dekker, New York, 1999, Chapter 7.
2. P. Herzog, 15th Munich Adhesives and Finishing Seminar, 1990, p. 3.
3. J.J. Elmendorp, D.J. Anderson and H. De Koning, ‘‘Dynamic Wetting Effects
in Pressure-Sensitive Adhesives,’’ Febr. 25, 2001, Williamsburg, VA, Proc. 24th
Annual Meeting of Adh. Soc., p. 267.
4. Bachofen & Meier AG, Bülach, ‘‘Reverse Gravure Coating of Emulsion PSA’’.
5. C.P. Iovine, S.J. Jer and F. Paulp (National Starch and Chem. Co.,
Bridgewater, USA), EP 0.212.358/04.03.1987, p. 3.
6. S.D. Tobing, O. Andrews, S. Caraway, J. Guo, and A. Chen. ‘‘Shear Stability
of Tackified Acrylic Emulsion PSAs,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr. 25, 2001, Williamsburg, VA, p. 273.
7. Adhes. Age, (3) 41 (1985).
8. D.J. Zimmer and J.S. Murphy, Tappi J., (12) 123 (1989).
9. C. Massa, Coating, (7) 239 (1989).
10. Coating, (11) 316 (1984).
11. Tappi J., (9) 104 (1983).
12. Coating, (4) 123 (1988).
13. G. Auchter, J. Barwich, G. Rehmer and H. Jäger, Adhes. Age, (7) 20 (1994).
14. H. Braun, ‘‘UV Curable Acrylic Based Hot-melt Adhesives,’’ Coating, (12) 477
(1995).
420 Chapter 7

15. ‘‘UV-Curable Acrylic Hot-Melts for PSA Application,’’ BASF Symposium,


2/15/96; in: S. D. Tobing and A. Klein, ‘‘Synthesis and Structure-Property
Studies in Acrylic Pressure-Sensitive Adhesives,’’ Proc. 24th Annual Meeting of
Adh. Soc., Febr. 25, 2001, Williamsburg, VA, p. 131.
16. K.H. Schumacher and T. Sanborn, ‘‘UV-Curable Acrylic Hot-melt for Pres-
sure Sensitive Adhesives-Raising Hotmelts to a New Level of Performance,’’
Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg, VA,
p. 165.
17. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.
18. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 1.
19. U.P. Seidl (Schneider Etiketten u. Selbstklebetechnik GmbH und Co.), OS/DE
3.625.904/04.02.1988; in CAS, 16, (3) (1988), 109, 23972b.
20. I. Benedek, Development and Manufacture of Pressure Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 8.
21. M. Bateson, Labels and Labelling International, (6) 36 (1986).
22. A.W. Norman, Adhes. Age, (4) 36 (1974).
23. R. Jordan, Adhäsion, (1/2) 17 (1972).
24. Verpackungsberater, (5) 108 (1980).
25. H. Muller, J. Türk and W. Druschke, BASF, Ludwigshafen, Germany, EP
01187267/11.02.1983.
26. J. Paris, Papier u. Kunststoffverarb., (9) 53 (1988).
27. W.C. Perkins, Radiation Curing, (8) 8 (1980).
28. R. Hinterwaldner, Coating, (4) 86 (1984).
29. M. Larson, Packaging, (11) 10 (1986).
30. J. M. Casey, Tappi J., (6) 151 (1988).
31. T. Marti and R. Bidie, European Tape and Label Conference, Exxon, Brussels,
April 1989, p. 171.
32. Coating, (2) 58 (1988).
33. L. Salmen, Tappi J., 7 (12) 190 (1988).
34. J. Young, Tappi J., (5) 78 (1988).
35. Coating, (9) 264 (1984).
36. L. Placzek, Coating, (11) 411 (1987).
37. E. Park, Paper Technology, (8) 14 (1989).
38. I. Kesola, European Tape and Label Conference, Exxon, Brussels, April
1989, p. 9.
39. P.E. Patt, Tappi J., (12) 97 (1988).
40. H. Senn, 12th Munich Adhesive and Finishing Seminar, Munich, Germany,
1987, p. 93.
41. Druckprint, (10) 32 (1987).
42. G. Jayme and G. Traser, Das Papier, (10A) 694 (1969).
43. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 396.
44. J.A. Fries, Tappi J., (4) 129 (1988).
Converting Properties of PSAs 421

45. R. Hinterwaldner, Coating, (4) 86 (1984).


46. H.P. Ast, Seifen, Öle, Fette-Wachse, (9) 289 (1985).
47. J. Verseau, Coating, (7) 189 (1971).
48. P. Thorne, Finat News, (3) 150 (1988).
49. W. Graebe, Papier u. Kunststoffverarb., (2) 49 (1985).
50. M. Bowtell, Adhes. Age, (12) 37 (1986).
51. Der Siebdruck, (5) 21 (1988).
52. Allgem. Papier Rundschau, (14) 340 (1987).
53. B. Hunt, Labels and Labelling, (1) 28 (1999).
54. SIP, (9) 40 (1997).
55. SIP, (9) 46 (1997).
56. SIP, (9) 38 (1997).
57. K. Taubert, Adhäsion, (10) 377 (1970).
58. Die Herstellung von Haftklebstoffen, Tl.2.2, 15d, Nov. 1979, BASF,
Ludwigshafen, Germany.
59. Der Siebdruck, (1)14 (1988).
60. A. Zosel, Colloid Polymer Sci., 263, 541 (1985).
61. J.A. Fries, New Developments in PSA, p. 27.
62. R. Lowman, Finat News, (3) 5 (1987).
63. Adhäsion, (1/2) 20 (1987).
64. Adhes. Age, (7) 12 (1986).
65. Adhes. Age, (7) 36 (1986).
66. H.K. Porter Co., USA, US Pat., 3,149,997; in Adhäsion, (2) 79 (1966).
67. Coating, (1)20 (1984).
68. Coating, (1) 89 (1984).
69. P. Holl, Verpackungs-Rundschau, (3) 214 (1986).
70. T. Zepplichal, Adhäsion, (9) 19 (1984).
71. G. Meinel, Papier u. Kunststoffverarb., (10) 26 (1985).
72. Adhäsion, (1) 11 (1974).
73. Adhäsion, (8) 296 (1984).
74. L. Placzek, Coating, (6) 140 (1980).
75. Kunststoffverpackung, (2) 7 (1988).
76. Topics, Courtaulds Films, (1) autumn, 1991.
77. Mobil, OPP Art, (5) 1 (1991).
78. Converting Today, (3) 5 (1981).
79. Converting Today, (11) 7 (1991).
80. P. Hammerschmidt, Allgem. Papier Rundschau, (4) 190 (1986).
81. D.A. Markgraf, Converting and Packaging, (3) 18 (1986).
82. E. Djagarowa, W. Rainow and W.L. Dimitrow, Plaste u. Kaut., (2) 100 (1970);
in Adhäsion, (12) 363 (1970).
83. Adhäsion, (5) 27 (1982).
84. K. Armbruster and M. Osterhold, Kunststoffe, (11) 1241 (1990).
85. W. Brockmann, Adhäsion, (12) 38 (1978).
86. W.H. Kreidl, Kunststoffe, (49) 71 (1959).
87. H. Buchel, Adhäsion, (10) 506 (1966).
422 Chapter 7

88. G. Kühne, Neue Verpackung, (34) 434 (1981).


89. J. Hansmann, Adhäsion, (12) 136 (1979).
90. E. Prinz, Coating, (12) 269 (1979).
91. R. Van Linden, Kunststoffe, (69) 71 (1979).
92. T.J. Bonk and S.J. Thomas (Minnesota Mining and Manuf. Co., St. Paul,
MN, USA), EP12070831/03.10.1984.
93. E. Prinz, Coating, (1) 20 (1978).
94. G. Höfling and H. Breu, Adhäsion, (6) 252 (1966).
95. Coating, (9) 119 (1969).
96. Coating, (9) 296 (1969).
97. E. Prinz, Coating, (2) 56 (1978).
98. P.B. Sherman, Papier u. Kunststoffverarb., (11) 62 (1986).
99. C. Jepsen and N. Rabiger, Kautschuk, Gummi, Kunststoffe, (4) 342 (1988).
100. Converting Magazine, (3) 54 (1986).
101. E. Prinz, Coating, (10) 374 (1988).
102. G. Menges, W. Michaeli, R. Ludwig and K. Scholl, Kunststoffe, (1)
1245 (1990).
103. R. M. Podhajni, Converting and Packaging, (3) 21 (1986).
104. D. Briggs, C. R. Kendall, A. R. Blythe and A. B. Wooton, Polymer, (24)
47 (1983).
105. B. Martens, Coating, (6) 187 (1992).
106. Softal, Report, (102) 2 (1994).
107. E.H. Prinz and K.H. Meyer, Papier u. Kunststoffverarb., (10) 1984; in (7) 505
(1988).
108. H. Grünewald, ‘‘Non-Poluting Plasma Pretreatment for Better Adhesion,’’
15th Munich Adhesives and Finishing Seminar, Munich, Germany, 1990.
109. F.A. Sliemens, Papier u. Kunststoffverarb., (4) 11 (1988).
110. W. Eisby, Allgem. Papier Rundschau, (829) 794 (1988).
111. G. Ellinghorst and C. Müller-Reich, ‘‘Openair Plasma-Fundamentals and
Effects of a New Tool for Surface Treatment at Atmospheric Pressure,’’ Proc.
24th Annual Meeting of Adh. Soc., Feb. 25,2001, Williamsburg, VA, p. 307.
112. P. Swaraj, Papier u. Kunststoffverarbeiter, (1) 52 (1988).
113. G. Liebel, Vorbehandlung von Kunststoff Oberflächen mit Niederdruck
Plasma,’’ Fachtagung, Kunststoffoberflächen des Praxis Forums, Berlin, Bad-
Nauheim, 1977.
114. M.D. Green, F.J. Guild and R.D. Adams, ‘‘Analysis of Functional Surface
Modification of Pre-treated Polypropylene, Using Multi-Modal XPS and
AFM,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg,
VA, p. 254.
115. D.A. Kaute and C. Buske, ‘‘High Performance, Truly Environmentally
Friendly and Cost Effective Bonding Solutions with Atmospheric Pres-
sure Plasma,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001,
Williamsburg,VA, p. 310.
116. S. Böhm, K. Dilger, R. Poprawe and E.W. Kreutz, ‘‘Surface Treatment of
Components by the Use of Lasers to Increase the Wettability and to Improve
Converting Properties of PSAs 423

Adhesion,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001,
Williamsburg, VA, p. 304.
117. Heger, US Pat., 4,041,192; in ref. 92, p. 3.
118. H. Fydelor, US Pat., 4,148,839; in ref. 92, p. 3.
119. Magat, US Pat., 3.252,880; in ref. 92, p. 3.
120. Garnett, US Pat., 4,179,401; in ref. 92, p. 3.
121. S. Yamakawa and F. Yamamoto, J. Appl. Polymer Sci., 25, 31 (1980).
122. S. Shkolnik and D. Behar, J. Appl. Polymer Sci., 27, 2189 (1982).
123. C.L. Aronson, L.G. Beholz, B. Burland and J. Perez, ‘‘Investigation of a New
Mechanism for Rendering High Density Polyethylene Adhesive,’’ Proc. 24th
Annual Meeting of Adh. Soc., Feb. 25, 2001, Williamsburg, VA, p. 294.
124. Adhäsion, (12) 506 (1966).
125. Adhäsion, (10) 6 (1983).
126. D.K. Perino, Paper, Film and Foil Conv., (11) 85 (1969).
127. Coating, (6) 174 (1972).
128. L. Devine and M. Bodnar, Adhes. Age, (5) 35 (1969).
129. L.G. Beholz, US Patents 6,077,913 and 6,100,343,2000, in ref. 123.
130. R. Milker and A. Koch, Coating, (1) 8 (1988).
131. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 3.2.1.
132. Papier u. Kunststoffverarb., (10) 2 (1985).
133. F.T. Kitchel, Tappi J., (2) 156 (1988).
134. Converting Today, (11) 9 (1981).
135. A.M. Slaff, Coating, (7) 198 (1973).
136. J. Patschorke, Adhäsion, (3) 37 (1970).
137. Adhes. Age, (11) 40 (1988).
138. F. Weyres, Coating, (4) 110 (1985).
139. L. Placzek, Coating, (8) 199 (1980).
140. Paper, Film and Foil Conv., (9) 28 (1989).
141. H.E. Kramer, Allgem. Papier Rundschau, (31) 843 (1988).
142. U. Eisele, Kautschuk, Gummi, Kunststoffe, (6) 539 (1987).
143. A. Masaaki, S. Sadaji, N. Hiroshi and O. Tatsuhiko (Nitto Electric Ind. Co.
Ltd.) Japan Pat. 63.142.086/14.06.88; in CAS, 25 (6) (1988).
144. Adhäsion, (5) 76 (1984).
145. Neue Verpackung, (5) 176 (1980).
146. I. Benedek, Development and Manufacture of Pressure Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 3.
147. D. Satas, Adhes. Age, (8) 28 (1988).
148. Der Polygraf, (5) 368 (1986).
149. K. Fust, Coating, (2) 65 (1988).
150. Pack Report, (11) 47 (1983).
151. R. Hummel, 12th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1987, p. 58.
152. Druckwelt, (13) 34 (198).
153. G. Perner, Coating, (5) 150 (1992).
154. Der Siebdruck, (1)14 (1988).
424 Chapter 7

155. M. Schlimmer, Adhäsion, (4) 8 (1987).


156. A. Zimmermann, Verpackungs-Rundschau, (3) 298 (1972).
157. Paper, Film and Foil Conv., (10) 39 (1971).
158. Adhäsion, (1/2) 27 (1987).
159. O. Fellers, L. Salmen and M. Htun, Allgem. Papier Rundschau, (35) 1124
(1986).
160. B. Wright, Adhes. Age, (12) 25 (1971).
161. Coating, (3) 64 (1974).
162. C.W. Drechsler, Coating, (1) 10 (1985).
163. H. Salmen and M. Htun, Allgem.Papier Rundschau, (35) 28 (1986).
164. E. Bohmer, Norsk Skogindustrie, (8) 258 (1968).
165. E. Brada, Papier u. Kunststoffverarb., (5) 170 (1986).
166. A.W. Dobmann and H. Braun, ‘‘Neues Engineeringkonzept für
Etikettenindustrie,’’ 10th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1985.
167. Allgem. Papier Rundschau, (37) 1046 (1988).
168. K. Fukuzawa and T. Uekita, ‘‘New Methods of Evaluation for Pressure
Sensitive Adhesive,’’ Proc. of the 22nd Annual Meeting of the Adhesion Society,
Panama City Beach, FL, Febr. 21, 1999, p. 78.
169. K. Matschke, Hoechst Folien, 5. Klebebandforum, Frankfurt/M, Germany,
Nov., 1990, p. 141.
170. Druck-print, (11) 33 (1986).
171. European Adhesives and Sealants, (9) 46 (1987).
172. J.A. Fries, Tappi J., (4) 129 (1988).
173. Westwald Co., US Pat., 2,140,710; in Coating, (7) 184 (1974).
174. F. Grajewski, A. Limper and G. Schwarter, Kautschuk, Gummi, Kunststoffe,
(12) 1188 (1986).
175. T. Timm, Kautschuk, Gummi, Kunststoffe, (1) 15 (1986).
176. Allgem. Papier Rundschau., (42) 1152 (1988).
177. H. Senn, Papier u. Kunststofjverarb., (4) 470 (1986).
178. G. Renz, Allgem. Papier Rundschau, (24/25) 860 (1986).
179. S.W. Medina and F.W. Distefano, Adhes. Age, (2) 18 (1989).
180. I. Benedek and L.J. Heymans, Pressure-Sensitive Adhesives Technology,
Marcel Dekker, New York, 1997, Chapter 7.
8
Manufacture of
Pressure-Sensitive
Adhesives

The manufacture of PSAs includes the manufacture of the raw materials as


well as their formulation. PSAs include solutions, dispersions, or hot-melt
adhesives. Water-based PSAs can be used in certain cases as such without
any additional formulation. Solvent-based acrylic PSAs also may be used
unformulated. The origin of the PSA manufacture is based on the blending
and dissolution of different raw materials, none of which exhibits a PSA
character by itself. Therefore the role of the formulator is especially
important in the development of PSAs.

1 MANUFACTURE OF PSA RAW MATERIALS

Manufacturing the raw materials refers to the manufacture of the


components of the PSA recipe. There are rubber-like (elastic) and liquid-
like (viscous) PSA components (see Chap. 5). Rubber-resin PSAs are
mixtures of both components. Certain synthetic PSA raw materials (e.g.,
acrylics, vinyl acetate, ethylene, and maleinate copolymers) possess
viscoelastic properties obtained through synthesis. In this case the
manufacture provides a viscoelastic material in the desired physical state
(solid state, solution, or dispersion) through the synthesis. As discussed in
detail in [1,2] the development of in-line manufacture of pressure-sensitive
adhesives imposed the use of monomers, oligomers, and macromers as raw
materials, which are transformed in viscoelastic materials after coating,
mainly by radiation curing. Thus the formulation by mixing of the
components of a PSA (i.e., the ‘‘classical’’ activity of the adhesive
manufacturer) was completed by the formulation of the base polymers

425
426 Chapter 8

(which is a synthesis-related formulation, known from the macromolecular


chemistry).

1.1 Natural Raw Materials


Natural raw materials with rubber-like properties and natural resins or
plasticizers may be used in order to formulate an adequate recipe (see Chap.
5). Of course these natural raw materials are not really manufactured, but
modified chemically or physically. Physical modification allows the use of
these materials in a suitable state (i.e., in solution or dispersion). As
described in [2] in some special cases certain base elastomers are used per se
without tackification.

Rubber
Natural rubber is used as a solution, and manufacture of the PSAs implies
dissolution of the rubber. In fact a slight mechano-chemical destruction of the
rubber occurs in parallel with the dissolution process. At least theoretically a
pre-calendering (i.e., a controlled depolymerization of the natural rubber)
is possible, but not economically viable. There are only a few special uses (like
protective films) where mastication of the rubber must be carried out.
Masticated rubber was used for tapes [3]. Pressure-sensitive adhesives exhibit-
ing low tack, for bonding onto themselves, also are formulated on the basis
of masticated natural rubber [4]. Mastication was discussed in detail in [5].
The high molecular weight portion of natural rubber is insoluble in
solvents and therefore natural rubbers must be milled to a Mooney viscosity
of 53–75 or below, to obtain solubility at 10–30% solids in solvent [6,7].
As shown in Fig. 8.1 and Table 8.1, unmilled natural rubber displays a
broad range of molecular weight (based on viscosity measurements) as a
function of its natural product character; milling reduces the viscosity
(Fig. 8.1) and the variation of the viscosity values (Table 8.1). Milling reduces
the molecular weight and thus the peel adhesion (see Chaps. 5 and 6).

Resins
Resins are used in solution, in dispersions, or as liquid resins. Manufactur-
ing the raw materials means a technical process in order to get the resin
in the desired end-use form. Unfortunately, natural resins possess a special
chemical structure which is oxidation and polymerization sensitive; there-
fore natural resins have to be modified chemically in order to be stabilized.
Generally, disproportionation and hydrogenation improve the aging stabi-
lity. Hydrogenation also increases resin compatibility [8]. Dimerization and
esterification lead to high melting point resins (see Chap. 5).
Manufacture of PSAs 427

Figure 8.1 Variation of the solution viscosity values for a given solid content, for
rubber/resin PSAs on the basis of milled rubber.

Table 8.1 Dispersion of the Viscosity Values


for Unmilled Rubber-Based PSA

Brookfield
Batch Viscosity (mPa.s)

01 11,000
02 20,000
03 17,000
04 19,000
05 14,000
06 12,000
07 16,000
08 10,000
09 18,000
10 13,000
A 1/1 (wet weight/wet weight) rubber/resin blend was
dissolved and diluted to the same solids content.

1.2 Synthetic Raw Materials


Synthetic raw materials for PSAs are mostly viscoelastic materials which
need no or only a low level of viscous formulating additives; hence
428 Chapter 8

manufacturing refers to the synthesis and formulation of a PSA adhesive.


On the other hand, the development of polymer chemistry in the last decade
has made it possible to obtain synthetic viscous components (tackifier resins
or plasticizers). Recent progress in macromolecular chemistry has led to
the synthesis of special (stereoregulated or physically or chemically cross-
linkable) elastomers. Today synthetic raw materials for PSAs are available
as elastomers and viscous components.

Synthetic Elastomers and Polymers


Manufacturing synthetic rubber (as a raw material for PSAs) implies the
synthesis of special rubbers (elastomers) and special viscoelastic materials.
The rubbers possess a special chemical composition or structure. In the class
of new elastomers with a special structure, the stereoregulated diene-based
tactic elastomers (used mainly in rubber-resin solvent-based formulations)
and the crosslinkable, block copolymers (used mainly as raw materials for
hot-melt PSAs) should be mentioned. In this range of materials the halogen-
containing elastomers should be noted also. Among the rubber-like raw
materials with a viscoelastic character there is a broad range of chemically
quite different polymers, such as the hydrocarbon-based carboxylated
styrene-butadiene rubber (CSBR; used mostly in aqueous dispersions), the
acrylic-based homo- and copolymers (used for solvent-based, water-based,
or hot-melt PSAs), the silicone elastomers (used mainly as solutions), and
polyurethanes (see Chap. 5).

Synthetic Resins
In Chap. 5 a detailed description of the resins used for PSA manufacture
was given and the chemical nature of synthetic resins discussed. Details of
the use, the level, and recipes of synthetic resin-based PSAs are given in
Chap. 6. Synthetic resins are mostly based on hydrocarbons, but derivatives
of heterocyclic compounds also are used. Like natural resins, these
compounds also can be used in liquid, solvent, or dispersed state.
Some of the synthetic resins are prepared by polymerizing special
monomers. Such monomers are usually derived from the so-called C9 and
C5 cuts in the fractionation of crude oil or similar materials. Such synthetic
hydrocarbon tackifiers generally have ring-and-ball softening point
temperatures between 10 C and about 100 C. The polycarboxylic acid
ester tackifier resins are polymerized from one or more monomers such as
acrylic acid which may be substituted with alkyl or alkoxyl radicals having
one to four carbon atoms or with alkyl or alkanol esters of such acids in
which the alkyl or alkanol moiety contains from one to about six carbon
Manufacture of PSAs 429

atoms. The most important raw materials are C4-C5 diolefins, styrene,
-methyl styrene, vinyl toluene, dicyclopentadiene, - and -pinene,
d-limone, isoprene, and piperylene. The polymers may be classified as
aliphatic, aromatic, alkyl aromatic, monomer-based, hydrogenated, func-
tionalized hydrocarbon resins, or terpene resins. In comparison with natural
resins, hydrocarbon resins display only a limited compatibility. Aliphatic
resins are suggested for aliphatic polymers or block copolymers where the
aliphatic midblock has to be tackified, such as natural rubber, SBS or SIS
block copolymers. Aromatic resins are used for aromatic or polar polymers,
such as acrylics, styrene-butadiene rubber, or ethylene-vinyl-acetate copoly-
mers. The aging stability of the resin is very important. Resins displaying
good aging stability include aliphatic hydrocarbon resins, aromatic hydro-
carbon resins, dicyclopentadiene resins, and pure monomer-based resins.

Additives
The additives used in PSA formulations are included in order to modify the
adhesive, converting, or end-use properties. Most of them produce changes
in the rheology of the adhesive; others serve to modify the liquid state
properties, the coatability of the adhesive, or the convertability of the PSA
laminate. There is a special class of additives used in order to improve the
chemical stability of the adhesive. Other additives may be selected in order
to improve the antimicrobial resistance of the adhesive. The adhesive raw
materials (i.e., base elastomers and tackifiers) are processed during the
manufacture of PSAs together with other additives (i.e., they are
formulated).

2 FORMULATING PSAs

Formulating means changing the adhesive composition, by addition of


different macro- or micromolecular compounds, having elastomeric or
viscous, surface-active or chemically stabilizing properties, in order to
prepare recipes for PSAs; it is advisable to add antimicrobial agents to the
adhesive. The moderate pH and presence of polyvinyl alcohol makes latexes
sensitive to bacteria.
Pressure-sensitive adhesives should have a Tg of about 15 to 5 C.
Blending a high Tg, hard resin with a lower Tg elastomer provides a desirable
formulating latitude. On the other hand, PSA coating and converting, and
end-use of the PSA label require other special performance characteristics.
These can be achieved only through formulating. The scope of the
formulation of PSAs is to produce tailored adhesives for each one of the
430 Chapter 8

intended end-uses. Generally, the coating technology of the adhesive, the


converting technology of the adhesive, the end-use properties, and
economical considerations influence the formulation of PSAs.
As described in detail in [9] formulation influences the manufacturing
technology (global product technology, adhesive coating technology, and
adhesive manufacturing technology), converting properties of the adhesive
(coatability, drying, running speed, etc.) and the end-use performances
of the pressure-sensitive products. The existing coating and converting
machines, as well as environmental considerations, determine the choice of
an adhesive. End-use properties of the label require special face stock
materials and/or PSAs with certain adhesive properties. Economical
considerations also limit the choice of the laminate components and
formulating components. Generally, PSA formulations contain polymers
(elastomers), tackifiers (resins or plasticizers), fillers, stabilizers, and carrier
agents (solvent, water). This constitutes a base formulation, where all the
components (except the carrier agent) contribute to the adhesive properties.

2.1 Adhesive Properties


There are many reasons for formulating. The first one is the need to improve
the adhesive properties. Rubber-like base materials used for PSAs do not
possess sufficient intrinsic tack and peel adhesion. Most viscoelastic
materials with built-in tack through the synthesis do not possess enough
peel adhesion to nonpolar surfaces, and soft, tacky PSA raw materials
exhibit too low a cohesion level. In other cases the removability of the PSA
must be improved. Therefore the whole range of the adhesive properties
has to be achieved through formulation.

Adhesion/Cohesion Balance
A proper balance between high tack, peel adhesion, and high cohesion is
necessary in most cases. It should be mentioned that formulating for high
tack and peel (discussed generally in terms of tackification), using viscous,
formulating agents (tackifiers and plasticizers) decreases the shear resis-
tance. A high cohesion level is given by the rubber-like component of the
adhesive or by partial crosslinking.
Tack. The main goal of the formulation is to improve the tack and
the peel adhesion of PSAs. Generally, this is achieved by the use of
plasticizers and tackifiers. Tackifying in order to obtain an improved tack is
easier. As can be seen from Table 8.2, a tackifier level of 10–20% is enough
for tack improvement of acrylic-based PSAs. A higher tackifier level is
imposed mainly for the improvement of the peel adhesion, not of the tack.
Manufacture of PSAs

Table 8.2 Tackifier Level Required to Improve the Tack

Formulating Components Chemical Composition, Parts by Weight (wet)

Code Supplier Nature 1 2 3 4 5 6 7 8 9 10 11 12 13 14

3703 Polysar CSBR 100 80 33 — 33 — 43 40 — — — — — —


CF52 A&W AC rosin — 20 23 — 23 — 19 26 — — — — — —
80D BASF AC — — 33 100 43 — — — — — — — — —
V205 BASF AC — — — — — 100 38 34 — — — — — —
Exoryl 2001 Exxon AC — — — — — — — — 100 100 — — — —
SE-351 A&W Ester rosin — — — — — — — — — 20 — 20 — —
5650 Hoechst AC — — — — — — — — — — 100 100 — 20
Nacor-90 National EVA copolymer — — — — — — — — — — — — 100 100
Rolling ball tack (cm) 15.0 3.0 1.5 7.0 2.5 6.0 1.8 2.6 5.9 4.5 3.0 2.8 11.0 8.0
431
432 Chapter 8

This statement appears very important for formulations of PSAs coated on


soft PVC, where the adhesive should contain a minimum resin level in order
to avoid resin/plasticizer interaction.
Trends for the improvement of the tack also are observed. New, wider
converting machines (more than 2 m wide) are being built, giving rise to a
higher variation of the coating weight (12–26 g/m2). New face stock and
substrate materials, based on polyethylene-polypropylene copolymers, are
being introduced with poorer anchorage and peel values. New PSA raw
materials with lower tackifying response (EVAc, CSBR) are now used.
Hence, new film coatings with better tack properties are necessary. More
removable and semipermanent labels (with lower coating weights) are
needed. All these changes require adhesives with improved tack properties
(e.g., more tackifier or more efficient tackifiers). Extreme application
and end-use conditions (e.g., flying labeling, adhesion to low or high
temperature-, humid or tensioned substrates, postapplication printing, etc.)
impose higher tack also.
Peel. High peel adhesion requires a certain tack level for bonding
and a certain cohesion for debonding. Tack increases continuously upon
adding soft, viscous components to the formulation. The dependence of the
peel on the ratio of elastic/viscous components is more complex, going
through a maximum as a function of the level of the soft component. The
same ambiguous influence of the crosslinking on the peel should be noted.
There are adhesive formulations with crosslinking agents leading to low peel
adhesion (i.e., removable PSAs), but in some cases crosslinking leads to high
peel adhesion (see Section 1.2 of Chap. 6).
Shear. The most important means to influence the cohesion are
tackification and crosslinking. Commonly used tackifier levels cause a
decrease in shear strength. The loss of the shear resistance due to the
tackification process is a function of the elastomer and tackifier nature and
their ratio. Figure 8.2 indicates that a tackifier level above 20% decreases
the shear strength of acrylic PSAs.
It is desirable to obtain as much tack as possible without losing a
significant amount of internal strength (cohesion). The better balance of
adhesion and cohesion properties is difficult to obtain through polymeriza-
tion only. The addition of hard monomers to increase the cohesive strength
generally results in a decrease of the tack [10].

2.2 Formulating Opportunities


Formulating by mixing includes tackification, cohesion regulation, detack-
ification, and polymer degradation [2]. Tackification covers blending with
Manufacture of PSAs 433

Figure 8.2 The influence of the tackifier level on the shear resistance. Hot shear
(50 C) dependence on the tackifier resin concentration.

high polymers, resins, and plasticizers. Tackification with high polymers


includes tackification with elastomers, viscoelastomers, viscous macro-
molecular compounds, and plastomers. It is possible to modify the adhesive
properties of a PSA by changing the base rubber-like polymer (i.e., through
blending different elastomers). For rubber-resin formulations this implies
the use of different qualities of natural rubber, or blends of natural rubber
with synthetic elastomers (e.g., stereoregulated ones). For synthetic PSAs,
mixtures of acrylics, or acrylics with EVAc or CSBR can be used. Although
generally used, this formulating freedom is less well known or applied than
blending the rubber-like component with low molecular, viscous materials
(mainly resins), known as tackification.

2.3 Tackification
Tackifying or tackification refers to the increase of the tack and peel adhesion
through formulation: as formulating agents low molecular weight, soft
materials such as resins and plasticizers or special elastomers can be used.
Tackification with Resins. The Tg and modulus of acrylic PSAs can be
adjusted by polymerizing with different ratios of acrylic monomers. Resin
addition offers an additional possibility for this modification [6]. Tackifying
434 Chapter 8

resins have found widespread use in the PSA industry since its inception.
There are practical and theoretical reasons for their use and there exist
technical and commercial reasons which make the use of tackifiers necessary.
It is a fact that the tack and peel adhesion (especially the peel on
nonpolar, untreated surfaces) of PSAs (independent of their chemical basis,
converting technology, and end-uses) do not meet the practical requirements.
Therefore, there is a need for formulating (compounding) the raw adhesives
with chemicals to provide a better tack/peel level. Traditionally these
materials were employed in rubber-based adhesives, where their use is often
necessary in order to achieve a desirable balance of properties. Acrylic PSAs
can be designed in such a way as to be suitable for some applications without
the need for modification. The addition of tackifiers to latex acrylic systems
can, however, widen the performance window available to these materials,
and thus can be a useful tool to complement the existing methods of
performance modification. Several methods of performance modification do
exist: the best known formulation methods include the compounding with
other PSAs and/or adding plasticizers. Unfortunately these methods only
have a limited use. The formulation of PSAs with other PSAs does not
always result in the desired tack/peel balance. Compounding with macro-
molecular plasticizers (like polybutenes) may give rise to compatibility and
cold flow problems. Addition of plasticizers changes the rheology of the
adhesive, involving the decrease of the peel and staining resistance. Thus
the most important method of tackification remains formulating PSAs with
resins. The successful incorporation of tackifier resins can lead to significant
cost savings. Therefore cost reduction also implies the use of tackifiers.
The resin functions as a solid solvent for the polymer. Generally, resin
tackifiers are macromolecular compounds with lower molecular weight and
modulus and a higher Tg than current PSAs (see Chap. 3). Table 8.3 lists
some data concerning the Tg and molecular weight of common tackifier
resins and acrylic PSAs.
Compounding PSAs with the tackifier resin lowers the modulus and
increases the Tg of the blend. The minimum modulus depends on the resin/
rubber solubility. On the other hand, a high resin loading increases the
modulus also. A lowered modulus must promote bond formation (creep
compliance), a higher Tg must make bond rupture difficult (debonding
resistance). It can be seen from Fig. 8.3 that tackification lowers the rubbery
plateau region. Suitable formulations must display a longer and lower
rubbery plateau region, but a rapid (almost vertical) transition (with
increasing strain rate) in the glassy region.
In order to meet these requirements, a good tackifier must have a low
molecular weight, a high Tg, and exhibit good compatibility (see Chap. 3).
In order to evaluate the tack of formulated PSAs, the following examination
Table 8.3 Glass Transition Temperature and Molecular Weight of Common Tackifier Resins and PSAs

Formulating Components
Manufacture of PSAs

Tackifier resin Base polymers

Code Supplier Nature MW Tg ( C) Code Supplier Nature Tg ( C)

Regalrez 1018 Hercules Hydrogenated HC-resin 407 22 80 D BASF AC 53


Regalrez 1033 Hercules Hydrogenated HC-resin 493 7 V205 BASF AC 40
Regalrez 1065 Hercules Hydrogenated HC-resin 723 17 102L BASF AC 58
Regalrez 1078 Hercules Hydrogenated HC-resin 819 29 A120 BASF AC 50
Piccolyte 590 Hercules Terpene resin – 32 S304 BASF SAC 22
Piccotex 75 Hercules Aromatic resin – 29 DH130 Hoechst EVAc 5
Piccotex 9191/92 Exxon C5 aliphatic HC – 46 DH137 Hoechst EVAc 25
Piccotex 323 Exxon Modified C5 HC – 12 AK600 Condea AC VAc 24
Piccotex 9251 Exxon Modified C5 HC – 12 AK610 Condea SAC 12
Piccotex 339 Exxon Cyclic hydrogenated HC – 29 XZ8754 DOW CSBR 43
Piccotex 9271 Exxon Modified C5 – 25 3703 Polysar CSBR 54
Piccotex 348 Exxon C5/C9 HC – 26 3958 Polysar CSBR 31
435
436 Chapter 8

Figure 8.3 The effect of the tackifier on the modulus. The different behavior of the
modulus for 1) a pure and 2) a tackified formulation versus 3) ideal comportment.
Modulus change with increasing temperature/decreasing strain rate.

criteria will be considered, namely the ease of tackifying, the tackifier level,
and the obtained tack. The suitability of PSAs to be tackified (ease of tacki-
fication) should be examined with respect to the chemical nature and the
physical state of the base polymer and of the tackifier. Elastomers with
different chemical bases used as main PSA components have different
tackifying responses (i.e., their compounding at a same level of tackifier
results in different levels of improvement of tack and peel, and a decrease of
the shear). As discussed in Chap. 6, acrylic PSAs display a special, high
tackifying ability compared with EVAc or CSBR. Acrylic hot-melt PSAs
display an excellent compatibility with the same tackifiers and plasticizers as
SIS block copolymers [11]. Acrylic PSAs display good compatibility with
common tackifiers (plasticizers or tackifier resins). They are compatible with
tackifier resins of a different chemical basis (e.g., rosin acid, rosin ester,
hydrocarbon, urea-formaldehyde resins). CSBR shows better compatibility
with hydrocarbon resins; EVAc possesses only a limited compatibility with
all kinds of resins. There is a general compatibility between low tack/high
peel and high tack/low cohesion acrylic PSAs, or between acrylic and
other PSAs (EVAc or CSBR). Compounding of acrylic PSAs with EVAc
yields high shear; compounding with CSBR leads to high tack. The data
of Tables 8.4 and 8.5 demonstrate the adjustment possibilities of the peel
adhesion based on PSA compounding.
Table 8.4 Peel Modification Through Tackification of AC PSAs with Other PSAs
Manufacture of PSAs

Formulating Components Adhesive Properties

1 2 Peel (N/25 mm)


Rolling
Level Level Coating Ball tack
Code Basis (pbw, wet) Code Supplier Basis (pbw, wet) Weight (g/m2) Glass PE (cm)

FC88 UCB AC 100 PC80 UCB AC – 18 8.0 4.0 3.5


FC88 UCB AC – PC80 UCB AC 100 19 6.0 5.0 2.5
FC88 UCB AC 80 PC80 UCB AC 20 30 – 4.0 3.5
FC88 UCB AC 70 2313 Nobel AC 30 23 11.0 5.0 4.5
FC88 UCB AC 50 2313 Nobel AC 50 24 8.0 3.0 7.5
FC88 UCB AC 50 880D BASF AC 50 21 13.0 8.0 7.0
FC88 UCB AC 70 880D BASF AC 30 21 10.0 5.0 6.5
V205 BASF AC 100 80D BASF AC – 19 – 4.0 5.0
V205 BASF AC 80 80D BASF AC 20 20 16.0 PT 9.0 10.0
Note: The samples 1 through 7 are film coatings; the samples 8 and 9 are paper coatings. pbw ¼ parts by weight; PT ¼ paper tear.
437
438 Chapter 8

Table 8.5 Adjustment of the Adhesive Properties via PSA Compounding

Formulating Components Chemical Composition, Parts by Weight (wet)

Code Supplier Base 1 2 3 4 5 6 7 8 9 10 11 12

V205 BASF AC 100 — — 30 40 — — — — — — 64


1360 Hoechst EVAc — 100 — 30 30 50 70 50 25 — — —
FC88 UCB AC — — 100 30 20 — — — — — — —
913 UCB AC — — — — — 50 — — — — — —
1370 Hoechst EVAc — — — — — — 30 — — — — —
85 D BASF AC — — — — — — — 50 100 100 — 16
3703 Polysar CSBR — — — — — — — — — — 100 20

Adhesive Properties

Peel (N/25 mm) from glass — 4.0 8.0 10.0 10.0 — — 11.0 12.0 12.0 10.0 10.0 PT
Peel (N/25 mm) from PE 4.0 2.0 4.0 5.6 5.0 0.2 5.0 — — 8.0 — —
RB tack (cm) 6.0 6.0 3.5 3.5 3.5 7.0 3.0 — — — — 8.0
Coating weight (g/m2) 18 18 18 24 24 21 23 20 21 20 18 —

Note: Samples 1 through 7 are paper coatings; samples 8 through 12 are film labels.

The use of chloroprene latex as a shear modifier is well known.


Neoprene 102 latex is a noncrystallizing low gel dispersion; its low gel
content allows good quick stick and a good holding power when properly
formulated [12]. Tack improvement can be achieved using soft, tacky acrylic
dispersions or polyvinyl ether derivatives [13].
It is difficult to predict the resin level required to achieve a certain tack
or peel value because of the dependence on the nature of the tackified
elastomer and on the nature of the tackifying resin. The nature of the face
stock/adherend should be taken into account as well. A level of up to
30–40% (by weight) tackifying resin is used in formulated acrylic PSAs [14].
Generally, acrylic PSAs require a lower tackifier loading level than CSBR-
and EVAc-based PSAs. Acrylic PSAs need 30–40% tackifier in order to
ensure a high level of polyethylene peel. The use of more than about 25 wt%
tackifier in the total adhesive is common, and more than 50% will lead to
the adhesive being nonpeelable.
The modulus increases at higher resin loadings. A modulus increase
will usually begin in the range of 40–60% resin loading [14]. Hence this is the
upper limit of the tackifier amount. On the other hand, it was mentioned
before that more than 20% tackifier resin leads to lower shear strength. The
minimum/maximum amount of the tackifier depends on the acrylic nature
(chemical composition, molecular weight), the nature of the tackifier, and on
Manufacture of PSAs 439

the converting and end-use properties (face stock, edge bleeding or oozing,
migration, die-cutting, etc.). The tackifier level is a function of the desired
adhesion/cohesion balance, and of the end-use properties of the laminate.
Theoretically, it also is a function of the unformulated PSA and of the
tackifier nature (compatibility). Thus for natural rubber, a two-phase
structure for natural rubber-resin mixtures was proposed [15]. More resin
and low molecular rubber have migrated to the surface. Only a 12% resin-
rich phase was identified optically [16]. In some cases of tackified PSAs, if
the ratio of the energy loss modulus and energy storage modulus (tan ) is
examined by dynamic mechanical analysis (DMA), another peak appears.
This extra peak may be related to a resin-rich phase; there could be some
phase separation between the resin and the rubber so that the rubbery
matrix exists as a two-phase material. When tackifying block copolymers
the compatibility of the additives with both phases must be taken into
account. Aromatic resins (e.g., coumarone-indene) associated with the
polystyrene domains give nontacky materials. Aromatic associating resins
influence the tensile strength less than resins associating with the elastomer.
There are compatibilizers like resin acids.
In hot-melt PSAs the aromatic resins are compatible with the
polystyrene domains, and increase (or decrease) the stiffness depending
on the melting point (>90 C, or <70 C). For better shear and peel
coumarone-indene resins (melting point ¼ 110 C) or -methyl styrene resins
(melting point ¼ 145 C) should be used [17]. The aliphatic resins are
compatible with the rubber segments (see Chaps. 2 and 5). Resins of choice
for CSBR-based adhesive systems are the rosin acid and rosin ester resins
and some of the terpene phenolics. The hydrocarbon resins as a general rule
are not as compatible with CSBR polymers as the rosin acids and rosin
esters. For nitrile and neoprene polymers the best choice remains the
phenolics and modified phenolic resins [18]. Some manufacturers of tackifier
dispersions provide data about the compatibility of the tackifier with PSA
dispersions of different chemical bases. Compatibility is examined according
to the optical appearance (clear, hazy, opaque, separation of components)
of the film [19]. Tackifiers may be included in either minor or major
amounts. The adhesives may contain very small amounts of tackifier, to
increase the tack of the total composition, or they may contain up to 150
parts (by weight) or more of tackifier per 100 parts of one or more of the
polymers. For the same amount of resin addition, a natural rubber-based
system behaves quite differently from one based on SBS block copolymers.
For the same resin loading the two systems would lie in different areas of
the window of application, due to the difference in the loss tan  peak
temperature of the polymers. For SBS multiblock copolymers the midblock
loss tan  peak temperature of the polymer is much lower than that of
440 Chapter 8

natural rubber and CSBR copolymers. At room temperature the storage


modulus value is greater by two decades than that of natural rubber; thus
the concentration of compatible resin in the styrene-butadiene system is
much greater (see Chaps. 2 and 3). This cannot be predicted by the Fox
equation because of the varying compatibility of a resin with the midblock
and the endblock of a block copolymer. A rheology modifier must be added
to reduce the storage modulus value, but not significantly affect the loss
tan  peak temperature; this modifier is usually an oil or a liquid resin [20].
As an example for natural rubber-based PSAs maximum tack is
obtained for 62 parts resin/100 parts rubber [19]. Maximum quick stick is
obtained at 87% resin and optimum peel adhesion is obtained at 100/100
tackifier level [21]. Tackifying natural rubber with a hydrocarbon resin gives
maximum tack at 50–65% resin, maximum quick stick for 85% resin, and
maximum peel at a 100/100 resin/rubber ratio. For polybutadiene 60–150
parts resin per 100 parts rubber should be added [22]. For SIS maximum
tack is given at 50–87% resin and optimum quick stick at 75–125% resin
[19]; a loading level of 150/100 is required for maximum peel. Rolling ball
(RB) tack and quick stick show quite different maximum levels as a function
of the resin concentration [23]. Maximum tack is achieved at lower resin
concentrations (55–70 parts resin/100 parts SIS rubber), quick stick has a
maximum at 95–100 parts resin/100 parts rubber. SIS rubber displays a
maximum rolling ball tack at 40–100 parts hydrocarbon resin/rubber ratio
and maximum peel adhesion is achieved at a 100/100 ratio [21]. As can be
seen from the above examples the tackifier level depends on the elastomer
nature. It should be stressed that a maximum peel level implies more
tackifier than a maximum tack level. Tack for acrylic hot-melt PSAs goes
through a maximum at 45–60 parts tackifier with properties also dependent
on the tackifier composition [24]. In experiments with ethylene/maleinate
copolymers, a 30/100 resin/elastomer ratio was used [25]. In tackifying
studies of acrylic latices using rosin derivatives 0–70 parts resin/100 parts
latex were used [5].
In a series of experiments testing electron beam-cured PSAs, the resins
and plasticizers were screened at a simple 50/50 polymer/resin ratio [26].
Generally, a lower resin/elastomer screening ratio is used for acrylic PSAs
(solvent-based or water-based), for water-based PSAs a 30–40% tackifier
level is suggested. Concerning the evaluation of different tackifiers for
rubber-resin formulations the rubber quality should also be specified. As an
example crêpe with a Mooney viscosity of 55 was used for the screening of
tall oil resins [27].
The peel of tackified CSBR adhesives increases with increasing
tackifier content and attains a maximum at a certain tackifier content [28].
A similar behavior may be observed for PSAs on other chemical bases.
Manufacture of PSAs 441

Table 8.6 Shear Dependence on the Adhesive/Tackifier Nature/Level

Formulating Components

PSA Tackifier Chemical Composition, Parts by Weight (wet)

Code Supplier Nature Code Supplier 1 2 3 4 5 6 7 8 9

3703 Polysar CSBR — — 69 52 37 36 43 38 24 — —


3958 Polysar CSBR — — — — — — — — 10 37 —
6157 Polysar AC — — — — — — — — — — 50
80 D BASF AC — — — — 34 — — 32 — —
V205 BASF AC — — — — — 16 39 33 — 32 —
— — — CF52 A&W 31 48 32 48 49 38 33 — 50
— — — CF301 A&W — — — — — — — 33 —
Hot shear (70 C) min — 368 — 270 136 50 — — 16
Hot shear (70 C) hr 22 — 29 — — — >22 >24 —

Generally, shear values decrease with increasing tackifier level, but in some
special cases (e.g., acrylic hot-melt PSAs) the cohesive strength increases
with the tackifier level. Table 8.6 illustrates the variation of the shear as a
function of the adhesive/tackifier nature and level.
The relatively high tackifier level required for high peel adhesion is
explained by its influence on the modulus (modulus increase begins in the
range of 40–60%). The resin level required for PSAs is generally higher than
that for ‘‘classical’’ adhesives. A difference between the current state and
future trends concerning the tackifier loading has to be made. Today the
best overall balance of performance properties is obtained using moderate
(e.g., about 30%) levels of high softening point tackifiers. Practically,
concentrations between 30 and 40% of resins with a 50–70 C softening
temperature are used. As an example for an acrylic a 30 parts (dry) tackifier
level of acid rosin is suggested. Unfortunately the shear level of these
formulations is not sufficient to give good die-cuttability with normal (not
voluminous or multilayer) face stock and release paper. There are, however,
trends concerning the improvement of the cohesion.
In the future the first possibility to improve the shear strength of the
tackified dispersion is a decreased tackifier loading level. For this purpose
one needs high tack dispersions which require only a minor amount of
tackifier in order to obtain a high tack/peel level. Table 8.7 illustrates the
range of some recent developments in this field; it can be seen, that special
dispersions need no more than 20% tackifier. The second possibility is to use
special tackifier-friendly acrylic-based PSAs which can absorb more than
40–50% tackifier, without the loss of the shear. Table 8.8 illustrates some
442 Chapter 8

Table 8.7 Tackifying with a Low Resin Loading

Adhesive Characteristics

Tackified PSA Peel at RT Shear at


Dispersion (N/25 mm) 50 C (min)

1090 (UCB) 20.0 PT 100


5650 (Hoechst) 16.5 70
2001 (Exxon) 4.5 60
85 D (BASF) 16.8 30
80 D (BASF) 6.5 90
20094 (Syntho) 0.2 120
A blend of 100/20 PSA/Tackifier (wet/wet) was used. Tackifier: SE
351 (Akzo-Nobel). Face stock: paper; RT, room temperature.

Table 8.8 Tackification of Tackifier-Friendly Acrylic PSAs

Formulating Components Adhesive Characteristics

AC-PSA WB tackifier resin


RB tack 180 peel RT shear
Code Level Code Level (cm) (N/25 mm) (min)

80 D 72
V 205 38 324 40 10 12 30
V 205 70 8/21 30 15 13 30
80 D 60 8/21 40 13 9 200
2395 60 TX 60 8 13 460
2395 60 124T 60 8 9 600
2395 60 301CF 60 9 12 34
1304 60 SE80CF 40 14 16 90
620 60 SE80CF 40 9 15 14
AC, acrylic; WB, water-based; RB, rolling ball; RT, room temperature.

new developments in this field. Although both categories of products


possess higher shear, the practical shear/die-cutting tests do not lead to
superior results compared to current products. The obtained tack level
depends on the base elastomer (chemical and macromolecular character-
istics), on its physical state, on the tackifying agent (chemical and other
characteristics), and on the tackifier level and technology. Generally, the
obtained tack level should be examined for the same kind of PSAs.
A detailed analysis of the obtained tack level is given in Chap. 6, including
Manufacture of PSAs 443

a comparative evaluation of the adhesive performance characteristics for


PSAs on a different chemical basis.

Choice of Tackifier. To modify properties many variations on the


starting formulation are possible, namely the ratio of hard to soft resin, a
change of the softening point of the hard or soft resin, and change in the pH
of the compound. Generally, tackifiers are employed to impart or control
one or more of the following properties in PSAs: tack, peel strength,
cohesive strength, staining, migration or adhesive bleedthrough, bleeding
or oozing, stringing or legging, and aging characteristics. Their use also
influences the most important coating characteristics (e.g., the rheology,
foam generation, substrate wet-out, and mechanical stability). Generally,
the most important criteria for the choice of tackifiers include the nature
of the base PSA (to be tackified), the balance of the desired adhesive,
converting and end-use properties, the tackifying technology, and econom-
ical considerations.

Compatibility of Resin. Compatibility is a key parameter for the


mixing of formulation components. The term compatibility is used of
polymers for their dispersing at molecular level. The main aspects of
compatibility and its influence on the formulation were discussed in [2,29],
in correlation with the theoretical basis of the tackification with resins.
Compatibility is a general requirement for optimum tackification [30] (see
Chaps. 2 and 3). The resin functions as a solid solvent for a portion of
the polymer, and the resin and the polymer must display similar solubility
characteristics if the resin is to act as a tackifier (mutual compatibility).
Several papers were published on the effect of the resin on PSA properties
and on the compatibility of resins with elastomers, especially in combination
with block copolymers [31,32]. Hydrogenated hydrocarbon resins (e.g.,
Regalrez, Hercules) may be selected according to their cloud points [33].
Cloud point measurements reveal aromatics. As discussed in Chap. 2, for a
first test of compatibility the Tg is used. For hot-melt PSAs the position of
the Tg shows whether the resin is compatible with the end- and midblocks
of the polymer. The tackification of the midblock influences the modulus,
elongation, and tack of the adhesive. Generally, tackifying of the aromatic
domain is suggested for contact adhesives, for PSAs tackification of the
midblock is required. Polystyrene displays a limited compatibility with
natural rubber only [34]; it is adequate for butadiene-styrene copolymers.
Poly(vinylcyclohexane) possesses good solubility in natural rubber, but it
cannot be dissolved in butadiene-styrene copolymers. Poly(butylstyrene)
may be used in natural rubber- and butadiene-styrene-based PSAs.
Traditionally most hot-melt PSA systems are based on SIS block
444 Chapter 8

copolymers. They are usually easier to tackify than SBS materials. Their
better compatibility with tackifying resins and greater elasticity have
ensured SIS block copolymers a larger part of the hot-melt PSA market.
In recent years more attention was paid to the tackification of SBS-based
block copolymers. These polymeric materials are cheaper than SIS
copolymers, but nevertheless offer interesting performance characteristics.
A clear color and good aging resistance for the tackifier resins are needed.
SBS does not exhibit enough elasticity, flexibility, or tackifier compatibility.
Tackifiers for these rubbers must display good compatibility, a light color,
good aging stability, and a satisfactory cohesion/tack balance. The most
important resins for tackifying SBS block copolymers are Escorez ECR-366,
ECR-368, and ECR-369 (Exxon) [35], where ECR-366 is a modified C5
aliphatic hydrocarbon resin.
As discussed in detail in [2] resin compatibility can be predicted using
the solubility parameter concept. Compatible tackifier works like a solvent.
The incompatible tackifier works like a filler. Tackification supposes mutual
solubility [31,36–38] and is achieved if on the mutual solubility diagram of
the elastomer and tackifier the solubility domain of the elastomer is enclosed
within the domain of the resin [39]. This does not happen for hydrocarbon
resins and rubber; however, as seen earlier, hydrocarbon resins are adequate
tackifiers for natural rubber. In the studies of Schlademan and Bryson
[40,41] the specific gravitiy of the resin was used to estimate its role as a
compatibilizer; the specific gravity was found to be a rough measure of
the solubility parameter (); later the refractive index was chosen. As stated,
the polarity, the molecular weight, and the structure of the macromolecular
compounds play a decisive role in their compatibility. The influence of the
tackifier on the modulus is another index of compatibility [2]. Tackifier
compatibility can be examined by DMA measurements. The resin molecular
weight controls the loss/storage peak (tan ). A compatible system exhibits a
maximum effect in lowering G0 during bonding.
Compatibility also depends on the molecular weight and MWD;
a narrow MWD indicates better compatibility. As an example Regalrez
resins have a MWD characterized by a ratio Mn/Mw ¼ 1.1  1.4 [33]. In a
similar manner for classical hot-melt adhesives the compatibility of the
waxes depends on their MWD and melting point, and influences the
bleeding [35]. Block copolymer acrylic-based PSAs display very good
compatibility with hydrogenated rosin esters and C5/C9 hydrocarbon-
based resins [24]. Generally, compatibility is a function of the melting point
of the resin and its composition. Regalrez as an example shows good
compatibility with acrylic-based PSAs up to a ring-and-ball softening point
of 78 C. Thus, on the basis of mutual compatibility, there are general
Manufacture of PSAs 445

recommendations concerning the use of certain tackifiers for different


classes of latexes [7]:

For natural rubber latexes, petroleum-based aliphatics, petroleum


aromatics, terpenes, and rosin esters are proposed.
For CSBR, rosin esters, petroleum-based aromatics, low molecular
weight, pure monomer aromatics, -pinene, and terpenes are
suggested.
For acrylic PSAs, rosin esters, pure monomer aromatics, low
molecular weight polystyrene, and copolymers of -methyl-
styrene-vinyl toluene are proposed.
Hydrocarbon-based resins display lower overall compatibility than
natural products. Aromatic resins are proposed for acrylic-, CSBR-, and
EVAc-based PSAs [42]. For example Alresen PT 214 is proposed by Hoechst
to be used as tackifier for Mowilith LDM 1360 EVAc. Because the main
criterion for the choice of a tackifier is its compatibility with a PSA, it is
possible to cover the needs of different PSAs with a limited range of tackifiers.
For example DRT (France) is supplying a disproportionated rosin, a terpene
phenol resin, and a stabilized rosin ester for chemically very different PSAs
[20]. There is a fundamental difference with regard to the choice of tackifiers
for PSAs and other adhesives. Bonding of PSAs is (mainly) a physical
process. No chemical interaction is needed between tackifier and/or
elastomer and substrate. Thus the chemical structure of the resin is less
important than its compatibility and its influence on the adhesion/cohesion
balance. The molecular weight should be less than 1000 and the resin should
have a narrow MWD (e.g., Snowtack CF 301, from A&W) [30].
For higher tack values UV-curable acrylic PSAs should be mixed with
non-UV absorbing resins, like Regalite R-9100 [43]. Electron beam-
crosslinkable SIS rubber develops a crosslinked network with some resins
(i.e., higher gel content) at a lower electron beam dose than with other
electron beam-cured PSAs. The average gel content for the adhesives using
the saturated resin was 0.7% at 3 Mrad, while the average for those using
the unsaturated resin was only 0.15%. Rolling ball tack values of PSAs
prepared with the saturated resins stayed at about the same level after
irradiation; rolling ball tack of PSAs using unsaturated resins increased
from 4.7 to 6 cm [26]. While there is a major difference between the adhesives
based on saturated hydrocarbon resins and those based on unsaturated
resins, there also are differences within these classes. For the adhesives with
unsaturated resins these differences can be large. Therefore the selection of
tackifier resins is of prime importance in obtaining the maximum amount of
compatibility with aliphatics and incompatibility with polymer crosslinking
446 Chapter 8

with a minimum of electron beam radiation. By EB curing of HMPSAs the


dosage increases thus increasing the aromatic content of resin. (The type
of plasticizer used may have a similar effect on the radiation-induced
crosslinking.) For electron beam-cured hot-melt PSAs based on Kraton D
1320 X (a branched styrene-butadiene copolymer with functionalized end
groups) hydrocarbon resins of the Escorez series, mainly the E-1310,
E-1401, E-5300, and E-5380 types were used. The details of the tackification
for postcuring are described in [44].
Hydrocarbon resins together with terpene-phenolic resins also are
recommended for polyvinyl acetate dispersions. A narrow MWD imparts
better compatibility also [45]. For rubber-resin PSAs -pinene derivatives
and for acrylic PSAs hydrocarbon-based resins were suggested as tackifiers
at the early stages of PSA manufacturing [46]. Hydrocarbon resins (like
Zonarez B115) are recommended for natural rubber-based or SIS-based
PSAs (e.g., Zonarez Z115). In both cases maximum tack is achieved at resin/
rubber ratios exceeding 50/100; maximum peel is obtained at resin/rubber
ratios above 1.
The major factors in determining the type of resin to use with a par-
ticular latex are its composition (chemical structure), and molecular weight.
Resins of appropriate structure are compatible, but only to some limited
concentration. Acrylic PSAs show the best tackifying ability, with a broad
compatibility with most tackifying resins, independently from their composi-
tion. Styrene-butadiene rubber, for example, with a higher styrene content
does not respond well to resin addition. EVAc shows lower compatibility
with tackifiers than acrylic- or CSBR-based PSAs. One can conclude that
the types of tackifiers used with acrylate copolymers and styrene-butadiene
rubbers are generally useful with the new EVAc-based polymers as well.
Somewhat lower softening points than normal may be desirable depending
upon the particular EVAc. Suitable tackifiers for EVAc dispersions are
polyvinylether, colophonium resins, low molecular resins, and plasticizers.
As an example a level of 30% polyvinylether for 70% MV 704 dispersion
(Hoechst) is recommended. As indicated earlier, hydrogenated hydrocarbon
resins display better compatibility towards EVAc with low vinyl acetate
content (< 40%). A narrow MWD also implies a better compatibility [47].
Taking into account the different factors influencing the compatibility
of the tackifier resin with the other components of the adhesive recipe,
it is evident, that the base elastomer determines decisively the choice of
the tackifier. As discussed in detail in [2] the tackifying versatility of various
base elastomers strongly differs. For solvent-based adhesives the versatility
range can be listed as follows:

NR >> CNR > APP > BR > SIS >> SBS > SBR ð8:1Þ
Manufacture of PSAs 447

For water-based PSAs the versatility range is given as follows:


AC > CSBR > NR > CNR > BR > EVAc > SBR ð8:2Þ
Depending on the nature of the rubber (natural or synthetic block
copolymer) a principal difference exists in the effects of the tackification.
As described in detail in [2] tackification of natural rubber leads to
simultaneous tack and peel increase, tackification covers the whole polymer
and reduces the shear resistance. Tackification of TPEs produces no
simultaneous tack and peel increase, does not cover the whole polymer, and
may increase the shear resistance.
Balance of Adhesive Properties, as a Function of the Resin Nature and
Softening Point. The addition of resin lowers the modulus of the rubber
and increases the Tg [48,49]; accordingly a suitable tackifying resin should
produce these changes. The lowering effect of the modulus and the increase
of the Tg are dependent on the resin’s own characteristics (E and Tg) and
its compatibility. In a simplified approximation for resin/elastomer systems
with good compatibility, the most important parameter is the softening
domain (melting point) of the resin (also related to E and Tg). In general,
soft resins impart aggressive grab and quick stick, while harder resins help
retain good cohesive strength and creep resistance; resins with a low
softening point impart tack, but give poor cohesive strength, while those
with a high softening point give good adhesive strength but poor tack. For
example, glycerine esters are tackier than pentaerythrite esters. On the other
hand, compatibility is a function of the molecular weight and softening
point of the resin. A higher softening point resin is less miscible with the
base polymer [50]. Typical rosins and hydrogenated rosin ester tackifiers
have ring-and-ball softening temperatures of about 25 C to about 115 C,
while preferred tackifiers have softening temperatures of about 50 C to
about 110 C. Table 8.9 illustrates the dependence of the adhesion/cohesion
balance on the characteristics (softening temperature and chemical nature)
of the tackifying resin.
High softening point resins and ester resins impart less tack (and peel)
than low softening point acid resins. The lower acid level increases the
sensitivity of tack and shear to molecular weight. In the tackification of
natural rubber with hydrogenated rosin esters, a low softening point resin
(75 C) gives maximum tack at higher resin concentration (110 parts resin/
100 parts rubber) than a high softening point resin (95 C), which displays a
maximum tack at 60 parts resin/100 parts rubber ratio [51]. Evidently, the
absolute value of the final tack obtained with the low softening point resin is
higher. In a manner quite different from tackifying SIS with hydrocarbon
resins, in some cases a maximum value of the peel adhesion was observed as
Table 8.9 The Influence of the Tackifier Resin Nature and Softening Point Range on the Adhesive Characteristics of WB AC
448

PSAs

Formulating Components Adhesive Characteristics

Tackifier resin dispersion Acrylic PSA


Coating 180 peel
Softening Chemical Level Level Weight Rolling ball (from PE, Shear at
Code Supplier point ( C) Basis (wet) Code (wet) (g/m2) tack (cm) N/25 mm) 50 C (min)

1 CF 52 A&W 55–65 Rosin acid 21.4 V205 35.7 21 13 25 PT 102


H5786 Hercules 95–105 Rosin ester 21.4 80 D 21.4
2 H5786 Hercules 95–105 Rosin ester 42.8 V205 35.7 16 > 30 14 PT 98
80D 21.4
3 MB698 Hercules Liquid Rosin ester 7.2 V205 35.7 21 5 20 PT 250
4576 Akzo 95–105 Rosin ester 35.7 80 D 21.4
4 MB Hercules 55–65 Rosin ester 42.8 V205 35.7 16 2 20 PT 37
6152 80 D 21.4
5 511 DRT 55–65 AC-rosin 48.4 V205 19.3 19 4 25 PT 55
80 D 32.2
6 DEG DRT 35–45 Rosin ester 48.4 V205 35.7 20 2.5 23 PT 40
80 D 21.4
7 DEG DRT 35–45 Rosin ester 14.3 V205 35.7 21 5 20 PT 47
80 DRT 60–80 Rosin ester 28.6 80 D 21.4
8 9251 Exxon – HC 48.4 V205 35.7 18 1.5 25 PT 15
80 D 21.4
9 9251 Exxon – HC 33.4 V205 35.7 19 2.5 20 PT 25
9271 Exxon – HC 15 80 D 21.4
10 OU80 Oulu – – 33.4 V205 35.7 20 7 17 PT 20
11 OU30 Oulu Liquid AC-rosin 15.0 80 D 21.4
Chapter 8

CF 52 A&W 55–65 48.4 V205 35.7 18 2.5 18 5


80 D 21.4
Manufacture of PSAs 449

a function of the resin concentration, which increases continuously with


increasing resin concentration when tackifying natural rubber with resin.
Resins with a softening point of about 50 C impart tack but give poor
cohesive strength, while those above about 70 C give good cohesive strength
but poor tack. To modify the properties the following variations on the
starting formulation are possible, namely the ratio change of the hard to soft
resin or a softening point change.
The addition of higher Tg polymers and/or the addition of higher
softening point tackifiers enhances the convertability and provides specific
converting properties [52]. Increased thermal or pressure sensitivity can be
obtained by increasing the level of tackifier and/or changing to a softer,
more heat sensitive tackifier. This may cause a reduction of the high
temperature bond strength. The tackiness of a high acrylate copolymer
can be increased by using a lower molecular weight (lower softening
temperature) tackifier. Tables 8.10 and 8.11 illustrate the influence of the
low softening temperature resins on the tack and shear. It can be seen that
chemically different resins (polar melamine derivatives, nonpolar hydro-
carbons) cause a similar shear decrease. On the other hand, it should be
noted that acid and ester rosins with the same softening point interval,
impart different adhesion/cohesion levels (Table 8.12).
The influence of the chemical nature and melting point of the tackifier
resin should always be examined with respect to the required adhesive
property. For properties where bonding affinity is more important a change
of the chemical nature is preferred. For mechanical properties of the bulk
polymer (inside the adhesive layer) a change of the melting point should be
considered preferentially. As an example, different tackifiers (stabilized rosin
acids, disproportionated rosin acids, hydrocarbon resins, rosin esters,
hydrogenated rosin esters) were tested for CSBR [53]. The best performance
characteristics concerning the peel were obtained with a stabilized rosin acid
(softening point 52 C). The best shear resistance is given by a high melting
point hydrocarbon resin (65 C) and the best loop tack is obtained when
using stabilized rosin acids. The best polytack is given for disproportionated
rosin acids. The complex dependence of the adhesion/cohesion balance for a
neoprene latex on the melting point and chemical nature was described [12].
The dependence of the shear (SAFT) of a rubber/hydrogenated resin
formulation is illustrated in Fig. 8.4. As can be seen from the figure the shear
resistance increases with increasing softening point of the resin. In order to
improve simultaneously the chemical nature and melting point dependent
properties, mixtures of resins with different melting points and on different
chemical bases (liquid resins) are suggested. Resin mixtures with different
melting points (1 to 1 blends of resins with melting points of 32 C and 64 C,
and 26 C and 82 C) were suggested for alkylene-maleinate copolymers [25].
450

Table 8.10 The Influence of Low Softening Temperature Resins on the Adhesive Characteristics of PSAs

Formulating Components

Tackifier resin Adhesive Characteristics


Coating
Softening point Level Weight Rolling ball 180 peel from Shear at
Product/Code Supplier ( C) Basis (parts, wet) (g/m2) tack (cm) PE (N/25 mm) 50 C (min)

Polyethyleneimine BASF RT Polyethyleneimine 20 17.0 3.5 12.0 15.0


Hercolyn D Hercules RT HC/ester 20 21.0 4.0 16.0 15.0
CF301 A&W 60–65 Acid rosin 100 19.0 6.0 10.0 200.0
Cymel Cyanamide RT Melamine 20 20.0 3.0 10.0 20.0
CF 52 A&W 55–65 Acid rosin 100 18.0 2.4 14.0 100.0
Base PSA: CSBR dispersion (100 parts).
RT, room temperature; HC, hydrocarbon.
Chapter 8
Table 8.11 The Influence of the High Cohesion AC Components and Liquid Tackifier Resins on the Adhesive

Formulating Components

Tackifier resin Base acrylic PSA Adhesive Characteristics

Softening Level Coating Rolling


point range Level (parts, Weight ball tack Peel Shear at
Manufacture of PSAs

Code Supplier ( C) (parts, wet) Code Supplier wet) (g/m2) (cm) (N/25 mm) 50 (min C)

CF 52 A&W 55–65 33 V 205 BASF 27 20 3.0 19 7


80 D BASF 20
CF 52 A&W 55–65 33 V205 BASF 47 19 3.0 6.0 10
80 D BASF 20
CF 52 A&W 55–65 33 FC88 UCB 5 20 6.5 19 18
20094 Syntho 5
V205 BASF 47
80 D BASF 20
FC88 UCB 10
CF 52 A&W 55–65 33 20094 Syntho 10 20 6.5 15 PT 20
V205 BASF 47
80 D BASF 20
CF 52 A&W 55–65 33 20094 Syntho 10 20 9.0 14 PT 20
V205 BASF 47
80 D BASF 20
CF 52 A&W 55–65 33 20094 Syntho 20 22 1.4 14 PT 19
V 205 BASF 47
Hercolyn D Hercules RT 10 80 D BASF 20
RT, room temperature.
451
452

Table 8.12 The Influence of the Resin Nature on the Adhesive Characteristics

Formulating Components

Rosin acid Rosin ester Acrylic PSA Adhesive Characteristics

Coating Shear at
Formulation Weight Rolling 180 peel 50 C
Code Supplier Code Supplier Code Supplier (parts, wet) (g/m2) ball (cm) (N/25 mm) (min)

CF 52 A&W 53251 A&W V 205 BASF 35.7 21 5.0 20 PT 25


80 D BASF 21.4
Chapter 8
Manufacture of PSAs 453

Figure 8.4 The influence of the softening point of the tackifier resin on the holding
power of a PSA formulation. The increase of the holding power with the increase of
the softening point of the resin.

A detailed discussion of the choice of tackifier depending on its nature and


softening point will be given later (see Section 2.8).

Influence of Physical State of the Tackifier. Both solvent-based and


water-based tackifying resins can be added to water-based acrylic PSAs.
Generally, it is not possible to add molten tackifiers to water-based acrylic
PSAs, but there are some special halogenated copolymer dispersions with
good thermal resistance. Relatively low melting point resins (95 C) may be
added to aqueous adhesive dispersions (polyvinyl ethers) in a molten state.
Some liquid resins can be added as such (without emulsifying) to compound
a latex (e.g., Cymel 301 in neoprene latex) using high speed agitation;
because of the good mechanical and chemical stability of the latex, there is
no coagulation problem. In HMPSA formulations liquid resins can replace
processing oils also. The use of liquid resins as tackifiers also depends on the
properties of the face stock material. Ordinary low cost papers without
an over-layer or special thermal papers are more sensitive to adhesive
bleedthrough. The use of liquid resins or a high surface-active agent
concentration in the tackifier must therefore be avoided.
Different acrylic latexes were tackified with polybutene and polybutene
emulsions [54]. From comparative testing in both systems it is anticipated
454 Chapter 8

that the tack and adhesion results are similar, independent of the system
used. On the other hand, the shear strength for carrier-free resin tackified
acrylic PSAs is higher than with resin emulsions [14]. In formulating practice,
the use of tackifying resin dispersions is more widespread because of
technological advantages [55]. In the early stages of PSA formulation,
rubber-resin PSAs were tackified with resin solutions. The first tackified
acrylic formulations also used tackifier solutions. Later tackifier dispersions
in combination with water-based PSAs were developed. Solvents may be
used as an additive (10–20%) in latexes [56,57]. In a similar manner, the
solids content of latexes may be improved with rubber solutions (20%) [56].
Kajiyama [58], showed that phase structures of PSAs prepared from
solution and emulsion differ, and prepolymerization tackification reduces
the molecular weight of the base viscoelastomers.

Tackification of Water-Based PSAs. The same tackifying resins used


for solvent-based PSAs also are envisaged for water-based PSAs. Theore-
tically, there are three different ways for the addition of tackifier resins into
water-based systems, i.e., as resin solutions, resin dispersions, and as a bulky,
molten material. From a technical point of view the most convenient method
is the use of resin dispersions. From the point of view of the adhesive
properties, solution or molten resin feed-in seems to be better because there
are no water-sensitive additives in the resin. Practically, resin solutions only
have a limited use (for PSA coating onto film); molten resin may be added in
heat-resistant dispersions (like chloroprene) only. Resin-tackified water-
based PSAs acquire a composite character with limited compatibility of the
emulsion-borne components; thus the tackifying effect is less effective than in
the case of solvent-based systems. On the other hand, resin fluidity in the
adhesive layer is enhanced by its composite structure, therefore strike-
through (penetration) is more prevalent in water-based systems and the shear
resistance of water-based PSA-tackifier systems reduced even much more
[59]. Water-soluble chemicals from the paper or film also may negatively
influence the aging properties of the resin. Because of the composite structure
of the resin particles, the heat resistance of the resin dispersions may differ
from that of the bulk material, influencing the storage and converting
properties of the laminate. In summary, the most important factors
influencing the choice of tackifiers for water-based systems are the limited
number of resins available as an emulsion, and the limited allowable loading
level of liquid low cohesion resins, because of the decrease of converting
properties and heat resistance, and increase of the penetration (adhesive
bleedthrough), and the limited number of resins with good aging stability.
A detailed study of the tackification of water-based dispersions with
rosin-type tackifier dispersions with a broad range of Tg and softening point
Manufacture of PSAs 455

(16 C to þ45 C and 25–90 C, respectively) and molecular weight between


720–1150 was done by Wood [60]. Wood stated that suitable tackifier resins
have to display low molecular weight, high Tg, and good compatibility.
There are some features common to the modification of rubber-resin-based
and water-based acrylic PSAs. The addition of certain resins reduces rolling
ball tack in direct proportion to the softening point of the resin and the
concentration employed, except for the 50% level of the 90 C softening
point resin [60]. On the other hand, the best overall balance of performance
properties is obtained using moderate (30%) levels of high softening point
tackifiers. A similar influence of Tg, molecular weight, and MWD on the
tackification of hot-melt PSAs (SIS triblock copolymers) was demonstrated
by Tse and McElrath [61]. They showed that if narrow MWD resins are
blended with SIS, the position of loss tan  on the temperature scale
corresponds to the molecular weight of the resin. This is because the resin Tg
is controlled by its molecular weight (see Chaps. 2 and 3). Film coatings
require high transparency and no yellowing during end-use. For this
purpose hydrocarbon resins are preferred.

Tackifiers and Converting Properties. As discussed in [2] the formu-


lation of the adhesive with tackifier resins may be carried out for multiple
purposes (regulation of the technological or adhesive properties). For PSA
formulations the main scope concerns the regulation of the adhesion/
cohesion balance. Generally resins influence the following characteristics of
the formulation: adhesion, peel resistance, shear resistance, temperature and
chemical resistance, gloss, transparency and barrier properties. Resins
generally decrease the melt viscosity of polymers and affect their flexibility.
Tackification with resins may be a technological necessity also (e.g.,
tackification for processing). Tackifiers affect the converting performances
of the adhesive (e.g., coatability) and of the pressure-sensitive product (i.e.,
confectionating, printing, and labeling). Good converting properties of the
aqueous dispersion (coating properties) and of the laminate (processing) are
essential (i.e., an adequate rheology, low foam generation, good substrate
wet-out, excellent mechanical stability, and high converting speed or drying).
Aqueous dispersions exhibit non-Newtonian shear thinning characteristics
and variations in the viscosity recovery. This difference may produce an
undesirable ‘‘corduroy’’ effect which is termed ribbing or striation. Acid
resin-tackified acrylic-based PSAs show better flow properties. Generally,
tackified dispersions display a less adequate rheology.
Tackified aqueous dispersions contain more surface-active agents than
untackified ones; hence they tend to generate more foam than pure PSAs.
The wet-out behavior of tackified water-based PSAs is generally a function
of the base acrylic (CSBR or EVAc) and of the tackifying resin. For the
456 Chapter 8

Table 8.13 Wetting Ability of Different Aqueous Tackifier Dispersions

Water-Based Tackifier

Code Supplier Wetting Agent (%) Wetting Ability

MBG-152 Hercules 3 Very low


DEG/G DRT 1 Average
MBG-64 Hercules 2.5 Low
50 D Oulu 3 Very low
EH-4576 Akzo 4 Very low
E-9251 Exxon 1 Good
E-9271 Exxon 1 Good
CF 52 A&W 1 Good
SE 351 A&W 1.5 Low
CF 8121 A&W 2 Low
The wetting ability was tested on solventless silicone-coated release paper at a
viscosity of 400 mPa.s; Lumithen IRA (BASF) was added as wetting agent.

same base, for PSAs using different tackifiers and sulfosuccinates as wetting
agents, the wetting out on the siliconized release paper depends on the pH;
at higher acidity one needs less wetting agent for an adequate wet-out.
Generally, the surface tension and the wetting out of tackifier disper-
sions from different suppliers vary quite a bit (as a function of the experience
and knowledge of dispersing technology). Table 8.13 illustrates the wetting
characteristics of different tackifier dispersions. Wetting ability was evaluated
by the amount of added surface agent necessary for good wet-out.
An additional deficiency of tackified water-based adhesives is the
tendency for coagulum to build up on the metering roll when conventional
emulsions are applied by a reverse roll coater. According to [62] a tackified
acrylic emulsion is more susceptible to high shear induced coagulation. The
use of higher solids emulsions increases opportunities for particle-particle
interaction, particularly when the emulsions are subject to mechanical stress.
If there is an insufficient protective layer on the particle surface to
repel other particles, the polymer particles will fuse, eventually yielding
a significant mass of gel particles. Such coagulum may arise either from
poor mechanical stability under shear conditions, or from imperfect
doctoring leading to a thin coating on the roll, which dries and cannot be
redispersed, so that these particles eventually transfer to the application roll
and disturb the appearance of the coating. Generally, acid rosin tackifiers
show better mechanical stability than ester-based ones. In order to increase
the mechanical stability of the tackified dispersion, it is proposed to use
Manufacture of PSAs 457

Table 8.14 Drying Speed of Pure and Tackified


Water-Based PSA

Weight Loss of the Wet Coating (%)

Time (min) Pure PSA Tackified PSA

0.0 0.0 0.0


1.5 0.8 0.9
2.0 1.4 1.5
3.0 2.3 2.7
4 3.3 3.8
6 5.4 5.6
8 7.4 7.2
9 7.7 8.0
10 8.5 8.7
12 9.9 10.0
Evaluation of a 100 m coating on Mylar of a V 205/80 D/CF
52 blend at room temperature.

ester-based tackifiers to neutralize the acrylic mixture, and to add the


tackifier dispersion in a second step. Laboratory tests show a slight
difference between the drying speed of the tackified and pure unformulated
acrylic-based PSAs (Table 8.14).
Industrial experience generally indicates a higher converting (coating)
speed for tackified water-based acrylic-based PSAs (a 5–20% increase
of speed). A possible explanation of this increase is based on the dependence
of the coalescence on the surface tension. Coalescence is proportional to
surface tension, thus tackifier dispersions with a higher surface tension
display higher coalescence and allow higher coating speeds (see Chap. 3).
Die-cutting is the most important feature of the laminate influenced by
the tackifier resin. Tackifying resins influence the cohesion of the adhesive,
and thus its shear strength and its diecuttability. It is to be expected that
high softening temperature resins impart better shear and die-cuttability than
the low softening point ones. On the other hand, laminate shear and die-
cuttability depend on the hygroscopy of the components also. Tackifiers
from different suppliers exhibit different die-cuttability properties (Fig. 8.5).
A comparison of the die-cuttability of different tackifiers (tackified
adhesives and laminates) indicates that ester-based tackifiers are superior
to acid-based ones.

Tackifier and End-Use Properties. As discussed in Chap. 5, the choice


of the base elastomer and in a similar manner the choice of the tackifier
458 Chapter 8

Figure 8.5 The influence of the supplier (different formulation of the resin
dispersion) on the cuttability of the PSA paper laminate. 1), 2), and 3) are formulations
based on the same resin from different tackifier resin dispersion suppliers.

influence the end-use properties of the PSA laminate and the label; quite
different end-use properties can be achieved. As an example, for clear face
stock material the adhesive color (resin color) is very important. Lightly
colored hydrocarbon tackifiers (liquid resin tackifiers) were suggested for
EVAc- and SBS-based adhesive systems [63]. Generally, water-white- and
UV-stable water-based PSA formulations can be obtained on the basis of
hydrogenated hydrocarbon resins [64].
Japanese PSA-based tape production started with natural rubber-resin
or polyterpene/natural rubber combinations [65]. The resin and polyterpene
were replaced by C5 and C5/C9 petroleum resins due to their cost/
performance balance and grade diversification for meeting the needs of the
quality-oriented Japanese PSA tape market. The main reason for the use
of hydrogenated water-white resins in PSAs seems to be their excellent
weathering performance rather than their color and transparency char-
acteristics. Phenolformaldehyde resins give better plasticizer resistance than
rosin esters [66]. Special resins and resin formulations are suggested for
repositionable and removable PSAs. As an example an aqueous dispersion
of a methyl ester or glycerol ester of hydrogenated rosin is recommended
[67] (see Chap. 6). The tackifying results demonstrate the necessity for the
Manufacture of PSAs 459

use of soft, essentially aliphatic resins, since the more aromatic resins exhibit
a high degree of initial tack, which increases upon aging, thereby no longer
retaining the removable character of the formulation [55]. The nonreactive
resins impart a tack increasing with time to synthetic rubbers, and the
reactive resins function as curing agents for synthetic rubbers [68]. For
removable aqueous natural rubber-based PSAs, preferred classes of
tackifiers are anionic aqueous dispersions of hydrogenated rosin esters
and anionic aqueous dispersions of hydrogenated rosins [67].
The resin stability influences the end-use properties of the PSAs also.
Abietic acid contains a conjugated double bond and will react with oxygen.
This will increase the softening point of the resins and the subsequent tack
of the adhesive will decrease with time. Crystalline tackifiers are more stable
than amorphous ones. Amorphous tackifiers provide higher tack at room
temperature, but they display lower tack at lower temperatures [69].
Regarding the use of tackifiers for SIS/SBS-based formulations, while
the temperature has little effect on the properties of PSA laminates produced
on the closed die coater, it clearly has a considerable influence on those of
the roller-coated products. At the same time, tapes coated with the die
coater are superior to those produced with the roll coater [70]. Therefore the
aging stability of the resin is influencing the formulation and end-use
properties of the PSA laminate. The tackifier resin type also influences
the staining of CSBR-based PSA formulations. Rosin acid-based materials
outperform hydrocarbon resins for example, as do higher softening point
resins within any type [53].
It should be noted, that for an adequate choice of the tackifier resin the
examination criteria of the desired adhesive/end-use properties have to be
chosen correctly. As illustrated in Fig. 8.6 the sensitivity of a formulation
towards the tackifier resin nature, measured by various adhesion tests
may strongly differ. For instance, for the same formulation the optimum
of a specific adhesion performance is achieved by the use of different
resins [71].

Figure 8.6 Dependence of the adhesive properties on the tackifier nature.


460 Chapter 8

Special Features. It should be stressed that it remains very difficult to


carry out an elastomer-independent evaluation of the tackifier performance
characteristics. Solvent-based PSA formulations use natural rubber (crepe
or smoked sheets), styrene-butadiene copolymers (SBS and SIS block
copolymers) or elastomers, as well as acrylic PSAs. Rubber as an elastomer
is used for hot-melt PSAs almost only as a block copolymer, or for water-
based PSAs almost only as CSBR. Tackification of natural rubber as a
function of its status (bulky or latex) leads to permanent or removable
products [72].
Therefore the differences in the chemical nature and structure of the
rubber-based elastomers cause differences in the behavior of the same resin
upon tackification. Additional effects arise from the dispersed state of the
components. Styrene-butadiene rubbers may have a linear, branched, or
crosslinked structure. Their properties depend on molecular weight, Tg,
colloid factors, and stabilizer systems. Block copolymers commonly have
a linear structure although some radial copolymers have now become
available. Generally, rosin and its derivatives are more closely associated
with emulsion CSBR latexes [73]. With CSBR the latex with the higher
butadiene content accepts more resin and the ultimate tack is nearly double
that of the reduced butadiene system. For this system, if an 80 C SP resin is
used, the optimum tack occurs at a slightly lower resin content than for a
52 C SP resin. On the other hand, CSBR generally has a higher modulus
and requires more resin. As discussed earlier in hot-melt PSA formulations
(based on SIS), the molecular weight of the resin is more important
(hydrocarbon versus rosin). The types of tackifiers used with styrene-
butadiene rubbers are generally suitable for EVAc-based polymers as well;
a somewhat lower softening point than normal may be desirable.
Tackification with resins has its own limits. As discussed in detail in
[2], in certain cases tackification is not possible or not effective. For instance,
tackifying with resins is not recommended for film coating because of
yellowing of the adhesive due to UV radiation. For ethylene-propylene
copolymers tackification leads to softening, but such products do not have
the required viscoelasticity. The use of a number of resins is limited in the
formulation of HMPSAs because of their thermal stability. Inroll aging is
decisively controlled by the stability of the resin. The color number is a very
important resin characteristic for certain PSAs. In low temperature
applications the rigidity of the resin leads to detackification. On the other
hand tackification is limited concerning its efficiency, duration, and
environmental sensitivity too. As described in detail in [2] and illustrated
by Fig. 8.7 tackification may be global, i.e., it includes the whole polymer
(e.g., tackification of acrylates and of natural rubber), or partial (e.g.,
tackification of block copolymers); it may be permanent (e.g., tackification
Manufacture of PSAs 461

Figure 8.7 Schematical presentation of the scope and limits of tackification.

of common elastomers and viscoelastomers) or temporary (e.g., tackifi-


cation of amorphous polypropylene or chloroprene rubber, where
crystallization leads to the loss of the pressure-sensitivity). Although
tackification is required to ensure dry tack (i.e., adhesion on dry substrates)
in some special cases wet tack, i.e., adhesion on humid surfaces is needed
(e.g., hydrogels, or deep freeze formulations).

2.4 Rosin-Based Tackifiers


Rosin-based tackifiers are actually the most important class of tackifiers for
PSAs. Their success is based on the experience of them in fields other than
PSAs and on the relatively broad range of data issued from tackifying
solvent-based rubber-resin and solvent-based acrylic PSAs. In order to
evaluate their advantages and disadvantages from a technical point of view,
their contribution to the adhesion/cohesion balance and their influence on
the converting, processing, and end-use properties will be examined. On
the other hand, the compounding technology influences their technical
462 Chapter 8

suitability as well. The adhesive properties of PSAs tackified with rosins are
characterized by the adhesion/cohesion balance and the time/environment
stability. First, the most important adhesive properties (e.g., tack, peel, and
shear) will be reviewed using acrylic PSAs as base materials. The main types
of rosin resins used include acid rosins and rosin esters.

Acid Rosins
The first group are thermoplastic acidic resins, derived from pine trees,
resulting in a mixture of organic acids (abietic acid and its derivatives). These
products are modified by disproportionation, polymerization, hydrogena-
tion, and esterification. For a tackifier to impart the desired performance to
an adhesive, it is necessary that the tackifier exhibits at least partial com-
patibility with the polymer. Typically, decreasing the polarity or increasing
the molecular weight of a tackifier will decrease its compatibility. In general,
the adhesive properties of acid rosin-tackified PSAs are superior to those
of the ester-based ones [52]. On the other hand, rosin esters display a higher
water resistance than acid rosins [74]. This feature appears important when
tackifying PSAs for film coating and also concerning the shear stability of
dispersions with a high resin loading. The holding power of an adhesive is
basically a viscosity driven effect. Samples with a high degree of entanglement
(higher viscosity) would be expected to exhibit a higher resistance to shear
than a sample with few entanglements (low viscosity). It is assumed that a
tackifier loosens the entanglement network and therefore increases com-
pliance in the entanglement structure. Increases of Tg are due to segmental
friction. This stands in contrast to a plasticizer, the addition of which causes
the Tg to decrease due to a loosened entanglement network and decreased
segmental friction. For the improvement of the shear of acrylic PSAs, it
appears preferable to have a broad molecular weight distribution resin
including some very high molecular weight species, so as to obtain a high
viscosity [75]. One could expect that a molecular weight increase (and a high
melting point) imparts a similar shear improvement when tackifying with
rosin esters; high melting point, high molecular weight ester resins yield better
shear strength. This hypothesis is only partially confirmed for the properties
of the bulk adhesive; molecular weight and softening point were only second-
ary factors in controlling shear/adhesion [52]; there are some parameters
which override the effect that softening point and molecular weight have
upon shear. The tackifier’s specific formulation (i.e., surfactant type and level
and other additives) appears to have the largest impact on shear (Fig. 8.8).
The evaluation of the tackifying effect of rosin acid versus rosin ester
is more important with respect to the tack/peel balance than the shear.
The tackification of different acrylic PSAs and other PSAs with various
Manufacture of PSAs 463

Figure 8.8 Cuttability dependence on the formulation of the tackifier resin


dispersion. Variation of the cuttability values (cuttability index) as a function of the
formulation tolerances. The change of the cuttability for a water-based acrylic PSA-
coated laminate, where the same tackifier resin has been used as dispersion from
different suppliers (1, 2, 3) during an extended period of time.

resins from several suppliers will be covered next. The most important tacki-
fier dispersion commercialized since 1985 is Snowtack CF52 (Albright &
Wilson); its main competitor is CF-301 (an acid resin with a narrower
MWD that imparts better shear and water resistance, but lower poly-
ethylene peel). Originally intended to be used at a level of about 20% (on a
dry basis) both resin dispersions were used in formulations with 30–40%
tackifier. Polymers with a different chemical nature require different
amounts of tackifier resin; addition of tackifier causes the decrease of the
modulus and an increase of the Tg. The latter appears more pronounced for
acrylic VAc than CSBR in comparison to acrylics alone. At high resin levels
the modulus and shear strength also are increasing; the resin concentration
at which this change occurs is higher in blends with acrylic dispersions than
in mixtures with vinyl-acrylic dispersions [76]. Consequently acrylic PSAs
exhibit a better tackifying ability. The molecular weight of the polymers also
influences their tackifying ability; acrylic PSAs show excellent tackification
with acid rosins, but the range of the obtained adhesion/cohesion values is
dispersion related (Table 8.15).
Table 8.15 Tackification of Different AC PSAs with the Same Acid Rosin-Based Tackifier
464

Base PSA Adhesive Characteristics

Chemical Level (parts, Rolling ball 180 peel Shear at


Code Supplier Nature basis Characteristics wet weight) tack (cm) (N/25 mm) 80 C (min)

V 205 BASF AC EHA Soft AC 67 2.5 18 PT 5


R3375 Harco AC EHA Soft AC 67 2.0 18 PT 3
FP2313 Nobel AC EHA Soft AC 67 1.0 18 PT 1
PC 80 UCB AC EHA Soft AC 67 2.0 17 PT 2
80 D BASF AC EHA-AN Medium AC 67 6.0 17 60
85 D BASF AC EHA-AN Medium AC 67 4.0 19 70
2338 Synthomer AC EHA-AN Medium AC 67 6.0 16 60
FC88 UCB AC EHA-AN Medium AC 67 7.0 14 40
552 Exxon AC EHA Hard AC 67 8.0 15 50
20088 Synthomer AC EHA-EA Hard AC 67 12.0 12 90
20094 Synthomer AC EHA-MMA Hard AC 67 > 20 10 200
V 205 BASF AC EHA Soft AC 45 2.5 22 PT 5
80 D BASF AC EHA-AN Medium AC 22
R3375 Harco AC EHA Soft AC 45 3 20 PT 4
80 D BASF AC EHA-AN Medium AC 22
PC 80 UCB AC EHA Soft AC 45 3.5 18 PT 4
FC 88 UCB AC EHA-AN Medium AC 22
3703 Polysar CSBR – Soft CSBR 67 1.5 16 100
Tackifier: CF 52 (A&W) at a level of 33 parts (wet/wet). All formulations tested as paper labels.
AC, acrylic; AN, acrylonitrile;
EHA, ethyl hexyl acrylate;
MMA, methyl methacrylate;
CSBR, carboxylated styrene butadiene rubber;
Chapter 8

PT, paper tear.


Manufacture of PSAs 465

Table 8.15 shows that the same tackifier acts differently on PSAs with
a different adhesion/cohesion balance. A medium (33%) resin concentration
generally provides good polyethylene peel adhesion and excellent tack
values for soft and medium PSAs. High molecular weight, highly cohesive
acrylic PSAs do not show enough tack and polyethylene peel; these
compounds cannot be tackified sufficiently without the presence of another
PSA. Measurements aimed at comparing the effect of CF 52 and
Dermulsene 511 do not reveal any differences concerning the adhesion/
cohesion balance, converting and processing properties of the finished
product using a V 205/80 D blend as reference. Table 8.16 summarizes the
adhesion properties of tackified acrylic PSAs, formulated with rosin-based
tackifier resins from the same supplier. The aggressivity of the formulations
decreases from 1 through 5 (i.e., from acid to ester), but acid rosin tackifiers
also may yield different results (1, 3, 5). Here, tackifier 3 is less aggressive
(lower tack and peel, although an acid rosin) but it brings higher shear,
better water resistance, and better aging properties. The better processability
of tackifier 2 is probably due to its lower water sensitivity. Acrylic PSAs
display the best tackifying response with rosins due to their general
compatibility. Acid rosins are superior, providing a more aggressive tack
than rosin esters. Thus it can be expected that rosin esters would exhibit
poorer adhesive characteristics with other PSAs as well. Comparative tests
of CSBR tackification with acid rosins and rosin esters show only a slight
tack decrease, but more polyethylene peel decrease for the same resin level.
The aging resistance of the acid rosin does not meet the future, high
requirements. Although the resin’s affinity towards crystallinity, and the
level, nature, and technology of the antioxidants/formulating affect the

Table 8.16 Tackifying the Same Base PSA with Different Rosins from the Same
Supplier

Adhesive Characteristics

Water-Based Tackifier Peel (N/25 mm)


Rolling Ball Shear at
Code Chemical basis Tack (cm) Glass PE 50 C (min) Cuttability

1 CF 52 Acid 2.0–2.5 21 PT 15 PT 5–10 Poor


2 CF 51 Ester 2.5–5.0 22 PT 10 PT 8–15 Fair
3 CF 301 Acid 4.0–5.0 20 PT 9–10 35 Good
4 SE 8/21 Ester 5.5 20 PT 16 PT 50 Fair
5 CF SP Acid 8.5–10.0 18 PT 15 100 Fair
Base AC PSA: blend of V 205/80 D. Tackifiers supplied by A&W.
PE, polyethylene; PT, paper tear.
466 Chapter 8

aging resistance, the less aggressive but more stable resin esters should be
used for demanding applications. Ester rosins tackify base polymers in a less
pronounced way than acid ones. Generally, the tack and polyethylene peel
of the rosin ester-tackified acrylic PSAs is lower than that of the acid-based
ones, and due to the migration (of the low molecular weight components)
paper coatings with these resins display a higher dispersion of adhesive
property values. Butyl acrylate-based formulations are less shear sensitive
and therefore may contain (at least theoretically) high (up to 50 wt%)
tackifier loading levels. This statement points toward the use of ester rosins
at a higher level, reaching the same tack level as obtained with rosin acids.
Unfortunately, although shear performance was improved, this formulation
approach does not yield a better adhesion/cohesion balance and processing
properties are not superior (Table 8.17).
Acid rosin-based dispersions possess a higher surface tension, a lower
solids content, and consequently poorer wet-out properties than pure PSAs;
therefore formulated PSAs need additional wetting agents. The amount of
added wetting agent depends on the product quality and formulator skill.
Because of the acid pH of the mixture, there are no stability or thickening
problems with sulfosuccinates. The order of blending of the components
is arbitrary. However, changes in drying conditions could give rise to wet
coatings, telescoping, or poor die-cutting properties. Acid rosins do not
possess enough aging stability. Thus, partial or total loss of the adhesion
(end-use property) may be observed.

Rosin Esters
Rosin esters are newer products in the development of PSAs. Their
dispersing technology requires other dispersing agents and levels than with
acid rosins. Rosin ester dispersions possess inferior converting properties,
such as a poorer wet-out, a lower mechanical (stirring, shear) stability of the
liquid dispersions, and a lower pH stability. The shear of the bulk PSAs
tackified with high melting point, high molecular weight rosin esters is
theoretically superior to that of acid-tackified formulations. For a dispersion
the shear depends upon the recipe of the adhesive/resin and on the humidity
balance of the laminate. Therefore die-cutting properties of the ester-
tackified formulations are not better than those of the acid-based ones. Die-
cutting depends on the anchorage of the adhesive; consequently it can be
expected that adhesive formulations with good anchorage display better die-
cutting properties. The anchorage on the face stock and substrate depends
on the acidity of the resin. Thanks to their high acid number certain resins
(e.g., Oulumer 70 and 75E) have very good adhesion to cellulosic materials.
However, their observed die-cutting properties are not superior because of
Table 8.17 Tackifying Acrylic PSAs with a Different Chemical Basis (from Various Suppliers) with Different Rosins

Formulation (parts, wet weight)

Nr Code Supplier Base 1 2 3 4 5 6 7 8 9 10 11 12 13 14


Manufacture of PSAs

1 V205 BASF EHA 100 38 39 50 50 50 50 50 — — — — — —


2 80 D BASF EHA-AN — 22 32 10 10 10 10 10 — — — — — —
3 CF 52 A&W Acid rosin — 40 29 40 — — — 39 — — —— — — —
4 CF 50X A&W – — — — — 40 — — — — — — — — —
5 CF 51 (SE351) A&W Rosin ester — — — — — 40 — — — — — — — —
6 SE 62 A&W Rosin ester — — — — — — 40 — — — — — — —
7 Bacote 20 — Crosslinked — — — — — — — 1 — — — — — —
8 E-2395 Rohm & Haas BuAc — — — — — — — — 60 — — — — —
9 3385 BASF BuAc — — — — — — — — — 60 70 — — —
10 SE80CF A&W Rosin ester — — — — — — — — 40 40 30 40 — 40
11 XZ95043 Dow BuAc — — — — — — — — — — — 60 — —
12 X795044 Dow BuAc — — — — — — — — — — — — 60 —
13 Revacryl 491 Harco EHA — — — — — — — — — — — — — —
PE Peel N/25 mm FTM 1 — 13 13 6.9 6.8 6.7 8.2 7.0 9.3 7.6 6.7 8 5 9
CB Peel N/25 mm — — — 18 11 14 16 13 8 7 6.0 7.2 6 7.2
Loop SS N — — — 22 21 27 26 23 26 18 24 23 12 21
Loop PE N — — — 15 12 14 10 12 16 13 13 10 9.4 5
Shear RT hr — 11 15 11 28 22 38 27.6 66 54 60 30 78 42
50 C min 18–50 5 5 16 114 42 168 40 12 — — — — —
Face stock: PET (Mylar).
467
468 Chapter 8

their higher water sensitivity. In this case the highest improvement of the
cuttability may be observed by lowering the tack. Generally, rosin esters
impart lower tack and polyethylene peel than acid rosins. Many formulations
with rosin ester tackifiers are binary mixtures with a low softening
temperature component. Migration of this component causes loss of coating
weight and lowering of the tack peel values in time. Hence, the end-use
properties of the rosin ester-based formulations are inferior to those of rosin
acid-based ones. Acid rosins may be blended in without regulating pH.
Direct compounding of rosin esters with common acrylic PSAs gives rise to
mixtures with a higher mechanical sensitivity. Therefore some suppliers of
rosin esters propose a two-step compounding process. In the first step the
formulation of the acrylic PSAs occurs; next, after increasing the pH of the
acrylic mixture to 7 or 8, the tackifier is added. Adjusting the pH may cause
thickening of the whole mixture and change its rheology.
Summarizing, rosin esters were introduced as replacements for acid
rosins, due to the concerns regarding the loss of adhesive properties with
time. Their unique advantage is their supposed better aging stability. Rosin
esters require a higher loading level for the same level of adhesion. Because
of their binary formulation, rosin esters cause migration and give rise to
higher dispersion of the peel-tack values. Rosin esters possess a lower
mechanical stability and display poorer wet-out. Compounding rosin esters
requires adjustment of the pH (associated with thickening), thus it is a more
complex operation than with acid rosins. Rosin esters do not provide any
better processing or end-use properties, and are more expensive than acid
rosins. Practically, the raw materials base for rosin remains limited as they
are derived from natural products. Resources of tall oil resins remain limited
also. Because of the classical nature of the raw materials and of the old
processing technology, there are no possibilities to reduce manufacturing
costs. Thus one can assume that the price level of rosins will continue to
increase.

2.5 Hydrocarbon-Based Tackifiers


In order to determine the technical merits and shortcomings of hydro-
carbon-based tackifiers, their adhesive, converting, and processing proper-
ties need to be examined in comparison to their main competitors (i.e., the
rosins). Hydrocarbon resins possess a limited compatibility, thus it can be
expected that from the point of view of the dry adhesive, they would show
better or similar tackifying properties only for certain PSAs (e.g., acrylic
PSAs). However, difficulties may arise when tackifying high molecular
weight BuAc. However, it appears possible to design better converting or
Manufacture of PSAs 469

processing properties of the laminate, depending on the dispersing skill of the


resin manufacturer. Consequently, hydrocarbon resins are not yet universal
tackifiers, but for the most important PSAs (e.g., acrylic- and CSBR-based
PSAs) they can provide technical advantages for the dry material as well as
for the aqueous dispersions.

Adhesive Properties of Hydrocarbon Resin-Tackified Formulations


Some years ago a trend was observed in the converting industry, namely a
switch from rosin acids to rosin esters, caused by the low aging stability
of the acid rosin-based formulations. Acid rosins are the best tackifiers,
providing a good adhesion/cohesion balance for all the commonly used
PSAs. Hydrocarbon resins are generally inferior tackifiers compared to
rosin-based ones. This holds true for the instantaneous adhesive properties
of the acid rosins. Aged formulations with acid rosins, including emulsions,
show poorer characteristics than hydrocarbon resin-based ones. Rosin ester
tackifiers generally display inferior adhesive characteristics to acid rosins;
thus, opportunities exist for hydrocarbon resins mainly in comparison to
rosin esters. Hydrocarbon-based tackifiers exhibit an adhesion/cohesion
balance similar to acid rosins. Slightly better shear values are possible, but
they are not high enough to improve the processing properties. For CSBR
high levels of hydrocarbon resins are necessary (like acid rosins) in order to
achieve sufficient polyethylene peel (Table 8.18).
As the tackifier loading increases from 25 to 30%, the failure mode
changes from adhesive failure to paper tear in both the 180 peel and loop
tack tests. However, if the tackifier level is increased above 35% the shear
strength drops dramatically. Results indicate that for rosin acid-tackified
acrylic PSAs, the optimum tackifier range is approximately 30–35 wt% [52].
Shear depends mainly on the molecular weight of the elastomer and the
resin level [14]. As the resin level is increased, the shear failure time generally
decreases exponentially. This behavior is generally valid (i.e., the resin level
has more influence on the shear than the nature of the resin). It can be
expected that an equal resin loading (independent of the resin nature) would
cause the same shear decrease. Consequently, at a level of 30–35%,
hydrocarbon resins should give the same tack and cohesion as acid rosins.
However, experimental results indicate a better tack of the hydrocarbon
resin-tackified dispersion at the same resin loading level (Table 8.19).
Acrylic-hydrocarbon resin mixtures exhibit slightly better tack values
than acid rosins at the same shear and peel level. Like rosin acids
hydrocarbon resins provide a more pronounced tack improvement than an
increase in polyethylene peel. Practically, the tackifying response in
470

Table 8.18 Tackification of CSBR Dispersions with Acid Rosin/Hydrocarbon-Based Tackifiers

Formulating Components Chemical Composition

Nr Code Supplier Chemical basis 1 2 3 4 5 6

1 CF 52 A&W Acid rosin 50 — 43 33 — 20


2 E-9251 Exxon Hydrocarbon resin — 50 — — 33 —
3 3703 Polysar CSBR 50 50 57 66 66 80
Adhesive Peel (N/25 mm) on glass 19 PT 17 PT 17 PT 15 PT 20 PT 14 PT
performance Peel (N/25 mm) on PE 14 PT 15 PT 16 PT 11 15 10
characteristics Rolling ball tack (cm) 1.5 2 1.5 2.5 2.5 3
Chapter 8
Table 8.19 Tackifying Acrylic PSA with Hydrocarbon Resins and Rosins
Manufacture of PSAs

Nr Component Supplier 1 2 3 4 5 6 7 8 9 10 11

1 Acronal 85 D BASF 21.5 25.0 29.0 83.0 77.0 74.0 83.0 66.0 83.0 40.0 21.4
2 Acronal V 205 BASF 35.5 34.0 32.0 — — — — — — 40.0 35.7
3 Snowtack CF52 A&W 43.0 41.0 39.0 27.0 23.0 26.0 — — 27.0 20.0 42.8
4 E-9251 Exxon — — — — — — 27.0 38.0 — — —
Properties
2
Coating weight (g/m ) 24 25 25 25 26 26 19 20 20 — 18
Peel (N/25 mm) 20 20 PT 20 PT 19 9 14 14 15 15 17 18 PT
Rolling ball tack (cm) 7.5 2.5 4.5 2.0 2.5 1.7 2.4 2.5 — 3.0 7.5
Shear resistance
Room temperature (hr) 40 40 30 — — — 45 29 40 — —
50 C (min) 16 23 20 — — — — — — — —
471
472 Chapter 8

combination with common acrylic PSAs may be considered similar to that


of acid rosins.

Aging Properties of Hydrocarbon Resins


Acid rosin-tackified acrylic formulations could suffer a partial or total loss
of the adhesive properties due to chemical changes in the resin. Under
environmental influences abietic acid undergoes chemical transformations
giving rise to hydroperoxides. Dimerization with metal salts (rosinates) also
is possible. This phenomenon is responsible for the increase of molecular
weight, and for the loss of the tackifying effect. Aging due to acid groups
should be avoided when using rosin esters or hydrocarbon resins.
Theoretically, hydrocarbon resins possess superior aging properties, but
this hypothesis needs to be confirmed in practice.

Converting Properties
Wet-out, mechanical stability, and foam resistance of hydrocarbon resin dis-
persions are equivalent to those of rosin acids and superior to those of rosin
esters. Generally, the additional wetting agent level needed for good wetting
reaches 1.0–1.5%, compared to the 2.0–3.5% necessary for rosin esters.

Processing Properties
At the same adhesion/cohesion balance die-cutting properties depend
mainly upon the humidity balance of the laminate. This also is a function
of the hygroscopy of the adhesive layer; the hygroscopy of the adhesive layer
depends on the nature and level of the wetting agent. Hydrocarbon resin
dispersions with less surface-active agents display lower hygroscopy and
better cuttability. The few experimental measurements reported do not
confirm this hypothesis. The dependence of the shear on the drying time
(i.e., humidity content) of the layer was demonstrated. Table 8.20 lists the
hot shear gradient (GHS) and cuttability results of different formulations.
Because of the exponential dependence of the hot shear on the temperature,
an increase of the cohesion of some orders of magnitude is required for a
better shear/die-cuttability. Unfortunately, hydrocarbon resins do not
impart this shear increase.
Compounding hydrocarbon resins with acrylic- or CSBR-based PSAs
does not cause problems with the pH level of PSAs. Unfortunately, the
mechanical stability properties of some hydrocarbon resin dispersions are
not sufficient. From the point of view of the instantaneous adhesive
properties, there are no advantages to the use of hydrocarbon resins. Their
adhesion/cohesion level is similar to the level reached by acid rosins, but
Manufacture of PSAs 473

Table 8.20 Hot Shear and Cuttability Performance for Acrylic Water-Based PSAs
Tackified with Rosin and Hydrocarbon Resins

Hot Shear Values (min) Water-Based Tackifier

Hot Shear Test CF 52 DEG/G MBG 64 E-9251 8/21


Temperature ( C) A&W DRT Hercules Exxon A&W

30 90 90 330 264 1440


40 100 162 90 102 192
50 12 72 72 90 192
60 10 45 38 46 51
70 11 22 23 40 26
Cuttability Fair Good Poor Fair Fair
Adhesive: Blend of Acronal V 205/80 D with 40% resin (wet weight) was used; face stock:
paper.

without a similar general applicability. However, hydrocarbon resins age


better than acid rosins. In a first approximation the user has to choose
between rosin acids with insufficient aging stability but general suitability,
and hydrocarbon resins with good aging stability but limited usability.
The development of hydrocarbon resin dispersions was slow. Therefore,
customers have replaced rosin acids with rosin esters which display a better
aging stability but carry a higher price tag and exhibit lower tack (see
Fig. 8.9). Hydrocarbon resins are superior to rosin esters, concerning the
adhesion/cohesion balance and the price level.

Raw Material Basis


Raw materials for hydrocarbon resins are naphtha derivatives; their supply
depends on the oil processing industry. Price fluctuations for crude oil are
possible, but there is no danger of a general raw material shortage, as the
refining/cracking capacity is more than satisfactory. Hydrocarbon-based
solid resins are manufactured in many different plants. The forecast
concerning a general shortage of the C4 fraction from steam crackers has
not been proven [77]. Steam-cracked naphtha leads to aromatic, polycyclic,
C5-modified hydrocarbons, and C5 hydrocarbon resins. Hydrocarbon-based
tackifiers do not perform better than acid rosins, but are superior to rosin
esters. The large raw material basis, the low price level of the petrochemicals
(compared to natural products), the simple manufacturing technology,
and the competition between suppliers ensures a lower price level for
hydrocarbon resins. Table 8.21 summarizes the main technical and
commercial advantages/disadvantages of hydrocarbon resins versus rosin
474 Chapter 8

Figure 8.9 General applicability, aggressivity (tack), and aging stability of


different tackifier resins for PSA formulation purposes.

resins. Consequently, hydrocarbon resins should be used in the following


applications:
permanent label (paper) adhesives, as hybrid tackifiers together with
rosin esters;
permanent label (paper) adhesives, as hybrid tackifiers together with
low tack rosin acids;
permanent label (paper) adhesives, as replacement of rosin esters,
providing a similar adhesion/cohesion balance to acid rosins, but
better aging stability;
permanent label (paper) adhesives, as a less expensive replacement of
acid rosins (but not generally applicable);
permanent label (film) adhesives for clear coatings onto PVC or
other films;
removable label (paper) adhesives, but only for formulations
containing a nontacky low peel natural latex and CSBR.
From a technical point of view rosins actually perform better than
hydrocarbon resins. However, there are three major disadvantages of the
rosin tackifiers, namely the low level of adhesive properties of rosin esters,
Manufacture of PSAs 475

Table 8.21 Comparison of Rosin and HC Resins: Technical and Commercial;


Strengths and Weaknesses

Technical

Strengths Weaknesses

Rosin Hydrocarbon Rosin Hydrocarbon

Experience in use — — Limited experience


General usability — — Restricted usability
General
availability
Broad range — — Limited range
of grades of grades
Good adhesive Good adhesive — —
properties properties
Fair converting Fair converting — —
properties properties
— Good aging Poor aging
resistance resistance
— Colorless Yellowing

Commercial

— Unlimited raw Limited raw —


material basis material basis
— Unlimited Limited —
technology technology
— Lower price Higher price —

the low aging stability of acid resins, and the inadequate coating/processing
properties of the aqueous dispersion (related to its water sensitivity).
Formulations containing some hydrocarbon resins and rosin esters
offer an adhesion/cohesion balance like that of acid rosins. Consequently,
there is a real opportunity for a partial replacement of rosins in
current recipes. Today about 60% of converters are using rosins. An
amount of 10–20% of the rosin ester could be replaced by hydrocarbon
resins (i.e., a maximum of 12% of the current resin consumption). About
70% of labels are paper laminates with a maximum of 40% tackifier resin,
or 28% of the whole PSA production is tackifier based. Assuming the
12% hydrocarbon resin level could be reached, the use of about 4% hydro-
carbon resin (related to the whole amount of raw materials) seems to be
possible.
476 Chapter 8

Polyterpene Resins
Terpenes are hydrocarbons with the formula C10H16, derived (theoretically)
from isoprene. There are acyclic, monocyclic, and polycyclic terpenes; the
polyterpene resins are their polymeric derivatives. Natural and synthetic
polyterpene resins are known. Polyterpene resins offer in an adhesive
formulation the lightest color ranges of natural tackifiers in addition to the
advantages of good thermal stability and color retention. They are good
choices for ‘‘water-white’’ formulations. Although hydrogenated hydro-
carbon resins are frequently selected for these applications, polyterpenes can
often meet the color appearance of a hydrocarbon with better heat, aging,
and tackifying properties [78].
Polyterpenes are available in a wide softening point range, from 10 C
to 135 C. They exhibit good compatibility with backbone polymers,
especially those with a high ethylene content (EVAc); they also display
good compatibility with SIS and SBS block copolymers. Liquid poly-
terpenes act not only as viscosity diluents (hot-melt PSAs) like waxes, but
also as co-tackifiers. Straight and aromatic polyterpenes (with a softening
point of 80–120 C), straight polyterpenes (with a softening point of 125–
140 C), and terpene point of phenol resins (with a softening 95–155 C)
are available. Aromatic, modified polyterpenes offer improved tackifying
capabilities with high molecular weight polymers (SBS, SIS) when compared
with hydrocarbon, rosin esters, and modified polyterpenes. Terpene-phenol
resins offer the best thermal stability of any adhesive tackifier [78]. Such
resins are synthesized from phenol and terpene derivatives. They are polar,
and give adequate adhesion and thermal and aging resistance. Terpene-
phenol resins and other heteroatom-containing polar tackifier resins (e.g.,
coumarone-indene, phenolic, ketone, functionalized, and reactive resins) are
described in detail in [79].

2.6 Special Tackifier Resins


Special tackifier resins are multicomponent systems containing various
resins and/or other (nonresin) components. The most important products
are the hybrid resins and the resin dispersions. Hybrid resins are mixtures
of chemically different tackifiers. Hybrid resin dispersions based on
colophonium and hydrocarbon resins were developed in order to improve
the adhesion on polyethylene or the cohesion. The formulation,
manufacture, and stability of resin dispersions is discussed in detail in
[79]. The economical aspects of the use of different tackifier resins are
described in [80].
Manufacture of PSAs 477

2.7 Tackification with Plasticizers


Tackification may be carried out using resins or plasticizers. Generally for
PSAs formulation with plasticizers is carried out mainly for softening of the
bulky (coated adhesive) or/and reduction of the viscosity of the coatable
adhesive [2]. Such effects are desirable for removable formulations and for
hot-melt PSAs. In such cases tackification manifested as an increase of the
tack is a side effect of the plasticizing. In some applications such a side effect
is required (e.g., tackification of hot-melts), for other products (e.g.,
removable adhesive formulations) it must be avoided. Certain crosslinked
formulations contain a plasticizer intended to make it easier for the polymer
chains to slide over each other. The plasticizer may provide wetting too.
Monomeric plasticizers, oils, and polymeric plasticizers (liquid resins and
liquid polymers) are used as plasticizers.
The tackifying mechanisms and the influence of the tackifying agent
on the Tg is quite different for plasticizers and tackifiers. The addition of
resins to an adhesive increases the Tg of the base polymer, the addition
of plasticizers decreases it [81] (see Chaps. 2 and 3). Tackifier resins may
increase the tack, without an important decrease of the shear resistance;
plasticizers increase the tack, but lower the cohesion. It remains difficult to
make a sharp distinction between tackifiers (resins) and plasticizers (solvents)
as a result of their chemical basis. Actually both low molecular weight liquid
resins and relatively high molecular weight plasticizer liquids (solvents) are
known. The use of liquid tackifier resins induces the same rheological changes
as with plasticizers. As an example, tackifying styrene-butadiene copolymers
is more difficult than tackifying SBS ones. A solvent-based copolymer with a
plateau modulus of 55  10 kPa needs a much lower Tg tackifying resin
in order to decrease the modulus than an SBS copolymer (plateau modulus
4.3  106); here the tackifier should act like a plasticizer [82].
The influence of the chemical nature, structure, and molecular weight
of solvents used as plasticizers was discussed [83]. Plasticizers used may be
solid (sulfonamide) or liquid (phthalates, phosphates, glycolates, etc.) [84].
Generally, plasticizers are reactive liquid components that interact with
PSAs. As discussed in [85] the interaction of the plasticizer with the polymer
is characterized by the specific interaction parameter (w). The value of this
parameter depends on the chain length, double bonds, and branching in
the plasticizer. The reduction of the Tg increases with the decrease of w,
i.e., of the solvation. Polymer incompatibility with the plasticizer results
in an increase of the glass transition temperature. Adhesives based on
polyisobutylene (PIB), EVAc, polyvinyl ether (PVE), and thermoplastic
elastomers (SIS, SBS) are very sensitive towards compounds like dioctyl-
phthalate [86]. Aromatic plasticizers like dioctyl phthalate (DOP) associate
478 Chapter 8

with the polystyrene domains and act like aromatic oils (i.e., soften the
polystyrene domains). Therefore, a barrier layer must be present between
PVC and Cariflex TR-1000 (Shell). Including plasticizer into poly(butyl-
acrylate) dispersions brings good removability after long time storage, but
produces migration, low cohesion, and adhesive transfer [87]. Plasticizers
increase tack but lower cohesion.
Classical plasticizers are used on a large scale in PSA applications. For
removable PSAs, recommended compounds include: 2-diethyl hexyl esters
of phthalic acid and 2-diethyl hexyl adipate, esters of sebacic and aceloinic
acids with 6–10 alcanols, alkyl esters of succinic, glutamic, adipic acid with
epoxidized soybean oil, phenol alkyl sulfonic acid ester, and propylene
glycol-alkyl-phenyl ether, di-2-ethyl-hexyl ester of thiodipropionic acid,
acetyltributylcitrate, and glycoldibenzoate; phthalic acid esters such as
di-n-butyl phthalate, di-isobutyl phthalate, di-2-ethylhexyl phthalate, and
polypropylene glycolmethylol ether [83,87]. The choice of the plasticizer
depends on the chemical basis of the PSA, application technology, and end-
use requirements. The compatibility of different adhesive raw materials with
plasticizers was discussed in a detailed manner [88]. The plasticizer nature
depends on the adhesive nature. As an example, for SBS butyl-benzyl
phthalate was suggested [42]. For electron beam-cured hot-melt PSAs
Parapol 350 (Exxon) and Flexon 1076 (Shell) are used as plasticizers. In this
case formulation ingredients like plasticizers have to be inert when subjected
to irradiation. To obtain optimum crosslinking with a minimum of
radiation, the selection of suitable resins and plasticizers is necessary (e.g.,
PSAs based on multiarmed SIS block copolymers crosslink much faster with
saturated than with unsaturated resins and plasticizers). On the other hand,
the rheology changes due to the plasticizer have a quite different scope in
permanent and removable adhesives. In permanent adhesives they improve
the tack, in removable ones they have to make the adhesive softer in order
to absorb the peel energy. Solvents and plasticizers increase the viscosity
of PSA dispersions [89]. The sensitivity of different plasticizers toward
antimicrobial agents also is quite different. Their environmental relevance,
toxicological findings, and suitability for medical products are other aspects
to be taken into account at the time of their selection [90]. A common level
of 20% (wet/wet) is proposed for water-based permanent adhesives.
Compounding water-based acrylic PSAs with more than 10% plasticizer
requires special skills (e.g., plasticizing of Acronal 120).
Common plasticizers are low molecular (micromolecular) compounds;
other macromolecular products also may be used. As an example Parapol is
used as a tackifier/plasticizer. It is a viscous, liquid copolymer of n-butene
and isobutylene. The viscosity of polybutenes ranges from light oil to highly
viscous oils. Their tackiness increases with increased molecular weight. In
Manufacture of PSAs 479

hot-melt PSAs they increase tack and also are applied to obtain softness and
low temperature flexibility, and to control cohesive strength [91]. For many
years polybutenes were successfully used as tackifiers/plasticizers in hot-melt
and solvent-based adhesive systems. Later the polybutene dispersions were
incorporated into water-based PSAs. Acrylic latexes were used because of
their increased compatibility with polybutenes. The successful incorporation
of polybutenes (at 20–30% addition levels based on the final ‘‘dry’’ film)
would result in significant cost savings. In hot-melt PSAs polybutenes
plasticize and provide tackiness by associating with the isoprene midblock,
which has negligible inherent tack. Polybutene contributes to the aggressive
tack and quick stick properties necessary for adhesive applications [92].
Generally, polybutenes can be incorporated into acrylic latex systems by
either of two methods, either by addition to the acrylic polymer latex,
where it acts as an external plasticizer, or by addition to the acrylic
monomers, prior to emulsion polymerization, where it acts as an internal
plasticizer. Polybutenes, polyisobutylene, and mineral oils were suggested
as plasticizers for C3/C4 polyalphaolefins also. Plasticizers are described in
detail in [93].
Common plasticizers known from the formulation of PVC are polar
substances. Such compounds are used mainly for polar elastomers based on
acrylics or EVAc copolymers. For the mostly hydrocarbon-based thermo-
plastic elastomers, nonpolar, hydrocarbon-based plasticizers are preferred
[2]. Plasticizers for HMPSAs include the aromatic, paraffinic, or naphthenic
extender oils. Paraffinic oils are compatible with the midblocks of
unsaturated or saturated rubbers, naphthenic oils are compatible with
both. Aliphatic plasticizers soften the SBCs, with a minimum effect on
their cohesion and adhesive properties. Aromatic plasticizers act on the
polystyrene blocks and destroy their association. The formulation with oil
gives different results for SIS and SBS block copolymers. The oil/resin ratio
does not influence decisively the shear resistance of SIS-based formulations;
the decrease of the shear resistance caused by the oil can be balanced using
an endblock reinforcing resin. SBS-based adhesives need mid-block active
resins and processing oils. In this case formulations with oil and resin give
higher storage modulus [94]. Generally the increase of the oil concentration
in HMPSA formulations leads to decrease of the tack and peel. A detailed
discussion of the formulation with oils is given in [95].
Plasticizers affect the processing and coating technology of the
adhesive. As discussed above viscosity regulation is one of the reasons for
their use in HMPSA formulations. In water-based formulations they
influence the drying speed also. For postcured adhesives the choice of the
plasticizer affects the curing speed. Migration of the plasticizer in the
adhesive may cause its diluting and changes in the adhesive, converting, and
480 Chapter 8

end-use properties thus formulation for plasticizer resistance needs special


skill (see Section 2.11).
In principle the scope of the formulation is to avoid polymer
degradation, but in some cases degradation is used as a manufacturing
method to modify or to transform one macromolecular compound in
another, or as a formulation method to provide a balance against excessive
molecular weight build-up. Degradation as a manufacturing method can be
carried out via mechano-chemical depolymerization or radiation-induced
depolymerization [2]. Generally such procedures lead to increased tack and
decreased peel resistance and cohesion.
The most important additives influencing the adhesive properties are
the tackifiers (resins and plasticizers). Detackifying agents are known too.
Principally detackification is the result of excessive immobilization of the
macromolecules. Therefore the methods which cause such immobilization
(e.g., crosslinking, functionalization, defunctionalization, etc. can lead to
detackification. Built-in nonpolar voluminous groups, deactivation of the
built-in polar groups, or improvement of the rigidity of the macromolecular
network provide detackification. The chemical possibilities for detacki-
fication are illustrated by the manufacture of release agents and by
formulation for removability (see Section 2.8). Detackifying resins are
known also. Detackification is discussed in detail in [96].
Tackifiers change tack and peel in parallel. Other ingredients such as
crosslinking agents have a similar effect (see Chap. 6). Their choice will be
discussed in Section 2.8. There are many formulating additives which are
used for other purposes, but also influence the adhesive properties; such
ingredients include fillers, antioxidants, surfactants, etc.

2.8 Cohesion Regulation


Generally cohesion decrease is a side effect of the tackification. In principle
cohesion can be improved by the increase of the molecular weight,
crosslinking, or reinforcing (filling) [2]. Cohesion increase is mainly required
for shear related applications and for removability.
The scope of the crosslinking is always the increase of the cohesion.
Such an increase may be moderate (in order to improve the peel resistance),
pronounced (to increase the shear resistance), or excessive (to allow
cuttability and die-cuttability). Although reinforcing tackifier resins are
also known, in adhesive manufacturing practice mainly crosslinking and
filling are used for the regulation of the cohesion. Crossslinking itself leads to
an increase of the molecular weight also and reinforcing with fillers can cause
crosslinking too. Crosslinking is based on the built-in or formulated reactivity
of the components of the PSA. The build up of a network reduces the chain
Manufacture of PSAs 481

mobility and improves or reduces the elastical deformability. Crosslinking


occurs during polymer synthesis also; post-synthesis crosslinking of pressure-
sensitive adhesives is used to control the adhesion-cohesion balance or to
provide other converting or application related properties. As discussed in
Chap. 5, chain entangling is responsible for rubber-like properties of
uncrosslinked elastomers. At sufficiently low frequency of deformation
linear chains begin to disentangle (reptation). One chemical crosslink per
chain is sufficient to prevent flow by reptation. However, entangling accounts
for 75% of the equilibrium modulus even at high degrees of crosslinking.
Crosslinking is based on chemical reactions between the macromolecules or
between the macromolecules and other components of the formulation. Such
chemical reactions can be carried out as common chemical reactions or as
polymerization. Polymers can be divided in the following classes concerning
their reactivity: noncrosslinkable, self-crosslinkable, and thermosetting
polymers. The best known curing methods include: thermal, self-cure,
moisture, and radiation crosslinking. The chemical basis for crosslinking was
described in Chap. 5, special aspects are discussed in Section 2.11.
Fillers can modify the rheology of viscous or viscoelastical systems,
without affecting their viscous character, or they can build up an elastical
network in such systems, transforming them into viscoelastomers or
elastomers [2]. Filled systems are special composites where the role of the
filler can be played by various components (e.g, crystallites, polymer
sequences, fillers, etc.). There are nonreactive fillers working physically and
reactive ones which can build up a chemically linked network (Section 2.11).
Fillers used for PSAs are described in detail in [97]. Reinforcing and special
fillers (e.g., neutralizing, crosslinking, conductive, flame retardant, pressure-
sensitive, water soluble, and expandable fillers) are used.

2.9 Coating Properties


In order to wet the face stock or release liner material during coating PSAs
must display a good wet-out ability. They must also resist shear forces
during coating (metering device) and thermal shocks during drying.
Formulating the adhesives in order to optimize the coating properties
intends to improve these characteristics. Wetting out depends on the receiver
surface. It is evident that direct coating on paper or film is less difficult than
indirect coating via a release liner, because of the fairly different surface
tensions of these materials. On the other hand, direct coating on paper is
easier (concerning the wet-out) than direct coating onto film as the nature
of the face stock film influences the wetting ability. Polar surfaces (PVC,
PC, PET) are easier to coat than polyolefins. The nature of the silicone
(solventless or solvent-based) influences the wetting behavior as well.
482 Chapter 8

Depending on these factors, different wetting agents at different levels can be


added to the formulation. When direct coating, bleedthrough (migration)
should be avoided; a high viscosity and high solids content is needed.
Drying, when direct coating, is limited by the temperature resistance
of the face stock material. Formulating is limited by the chemical resistance
of the face stock material. Different coating devices also require different
coating viscosities. The drying oven capacity implies a lower or higher solids
content. These few examples illustrate how coating properties of PSAs affect
their formulation. Next, the influence of the coating technology on the
formulation will be discussed.

Influence of Coating Technology on Formulation


The most important feature of the coating technology is its direct or transfer
character. In direct coating the adhesive is deposited onto the face stock
material, mostly as a low viscosity liquid. In transfer coating PSAs are
temporarily coated onto a siliconized web, the adhesive transfer onto the
face stock material is carried out later after the drying operation (i.e., the
coating properties of the liquid adhesive have to be designed for the release
liner). Therefore quite different formulations are required for direct and
indirect or transfer coating. In some special cases (Monoweb) the PSA
coated onto the face stock material is protected by the back side of the face
stock material. Such applications are encountered mainly for protective
films or tapes where a release layer is coated on to the back side of the face
stock material or the chemical composition of the face stock material allows
the peel off of the adhesive.
The quite different requirements for the formulation of PSAs as a
function of the coating technology are illustrated by the design of water-
based adhesives. In this case, for direct coating neither liquid resins nor peel
modifiers should be used, and only a low level of crosslinking agents; highly
viscous, thickened dispersions are preferred. When necessary, surfactants
with good water resistance should be used. For indirect or transfer coating,
as a function of the liner nature, varying levels of wetting agents are needed.
Defoamers may be added also, but only at a low level. A higher mechanical
stability of the dispersion is required. The formulation for direct/transfer
coating influences the release force from the release liner also. Aqueous
dispersions damage the siliconized release paper, and therefore the release
force between a transfer-coated adhesive and the release liner is higher. An
adhesive required to be coated directly onto a web via Meyer bar appli-
cation, requires a different set of intermediate properties (e.g., wet-out, total
solids, viscosity) compared to one designed for transfer coating on a reverse
roll coater via a silicone release liner.
Manufacture of PSAs 483

Influence of Face Stock Material on Formulation


Wetting out on paper is easier than on films and a higher surface tension
can be tolerated; for the same reasons a lower viscosity may be used.
Unfortunately, low viscosity dispersions or solutions migrate easier through
porous paper. Wetting out on polymer films is more difficult. Here the
polarity and the roughness of the film influence the coatability. As an
example, white or clear PVC displays a different wet-out ability and needs
different surface-active agent levels. Polyolefins are difficult to be coated as
their slip agent content influences the wettability. The porosity of paper and
the nature of the coating method influence the coating weight as well as the
adhesive/converting/end-use properties.

Influence of Coating Machine on the Additive Choice


Shear on the coating machine depends on its construction and functional
parameters, as does foaming. Shear contributes to good wetting; foaming
can be avoided with the use of defoamers, therefore it works against wet-
out. Too high shear on the metering device or in pumps produces coagulum;
hence post-stabilizing with surface-active agents can be necessary. Insuffi-
cient thermal (drying) capacity requires the improvement of the volatility of
the adhesive carrier (solvents or water).
A given metering device can be used for a certain viscosity range.
Engraved cylinder rolls with different line screens use different viscosities
and solids contents. Different flow properties are required for different types
of metering devices (e.g., Meyer bar or gravure cylinder) and even the same
formulated adhesive coated on different coating devices may lead to quite
different coating weights and adhesive properties (Table 8.22).
Therefore different versions of the flow and adhesive properties are
needed as a function of the coating machine. Adjusting the adhesive’s

Table 8.22 The Influence of the Coating


Technology on the Adhesive Properties

90 Peel (N/25 mm)

Coating device
Theoretical Coating
Weight (g/m2) Meyer bar Gravure roll

0.7 0.37 0.18


0.8 0.42 0.27
0.9 0.48 0.31
1.0 0.60 0.40
484 Chapter 8

viscosity (via formulation) for any machine is important for all kinds of
adhesives. For water-based PSAs, dissolving nonionic alkylphenol-oxyethy-
lates may be accompanied by an increase of the viscosity. Therefore warm
water should be used; later it can be diluted with cold water [98]. In some
cases (certain sulfosuccinates) the improved wettability is given by the
increase of the viscosity (and decrease of the surface tension). In other
gravure coating applications dilution of the raw dispersion and subsequent
rethickening are used in order to avoid gravure-induced structures (i.e.,
optical defects in the coating).
The choice of the processing oil has the greatest impact on hot-melt
PSAs performance. The greater solvency power of naphthenic oil produces
lower adhesive viscosities and lower shear adhesion failure temperatures,
while increasing the tack. Peel values strongly decrease with the use of fully
saturated mineral oils [99]. The flow properties of styrene-diene block-
copolymer-based hot-melt PSAs are improved by polyethylene-based
polymers. Modification with low molecular weight polyethylenes results in
significantly better flow properties, allowing for less energy intensive mixing
and more uniform deposition of the molten adhesive [100].
Polyacrylate rubbers with relatively low molecular weight can be
dissolved directly in a polar solvent without milling. However, milling the
elastomer would further help in lowering the viscosity, this being required
for certain coating devices. On the other hand, in general, milled elastomers
possess less cohesive strength (i.e., require less tackifiers).
Precautions must be taken to avoid degradation of hot-melt PSAs by
oxidative attack during the application of the adhesive. During coating the
thin film of adhesive is spread on the web at elevated temperatures. In roller
coaters where the adhesive is transferred from the heated roller, the adhesive
is exposed to air, while in die coaters the adhesive is extruded onto the
web in a closed system and air exposure is less critical. Therefore, when
formulating, the type of coater should be considered [101]. The nature of
the adhesive influences the drying process [102]; different adhesives require
different driers (length of drying oven, temperature, air volume, temperature
gradient). Conversely, for a given oven, the adhesive formulation should be
adjusted in order to obtain an optimum drying velocity.

2.10 Converting Properties


The processing properties of the laminate will be discussed separately. In
this section only the most important of them will be covered; they are
printability and cuttability. These are influenced by properties such as
migration, lay-flat, and reel stability.
Manufacture of PSAs 485

Migration
The influence of the face stock material on migration was covered in Chap.
2. It can be difficult to store label stock for any length of time due to the
aging/bleeding problems. The adhesive bleedthrough often interferes with
the quality of the art work and printing. Reels often bleed at the bottom and
dry out at the top (gravity bleeding). There are a lot of difficulties with high
gloss labels bleeding through. Generally, migration leads to the deteriora-
tion of the face stock, deterioration of the adhesive properties by lowering
the coating weight, phase separation in the adhesive, and enrichment of one
of the components. Phase separation is caused by the different melting
points of the components and different molecular weight distributions.
Acrylic PSAs are sensitive to this phenomenon, especially when used on face
materials which promote migration of oils into the adhesive, thereby
lowering the cohesive strength [103]. Generally, the following factors
influence the migration (penetration, bleedthrough, staining) of the adhesive
in the face stock: the characteristics of the adhesive and face stock, and the
machine conditions. As discussed in [104] the adhesive related parameters of
the migration include the base elastomer, the tackifier, and the plasticizer.
The nature, molecular weight, and gel content of the base elastomer strongly
affect the migration. The nature and softening point of the tackifier and the
nature and boiling temperature of the plasticizer influence the migration
also. The top layer of the paper can include mineral oil soluble or insoluble
binders. For dispersed systems the solids content, the surfactants and the
thickener nature and concentration affect the migration too. The influence
of the face stock on the migration was discussed earlier (see Section 3.2 of
Chap. 2). A detailed analysis of the carrier related parameters of the
migration is given in [105]. Next the influence of the adhesive characteristics
on the migration will be covered.

Influence of Adhesive Characteristics. The chemical composition of


the adhesive, the formulation of the adhesive, the tackifier, the additives,
and their chemical and macromolecular characteristics all influence the
migration. Migration starts with an almost sudden temperature increase
which depends only on the formulation of the adhesive, not on the paper
properties. The depth of the migration is given by the apparent density of
the paper. The phenomenon of penetration has been well studied for paper
manufacturing [106], where ink penetration depends on the concentration,
the viscosity, the water retention properties, the temperature, and the
chemical composition. In a similar manner the same parameters influence
the migration of PSAs. Water retention and viscosity of water-based PSAs
depend on the surface-active agent content and its nature. In the particular
486 Chapter 8

case where the adhesive is to be used to form laminates and at least one of
the surfaces is a printed surface, the presence of any residual surfactant can
lead to discoloration or bleeding [107]. Staining is a common characteristic
with water-soluble adhesives [108].
There are two separate conditions causing staining: pressure (reel-
wound laminates) and heat (storage). For CSBR, in both cases the level
of staining is related to the butadiene content and molecular weight
distribution of the base polymer, the tackifier resin type, and its softening
point and emulsification system [52]. Generally, staining increases with
increasing butadiene content, but decreases at higher gel contents, although
not proportionally. The tackifier resin type has a definite effect; rosin acid
tackifiers outperform hydrocarbon resins for example, as do higher
softening point resins within any type. Antioxidants and stabilizing agents
can migrate too.
Molecular Weight of the Tackifier Resin. Hot-melt PSAs usually
contain SIS or SBS rubber block copolymers, tackifiers, and hydrocarbon-
based oil. The use of low softening point tackifiers and/or a high
concentration of an oil may be necessary to achieve higher tack, better
low temperature properties, reduced melt viscosity, removability, and lower
cost [100]. However, upon aging the much lower molecular weight
components can migrate from the hot-melt PSAs. Plasticizers migrate
from flexible PVC film and upon entering a PSA cause loss of adhesion
due to a lowering of the PSA’s cohesive strength. Because they are based
on relatively low molecular weight components, hot-melt PSAs are more
adversely affected by migration of plasticizers. Generally, low molecular
fractions penetrate more. Measuring the molecular weight distribution in
the area where migration occurred, it was found that a high molecular
weight fraction has the same concentration in the middle and at the
extremities (i.e., center and outer range), but the low molecular weight
portion has a higher concentration at the extremities. Some agreement could
be found between the starting temperature of migration and the softening
point temperature, as measured by the ring-and-ball method [45]. According
to the theoretical model, the self-diffusion rate Rd is inversely proportional
to the molecular weight [109]:

Rd ¼ f ð1=MW2 Þ ð8:3Þ

For rubber, the gel/sol ratio influences both cohesive strength and
migration. For those skilled in the art of formulation the dependence of
the migration on the molecular weight is well known from the practice of
thickeners. Thus different grades of polyvinylpyrrolidone (PVP) with
Manufacture of PSAs 487

different molecular weights cause different levels of migration. As an


example PVP-K-15 (low degree of polymerization, DP) gives migration,
whereas PVP-K-90 (high DP) does not lead to any penetration in paper.
When Hyvis 10 polybutene is used, the polybutene exudes from the adhesive
and soaks through the backing release paper; therefore only the high
molecular weight Hyvis grades (i.e., 200 and above) should be used [54].

Machine Influence. The coated release liner may be laminated to the


face stock, in a pressure nip between a rubber roll and a steel roll. This
technique accomplishes the transfer of the adhesive mass to the face stock
with a minimum of penetration [110].
The incorporation of 1.5% amorphous co(polyethylene) (A-CPE) in
an SBS copolymer-based hot-melt PSA prevents paper face stock
bleedthrough upon aging at 158 F for 2 weeks. The ability of these low
molecular weight polyethylene and polyethylene copolymers to gel organic
liquids is responsible for the reduced migration of low molecular weight
tackifiers and oils [100]. Another possibility to avoid bleedthrough is the use
of barrier coatings. Resin compatibility also influences migration; colloidal
silica reduces the bleeding as well.

Cuttability
In Chaps. 2 (rheology) and 7 (converting properties) the parameters
influencing the cuttability were outlined. The adhesive nature (solvent-
based, hot-melt, or water-based) influences the cuttability, but within the
same class of adhesives the adhesion/cohesion balance also plays an
important role. Hot-melt adhesives display good overall adhesive perfor-
mance on a wide range of substrates including polyolefins, and show high
resistance to humid conditions.
In the United States, hot-melt technology represents about 50% of the
label stock production. In contrast, the convertability of hot-melt adhesives,
which (together with the thermal and aging resistance) had been questioned
at the early stages of development of hot-melt technology, has limited the
potential in Europe [111]. Poor cuttability of hot-melt PSAs limits their
use for labels. The formulator has the option to use other PSAs when good
cuttability is required. Later on formulations were modified in order to
improve the cuttability of hot-melt PSAs. Thus silicones (polysiloxane
additives with a molecular weight of up to almost 10,000) are mixed with
tackified synthetic rubber-based PSA compositions to reduce edge ooze
or cold flow upon cutting sheets coated with such components. Either
nonreactive or reactive polysiloxanes can be utilized [112]. Figures 8.10
and 8.11 illustrate the die-cuttability of film laminates, as a function of
488 Chapter 8

Figure 8.10 Cuttability of film laminate as a function of the peel. 1), 2), and 3) are
different water-based acrylic PSAs.

Figure 8.11 The influence of the formulation on the cuttability. Cuttability of


paper laminate as a function of the peel.
Manufacture of PSAs 489

the peel for water-based and solvent-based formulations. The formu-


lator should use quite different recipes for a given peel or cuttability target
(Figs. 8.10 and 8.11).

2.11 End-Use Properties


Technical and commercial factors influence the choice and formulation
of PSAs. The most important technical parameters include the end-use,
coating, and converting properties of the adhesive. In this section, the
influence of the end-use properties of the adhesive on its formulation will be
examined. The end-use properties of labels differ according to the face stock
material used (paper or film) and the character (permanent or temporary) of
the adhesive bond. Therefore it is recommended to examine the principles of
formulating separately, in light of these criteria.
There is a wide variety of quality requirements corresponding to
different end-uses. In some cases different performance characteristics such
as clarity, temperature resistance, removability, water resistance and
solubility, etc. are required. Generally, both of the main adhesive
components affect these performance characteristics (see Chap. 5), but in
many cases formulating additives play a decisive role. As an example, acrylic
hot-melt PSAs are of great interest for clear film applications and other
applications where UV stability is critical. Hydrocarbon resin-based
tackifiers also are recommended for such uses. Hydrogenated pure
monomer resins have a better color stability [45]. On the other hand, a
crystalline rosin is thermally more stable than an amorphous one [113]. An
improvement of the UV stability may be achieved with stabilizing agents.
Medium softening point tackifiers provide balanced adhesive pro-
perties, including strong specific adhesion to polyethylene and other
polyoleflns. For deep freeze labels a hydrogenated, water-white liquid
resin was proposed, with good tack at temperatures down to 30 C [114].
Lower softening point tackifiers help to improve the formulations’ low
temperature adhesive properties for frozen food packaging, cold tempera-
ture tapes, and labels [78].
The resin level has its influence also. For the curved panel lifting test
(carried out at 150 C) one should use more than 10% tackifier; above 40%,
the label is less likely to be stripped off cleanly. Water-based ethylene
copolymers possess a better resistance towards plasticizers migrating from
PVC [115]. A comparison of acrylic PSAs with CSBR and vinyl acetate/
ethylene copolymers shows that acrylic PSAs possess an adequate plasticizer
resistance, aging stability, and heat resistance, CSBR exhibits better
adhesion to polyolefins, and EVAc displays good plasticizer resistance,
490 Chapter 8

aging stability, heat resistance, adhesion to polyolefins, and an adequate


rheology [116].
Plasticizer resistance may be improved by the appropriate choice of
the formulating additives. Phenolformaldehyde resins provide better
plasticizer stability than rosin esters [66]. The addition of rubber dispersions
may improve the plasticizer stability also [117]. Test data indicate that the
incorporation of 1–5% acid-polyethylene derivatives into hot-melt PSAs
contributes significantly to obtaining a stable bond to plasticized PVC [100].
The choice of an adequate base elastomer with plasticizer or low tackifier
level leads to removable PSA formulations. On the other hand, the proper
use of crosslinking agents, plasticizers, or slip agents can also ensure
removability. High temperature resistance may be obtained with highly
cohesive base elastomers containing self-crosslinking units, with high
melting point tackifiers or crosslinking agents and fillers. A high tem-
perature resistance may be given by small glass microbubbles [118]. On the
other hand, incorporating glass microbubbles also may impart a good
removability [119].
For many different end-uses a general problem remains the adjustment
of the release force (i.e., formulating for release and labeling ease), which
depends on the interaction between the adhesive and the release liner. This
interaction depends on the adhesive’s nature, the release liner nature, as well
as on coating and converting conditions. Adhesives with a different chemical
basis show different release levels from the same release material. On the
other hand, solvent-based, water-based, or solventless silicones display
different release forces with the same adhesive. A chemical interaction
between water-based acrylic PSAs and solventless silicone is possible also.
The release force also is a function of the age of the laminate.

Permanent and Removable Labels


Pressure-sensitive adhesives for labeling applications are usually classified as
either removable or permanent. The removable adhesive, once applied to a
substrate, must be able to be removed after a residence time on the substrate
and at the various temperatures the substrate may be subjected to [120].
High tack, high peel adhesion, and in most cases high cohesive strength
PSAs are used for permanent labels. Good anchorage on the face stock and
on the substrate is needed. Affinity towards nonpolar surfaces is necessary.
Low peel, medium tack, and medium cohesion are needed for removable
PSAs. No adhesive build-up and adhesive break are allowed. Evidently such
very different requirements lead to quite different formulations. Adhesive
break and cohesive failure are characteristic for common hot-melt SIS-based
PSAs. Therefore common hot-melt PSAs are not adequate for removable
Manufacture of PSAs 491

labels. In Chap. 2, the parameters influencing the removability and the raw
material basis were examined in detail. Here, the special features for the
formulation of removable PSAs are covered.
The nature and level of the tackifier finally depend on the end-use
requirements of the laminate. Removable or permanent use, and film or
paper coating require different tackifiers and perhaps a different tackifying
technology as well. Cost factors (the price level of the final label) also
influence the choice of the tackifier. The properties which contribute
to peelability are principally limited tack, a low build-up factor, and a
relatively soft adhesive. These properties can be achieved conventionally by
omitting or only using a low level of tackifier, including tack deadeners
(such as waxes), and by including plasticizers. Some removable formulations
include natural rubber and/or CSBR latexes; hence they need hydrocarbon
tackifiers. On the other hand, soft removable adhesives tend to migrate;
therefore they do not need liquid tackifiers. When using natural rubber-
based formulations, the only possibility is to change the molecular weight of
the base elastomer (i.e., to use milled or unmilled rubber) in order to modify
the cohesion and the peel adhesion (i.e., removability) of the adhesive.
In this case the choice and level of tackifiers used is more important.
Another possibility exists in the use of plasticizers or crosslinking agents.
Experimental results demonstrate the necessity for the use of soft, essentially
aliphatic resins, since the more aromatic resins exhibit a high degree of
initial tack which increases on aging, so that they are no longer removable
[120]. Glycerol esters of highly hydrogenated rosins also may be used [121].
Common liquid plasticizers include polybutene/isobutylene derivatives.
Polybutene and polyisobutene display a quite different storage and aging
behavior; polybutene suffers upon aging and crosslinking, and becomes
hard. Polyisobutene depolymerizes during aging and becomes soft and
tacky. The natural rubber latex component also imparts a releasability
characteristic to the adhesive, such that the adhesive-coated face material
after it has been pressed onto a contact surface, can be easily removed by
simply manually lifting the face material from the contact surface. The
natural rubber latex component in combination with the tackifying resin
has the additional property that little adhesive residue is left behind on the
contact surface when the adhesive-coated material is removed. A level of
about 30% natural rubber latex in a vinyl acetate-maleinate dispersion
yields a removable PSA, leaving no residues behind upon debonding.
Styrene multiblock butadiene/styrene copolymers (43% styrene) were
proposed as removable hot-melt PSAs [99]. The suitability of a block
copolymer as a removable PSA is heavily dependent on the other ingredients
used in the adhesive formulation, such as tackifiers and processing aids; for
hot-melt PSAs, SIS-type rubbers are preferred since SEBS shows excessive
492 Chapter 8

build-up of the peel adhesion. An SBS rubber (30% styrene/70% butadiene)


in addition to exhibiting some build-up of the peel upon aging, leaves a
substantial amount of residue on the substrate [120]. Hence removable hot-
melt PSAs should be formulated on the basis of SIS block copolymers, low
softening point aliphatic resins, and salts of fatty acids. The resins have
a softening point below 30 C, as determined by the ASTM E-28 ring-and-
ball method. Commercially available low softening point aliphatic resins
recommended for removable hot-melt PSAs include Regalrez 1018
(Hercules), Escorez 1401 and ECR-327 (Exxon), Wingtack 10 (Goodyear),
and Zonarez 25 (Arizona/Bergvik). These resins are used in removable hot-
melt PSAs in amounts of 20–50 wt% [120]. Removable hot-melt PSAs may
contain up to about 25 wt% of plasticizing or extending oil in order to
provide wetting and/or viscosity control; such components include paraffinic
and naphthenic oils.
The use of liquid components is characteristic for the first removable
water-based formulations as well, where plasticizers (up to 20 wt%) [13,122]
were suggested in the recipe. Such a high plasticizer content, however,
increases the danger of bleeding; polyisobutylene (PIB) may be used as a
liquid softening component. The lack of age hardening has made butyl and
PIB the first choice as the elastomer base for hot-melt removable label PSAs.
Latex PIB dispersions were suggested for water-based PSA formulations.
However, the staining is more pronounced for removable PSA labels with a
liquid component in the formulation.
The removability of acrylic-based PSAs is improved by the addition
of small amounts of organofunctional silanes (e.g., initial peel strength of
0.057 kg/cm after 20 min at 22 C, and a final peel strength of 0.065 kg/cm
after 7 days at 22 C compared with 0.097 and 0.389 respectively for a similar
adhesive without silane) [123]. Modified, crosslinking methyl polysiloxanes
also may be used [124]. Polysiloxane grafted copolymer PSAs give
repositionable labels [112]. These copolymers have pendant polysiloxane
grafts which cause the exposed surface to initially have a lower degree of
adhesion.
Fillers also change the peel adhesion value as contact hindrance caused
by filler particles reduces the peel. Therefore in some cases inert fillers are
used in removable formulations. Different mathematical formulas were
proposed by different authors to describe the modulus increase by fillers
[125]. Fillers reduce wetting and flow performance characteristics of hot-
melt PSAs, but may improve their cold flow [126]. Large amounts of
nonreactive pigments, such as clay and calcium carbonate, can be tolerated
without noticeable effect on the peel. Zinc oxide and in some cases colloid
silica are active pigments and much smaller amounts can be tolerated. The
properties of fillers used for PSAs have been described [127–130]. Aqueous
Manufacture of PSAs 493

PSA dispersions and polymer solutions can contain more than 2%, often
more than 5%, and even more than 10% fillers, colorants, and/or extenders.
Fillers to be used in acrylic PSAs are discussed in [131]. BaO, CaCO3, and
kaolin do not decrease tack, but ZnO and TiO2 do. Addition of zinc oxide
in formulations with carboxyl functional groups dramatically reduces the
pressure-sensitive character of the adhesive [132]. Normally zinc oxide is
used in PSAs for special tapes, improving the cohesion and reducing the
cold flow [133]. In carboxylated neoprene latex zinc oxide fulfills a dual
function: it acts as an acid acceptor to neutralize hydrogen chloride which
is formed, and it also crosslinks with the carboxyl groups on the polymer,
thereby increasing the cohesive strength. The fillers act as contact regulators,
pH regulators, crosslinking agents, and stiffeners according to their and the
adhesive’s composition; all these effects influence the peel. Crosslinking is
another way to achieve removability (i.e., through the choice of a self-
crosslinking PSA, or an external crosslinking agent).
This use of reactive fillers and the use of isocyanate as a crosslinking
agent for rubber-based PSAs is well known from adhesive practice; reactive
resins may be used also. Polyacrylic esters may be crosslinked by isocyanates
[87,134]. Crosslinking also may be carried out by more sophisticated,
built-in multifunctional monomers such as N-methylolacrylamide [102].
Polyacrylamide was used as a crosslinking agent for quite different polymers
[67]. Other functional polar monomers with carboxyl, amide, or hydroxyl
groups may be built-in [135]. Zinc oxide and zirconium ammonium
carbonate can be used as reactive metal derivatives [106,136]. It is possible
to obtain a more sophisticated structure, where a crosslinked polymer is
dispersed, like an inert filler in a PSA matrix [137].
Contact hindrance (i.e., the decrease of the adhesive contact surface) is
a general method to improve removability. It can be achieved by coating
only a minor portion of the substrate [67], coating it discontinuously
[138], coating it with a structure of adhesive/nonadhesive strips [139],
or with a mixture of adhesive/nonadhesive material [140]. In the latter
case the formulator has to design the adhesive blend, in the former cases
(discontinuous structures) only its rheology. As a special case of formulation
the use of raw adhesives having a built-in emulsifier (surface-active
groups) can be mentioned. Small amounts of vinyl-unsaturated emulsifier
monomers such as sodium-sulfoethylmethacrylate or salts of styrene
sulfonate may be used [141]. For repositionable, readhering labels
(like ‘‘Post It’’) and for protective films, both technical possibilities (i.e.,
coating of structured adhesive layers and the formulation ‘‘without’’
external surface-active agent) are used. As described in detail in [142]
partial fragmentation of a rigidized adhesive surface, pyramidal PSA
coated sites, or expandable fillers were also tested in order to avoid adhesion
494 Chapter 8

build-up for removable adhesives. Microspheres used in the formulation of


PSPs are listed in [143].
Another field in the formulation of removable adhesives concerns
the design of primer coatings. Peelable pressure-sensitive adhesive-coated
laminates are comprised of a face material with a primer coating (which
includes a contact cement) and the PSA contains the contact cement as well.
The use of a contact cement as a primer and in the adhesive enhances
cohesion of the adhesive coating and anchorage to the face material which
in turn improves peelability. Generally for a contact cement, a harder
elastomer may be used [144] (e.g., CSBR for natural rubber), a harder (high
melting point) tackifying resin, or the same adhesive in a more crosslinked
status. The primer coating (with a quite different or the same adhesive
composition) and the PSAs may contain quite different additives (for the
coating). The primer coating is applied to label grade paper and dried to
form the first layer. The peelable PSA is coated onto a paper release liner
with a cured silicone release layer, and than dried. The primer coating and
the adhesive-coated liner are subsequently laminated together.
Polyisocyanate and polyurethane derivatives used as formulating
additives for removable compositions may play quite different roles. In
general, they are used as crosslinking agents, but it is possible to apply them
as primers where they will also act as crosslinking agents. In this case
the isocyanate polymerizes first with reactive groups in the face material
to create a strong chemical bond in addition to conventional physical
adhesion. The remaining isocyanate reacts with moisture and with the PSAs.
The best results are achieved using solvent-based polyisocyanates. Disper-
sion-based polyisocyanates penetrate and break through (migration);
solventless ones do not penetrate enough and increase the stiffness of the
face material too much. The formulation of primers is discussed in detail in
[145] and [146].

Paper and Film Labels


General Purpose Permanent Paper Laminates. Permanent paper
laminates require tackified formulations where the tackifier level depends
on the sheet/roll nature of the laminate. Dispersion-based tackifiers should
be used and, as a function of the base PSA and face, rosin esters (acrylic),
hydrocarbon (acrylic, CSBR), or rosin acid resins (acrylic, CSBR) should be
selected. For light release (solventless liner) a high wetting agent level should
be used, mainly on a sulfosuccinate basis; the level of defoamers should be
limited. The release force depends on the direct or transfer coating method
used. For aggressive paper labels (roll material, for labeling machines) high
tack, high peel adhesion, and low controlled release is required. For sheet
Manufacture of PSAs 495

material less tack, but high peel, cohesion, and excellent die-cutting
are required. Time/temperature stability of the adhesive properties also
is required. Tackified acrylic adhesives meet these requirements; both
solvent-based and water-based PSAs can be used. Tackified rubber-resin
solvent-based adhesives are adequate for roll/sheet material also; tackified
EVAc-based PSAs are more suited for sheet laminates.

Removable Labels. The guidelines for the formulation of removable


PSAs were outlined earlier. The main requirements can be summarized as
follows: low peel, good removability, low migration, no bleeding, and high
tack. Crosslinked solvent-based acrylic PSAs and special primed (or
primerless) water-based acrylic PSAs meet these requirements; solvent-
based rubber-resin PSAs and water-based PSAs meet those same require-
ments for other self-adhesive products (e.g., protective films, etc.).
Nontackified or slightly tackified formulations should be used; plasticizers
as tackifier are preferred. Crosslinking agents and peel modifiers should be
used, with low levels of wetting agents (sulfosuccinates are not preferred).
Direct coating appears to be the preferred coating method. Screening
formulations for removable labels are presented in [147].

Permanent Film Laminates. These films are engineered to be durable,


conformable, dimensionally stable, and able to withstand severe weather
and handling conditions. They can be processed by screen printing, roll
coating, steel die-cutting, thermal die-cutting, and premasking. High
transparency, high water resistivity, and dimensional stability are required.
Roll and sheet materials are coated with solvent-based and water-based
acrylic PSAs. The market segment for high polyethylene adhesion is
partially covered by ethylene copolymers. Mostly nontackified formulations
are used; the level of dispersion additives should be limited. Sulfosuccinates
are not preferred; wetting agent free, thickened formulations (if possible) are
suggested.

Removable Film Laminates. Low peel, good removability, high tack,


good transparency, dimensional stability, and water resistivity constitute
the performance requirements, which are met by solvent-based acrylic and
water-based PSAs. Plasticized formulations are possible. A low level of
dispersion additives should be used (direct coating is proposed); crosslinking
agents may be included in the formulations.

PVC-Based Laminates. A special case of adhesive formulations


concerns coatings onto PVC. In order to understand the principles of the
formulation of PSAs for PVC coatings, the requirements for PVC labels will
be studied first (except for plasticizer migration). The properties of adhesives
496 Chapter 8

Table 8.23 Adhesive Properties of a PSA-Coated PVC Label Material

Adhesive Properties Values, Application

Property Method Substrate Comments Sheets Roll

Peel adhesion 180 Glass N/25 mm


Immediate 7 10
20 min 8 12
1 hr 8.5 14
24 hr 10 16–20
Stainless steel Immediate 6 –
20 min – –
1 hr – –
24 hr – –
PE film Immediate 4 5
20 min 5 6
1 hr 5 –
24 hr 6 8–12
Tack Rolling ball cm 2.5–3.5 –
Loop tack Glass N/cm 6–10 10
PE N/cm 4 6
Shear Hot shear Stainless steel min 60 20

for coating onto PVC include the adhesive performance characteristics


(Table 8.23), water resistance, shrinkage or dimensional stability, optical
appearance, aging, processability, and FDA or BGA approval. It is
apparent that different peel, tack, and shear values are required for sheet
or roll material applications.
In Table 8.24 the other characteristics of PSAs used for film labels
(e.g., water resistance, shrinkage, and optical appearance) are summarized.
In order to clarify the importance of the given values, a more detailed
discussion of these properties follows. First the water resistance is examined.
Water resistance encompasses the stability of the adhesive and end-use
properties after immersion in water. For certain practical uses, the labels
(label sheets) will be immersed in water in order to achieve a partial and
temporary loss of the tack for a better repositioning. Clear, transparent
labels should not suffer whitening under these conditions and the loss of
tack should persist for a short time only in order to achieve rapid adhesion
onto the final substrate. Consequently one should distinguish between label
sheets (hand applied, voluminous film coatings), and labels (roll or sheet
material, applied by high speed labeling machines, labeling guns, or by
hand). For label sheets affixed mostly on polar surfaces, water-whitening
Manufacture of PSAs 497

Table 8.24 End-Use Properties of PSA-Coated PVC Label Material

Water Resistance

Property Method Values Units

Water-whitening Subjectively > 10–17 sec


< 3–8 %
Loss of transparency Colorimetry,
after 7 min
Wet adhesion Subjective push
away after:
on glass 1 min
5 min
10 min
30 min
60 min
Squeeze bottle Subjectively No wings, lifting,
wet anchorage after 24 hr high peel adhesion,
subjectively
Shrinkage

Properties

Position Method Values Units

Unmounted 7 days, 60 C


clear PVC MD 0.9 CD 0.6 %
white PVC MD 0.5 CD 0.4 %
Mounted 7 days, 60 C
Clear PVC MD 0.5 CD 0.4 %
White PVC MD 0.2–0.3 CD 0.15 %
Optical Appearance

Property Method

Coagulum No grit
Optical appearance Subjectively
Optical clarity Colorimetry
MD, machine direction; CD, cross direction.

(loss of transparency) and wet adhesion are the most critical characteristics
(Fig. 8.12).
For labels affixed at high speed, mostly on polyolefin bottles, wet
adhesion and squeeze bottle anchorage are the most important properties
498 Chapter 8

related to water resistance. There is no correlation between water-whitening


and the other water resistance-related properties, but there is a correlation
between wet anchorage and wet adhesion on glass or plastics (after
immersion in water or diluted solutions of detergents, over a period of time).
Wet anchorage is the adhesion of the PSAs onto the face stock (film)
(Fig. 8.13 and 8.14). Hence, wet adhesion includes two separate aspects.

Figure 8.12 Water resistance of label sheets. Interdependence of the water


resistance performance with the adhesive characteristics.

Figure 8.13 Water resistance of PSA labels. Wet anchorage and wet adhesion,
where and how they work.
Manufacture of PSAs 499

Figure 8.14 Water resistance of PSA labels. Interdependence of the water


resistance parameters, peel, and tack-related parameters.

First, one observes the adhesion of the wet PSA coating on the substrate,
and fast (or slow) increase of the adhesive forces between substrate and
coating, resulting from the drying of the coating (over a short period of
time/minutes). Testing occurs by trying to move the label parallel to the
surface. Next, the adhesion of the PSA coating on the substrate is measured,
the substrate with the affixed label being subjected to immersion in water (or
water and surface-active agent) for hours, or days. Testing occurs by peeling
the label from the substrate surface.
Consequently, wet adhesion on the substrate tested through ‘‘push-
away’’ is a bonding wet adhesion (BWA) test. Wet adhesion tested through
peeling off is a debonding wet adhesion (DWA) test (Fig. 8.15). Testing of
the wet adhesion-related product performances will be described in detail in
Chap. 10 discussing the test methods for PSAs.

Formulating for Special Uses


Aging. The synthetic and natural polymers utilized in the adhesives
industry are frequently susceptible to degradation and this degradation is
markedly accelerated by interaction with oxygen at elevated temperatures as
well as with light. On the other hand, warranty times of at least 2–4 yr are
required for quality labels [148]. There are several practical data showing the
influence of the formulation on the aging of PSAs. Adhesives working under
extreme temperature and humidity or UV light require aging resistant
500 Chapter 8

Figure 8.15 Water resistance of PSA labels. Test of the wet adhesion.

formulating components. In some cases (film coating) water-white UV-


stable PSA formulations are needed [33]. For special uses, like cereal inserts,
water-based PSAs with low edge ooze and adhesion bleed that comply with
US Food and Drug Administration (FDA) regulation 175.105 are required.
Such formulations need special base elastomers and surface-active agents.
Any adhesive composition must include an antioxidant to inhibit oxidation
of the tackifying agent and consequent loss of tackiness as the adhesive
composition ages [10]. The saturated hydrocarbon-based long-chain
structure of acrylic PSAs provides good chemical stability towards aging
through oxygen and light. Hence acrylic PSAs can be used without
protecting agents for opaque as well as for transparent face stock. Rubber-
based PSAs are inadequate for transparent webs. Thus there is a general
trend to give several years warranty for these laminates.
Tackified formulations show less stability upon aging. Tackified,
water-based acrylic PSAs can interact with special face materials (e.g.,
thermal paper) giving rise to the partial or total loss of the adhesive
properties. Tackified film laminates may show discoloration. The most
important problem is the frequent interaction between metal ions from the
paper (face stock) and acid rosin tackifiers causing the loss of tack and peel
adhesion. This phenomenon influenced the formulators to change the resin
base from acid- to ester-based ones, although the latter show less tack and
peel adhesion. In this context hydrocarbon-based tackifiers can compete
with the less aggressive rosin esters. As can be seen from the DSC data in
Fig. 8.16 differences in Tg for fresh and aged material are very pronounced
(35.5 C versus 1.5 C respectively). This is probably related to the
Manufacture of PSAs 501

Figure 8.16 The increase of the Tg during aging for a PSA formulation using an
insufficiently stabilized acid rosin as tackifier. The shift of the DSC peak to positive
values for the aged adhesive (B).

polymerization of the resin. The exact nature of the phenomenon is still


unknown.
Abietic acid (in rosins) contains a conjugated double bond and will
react with oxygen. This will increase the softening point of the resin, and the
subsequent tack of the adhesive will decrease with time. Development work
was carried out to improve the aging stability of hydrocarbon-based
elastomers and tackifiers. The unsaturation that exists in the diene
midblocks is responsible for the aging of SBS. The experience with
PVA-stabilized carboxylated neoprene latexes indicates that they are more
environmentally stable than conventional neoprene [149]. On the other
hand, the structure and properties of SBS block copolymers are nearly ideal
for PSA formulations since they combine high cohesive strength and good
aging characteristics. The importance of the compatibility between PSA
and tackifier was demonstrated earlier. Compatibility influences the time/
temperature stability of the end-use properties; it also is influenced by the
chemical nature of the resin. For example, unsaturation of the resin will be
eliminated by disproportionation and hydrogenation. Hydrogenation brings
more compatibility and aging resistance. Practically, the choice of an
adequate base elastomer and tackifier, together with the use of antioxidants
and UV-stabilizing agents, impart a good aging stability.
502 Chapter 8

There are many components with a well-known chemical composition


that are used as antioxidants for adhesives. Auto-oxidative degradation can
be reduced by heat stabilizers: primary and secondary antioxidants. Primary
antioxidants are free radical scavengers. They deactivate the aggressive
peroxy radicals and the new build-up of radicals. Secondary antioxidants
reduce the hydroperoxides formed during oxidation. (A detailed discussion
of the chemistry of antioxidants is given in [150].) According to theory,
secondary antioxidants (which are less expensive) should be added at double
or greater the level of primary antioxidants. Other specialists prefer a higher
level of primary antioxidants to avoid deficiencies associated with secondary
antioxidants. The most important primary antioxidants are the hindered
phenols [10], ortho-substituted or 2,5-disubstituted hindered phenols, in
which each substituent group is a branched hydrocarbon radical, with 3–20
carbon atoms (e.g., tertiary butyl or tertiary vinyl phenol). Other hindered
phenols include para-substituted phenols where the substituent group is
OH, R being methyl, alkyl, etc. Among the sulfur-containing organometal
salts, the nickel derivatives of dibutyl-dithiocarbamate are used. Hindered
phenols or sulfur-containing organometal salts are preferred for crosslinked
acrylic PSAs, dibutyl-dithiocarbamate may be used for thermoplastic
rubber-based hot-melt PSAs. Oxidation sensitivity depends on the formu-
lation, on the tackifier level, and on the nature of the PSA. The effect of the
antioxidants is a function of the type of aging, nature of additives, curing
systems, chemical structure of the polymer, and environmental conditions.
Thus at low resin content (for SIS-based PSAs) there is little or no tack
retention when exposed to air; however, the tack retention is increased if the
resin level is increased to 60%.
UV stabilizers are 2,4-dihydroxydibenzophenones, substituted hydro-
xyl-phenyl-benzoates, octyl-phenyl-salicylates, and resorcin-monobenzoles.
Stabilizers used for block copolymers are sterically hindered phenols,
thiotriazines, trimethyldihydroxy-chinolin polymers, alkylated bisphenol;
for PSAs the best would be Zn-dibutyl-dithiocarbamate. Hot-melt adhesive
compositions generally contain 0.2–2% antioxidant. Antioxidants for hot-
melts should fulfill the following requirements: low volatility to provide
better maintenance, to reduce viscosity changes during processing, to delay
skin formation, and not to impart colour. Among the preferred stabilizers
or antioxidants utilized are high molecular weight hindered phenols and
multifunctional phenols such as sulfur- and phosphorus-containing phenols.
Hindered phenols may be characterized as phenolic compounds, which also
contain sterically bulky radicals in close proximity to the phenolic hydroxyl
group. (From this class BHT, i.e., 1,6-di-t-butyl-4-methylphenol, exhibits
excellent processing characteristics but pronounced volatility, which limits
its application.) The performance of these antioxidants may be further
Manufacture of PSAs 503

enhanced by utilizing them in conjunction with known synergists such as


thiodipropionate, esters, and phosphites [120]. For removable hot-melt
PSAs that are based on styrene/butadiene multiblock copolymers, 0.3%
Irganox 565 (Ciba Geigy) and Polygard (Uniroyal) were proposed [99].
Furthermore it was shown that after 7 days at 70 C an SIS rubber loses
99.5% of its tensile strength, with antioxidant of only 10%.
A preferred antioxidant for natural rubber-based formulations is
available under the trade name Heveatex B 407A antioxidant, which is a
ball-milled dispersion of 4,4-butylidene bis powder in an aqueous base. A
butylated reaction product of paracresol and dicyclopentadiene and a liquid
carrier such as ditridecylthiodipropionate (Wingstay L by Goodyear) or an
aqueous dispersion containing zinc dibutyl-dithiocarbamate were proposed
as nonstaining, nondiscoloring antioxidants [67]. Oxidation of neoprene
yields hydrogen chloride, which in the presence of moisture is converted
to hydrochloric acid. Neoprene may be protected by nondiscoloring,
nonstaining antioxidants of the hindered bisphenol class (like Antioxidant
2246 or Wingstay). A 50% dispersion of Wingstay L under the trade name
Bostex 24 may be used for aqueous PSA dispersions. For these 0–3% w/w
DDA-EM 50%, KSM-EM 33%, or Vulkanox TD-EM (Bayer) may be used
also. Suitable antioxidants for polyvinylethers include Ionox 330 (Shell),
Irganox 2010 (Ciba Geigy), or ZKF (Bayer).
The aging of SBS and SIS rubbers is usually noticeable by the loss
of tackiness and cohesive strength, and it is reasonable that some of these
changes may be related to the incompatibility of the two-phase system.
However, it is more likely that the unsaturation that exists in the diene
midblocks is responsible for this instability [151]. Generally, for SIS-based
hot-melt PSAs, amine- and phenol derivatives are used as chain stabilizing
agents (antioxidants). Phosphites and thioesters are used as ‘‘killers’’ for
peroxides. Oxybenzochinones, oxybenzotriazoles, and nickel chelates are
used as UV stabilizers [152].
As can be seen from these examples, there are many antioxidants
available with different structures, formulations, and trade names. Their
choice is made on the basis of their working mechanisms, their dispersability,
and on the basis of economical considerations also. The most important
criteria in the choice of antioxidants and UV stabilizers is, beside their
efficiency, their inert character towards the face stock and the printing
materials. Nonstaining compounds are suggested only. Their selection is
made mostly on the basis of practical tests (see Section 2.12 and Chap. 10
also). The main antioxidants and their suggested use are listed in [93].

Water Resistance. Wash off labels, splicing tapes, certain medical


labels etc., need good water solubility; labels for soft drink containers need
504 Chapter 8

time-dependent water solubility, whereas deep freeze labels, special medical


labels, and labels for outdoor use and for textiles need no water solubility
(i.e., the water resistance should be adjusted as a function of the intended
end-use). Recycling of paper-based pressure-sensitive products requires
water solubility too. Water resistance is a function of the nature of the base
polymer, the particle size and particle size distribution, the surface-active
agent nature and content, and the coating and converting conditions.
Satisfactory results can only be obtained with crosslinked products. One-
component products with built-in crosslinking capabilities are required. A
high degree of crosslinking, however, cannot be achieved for PSAs without
the loss of the pressure-sensitive performance characteristics. On the other
hand, high tack/peel adhesion are very important for both debonding
resistance and bonding velocity under humid circumstances. As discussed
in [153] wet tack is a special requirement for several products and can be
achieved by an apparent adhesivity due to the rheology of the dispersed
system (Fig. 8.7). Such systems include solubilizers or are based on
hydrogels (see Section 2.12 also). Humidification agents (for the stabilizing
of the water content of the adhesive) are added too. The role of the
solubilizer can be played by the main elastomer (e.g., vinyl pyrrolidone,
pyrrolidone-vinylacetate copolymers, polyethyloxazoline, polyamides,
ethylene-vinylacetate copolymers etc.) or by special additives (e.g., fatty
acids, protective colloids, surfactants, water soluble waxes etc.). Such
additives do not have a pressure-sensitive character. Therefore the
formulation of water-resistant or water-soluble PSAs is always associated
with the loss of the main adhesive properties (i.e., water-resistant/soluble
PSAs do not perform as well as common PSAs for permanent applications).
The final performance of the PSAs depends on the skill of using an adequate
ratio of solubilizer (being less tacky) either built-in (base polymer) or added
(wetting agent), tackifier (danger of coalescence) needed for wet tack, and
chemical base. The most important test criteria and characteristics are
discussed in Chap. 10. Solubilizers are listed in [93].

Thermal Resistance. Formulating for thermal resistance requires the


use of temperature-resistant elastomers and/or tackifiers. Chloroprene
latexes display good thermal resistance. Self-crosslinking, acrylonitrile-
containing latexes were suggested also. Reactive, high melting point resins
and/or crosslinking agents also may be used.

Deep Freeze PSAs. Special demands are placed on the chemical


composition and physical properties of low temperature PSAs (i.e.,
adhesives intended for use at relatively low temperatures). Often PSAs
which have adequate cohesive and adhesive strength at low temperatures are
Manufacture of PSAs 505

so gummy at ambient conditions that they complicate adhesive handling at


ambient temperatures and the manufacture of adhesive-containing articles.
Such gumminess also causes creep and bleedthrough on labels and other
PSA-coated face materials.
Theoretically, there are two different formulating approaches in order
to achieve good deep freeze properties. The first one is the choice of very soft
elastomers (acrylic PSAs) with no need for tackifying resins. The second
consists in the use of low molecular weight liquid tackifiers. Both
approaches can be worked out more easily using solvent-based adhesives;
better results are obtained without tackifiers. For deep freeze adhesives, the
application temperature is 5 to 25 C, the service temperature ranges from
10 to 30 C [154]. PSAs should have a lower Tg than their application
temperature. Such soft PSAs often give rise to edge oozing, bleeding, and
converting problems. For rubber-based formulations, measurement of peel
adhesion and polytack at 5 C and 20 C provides a reasonable indication
of the final performance, both for low temperature (i.e., chill) and for deep
freeze applications. As the butadiene level increases beyond 80% the
suitability for deep freeze applications improves. Gel content appears to
have a diminishing influence as temperature decreases [53]. Wet tack
formulations are similar to those for low temperature applications [155].

2.12 Influence of Adhesive Technology


The state of the adhesive (liquid or molten) limits the use of tackifiers or
other formulating additives. Generally, a good mixing of the components
supposes a liquid state (solution, dispersion, or melt). Solvent-based
adhesives are tackified with resin solutions, but water-based ones also
have been successfully tackified with resin solutions. Latex-resin dispersion
mixtures were developed. The use of solid (molten) resins is limited by the
thermal stability of the water-based PSAs. Chloroprene dispersions are
sufficiently stable to be tackified with molten resins. Molten resin (melting
point < 90 C) may be fed into aqueous dispersions of vinyl ether and acrylic
PSAs also [155].

Solvent-Based Adhesives
Solvents are used as temporary components of the adhesives, primers,
release coatings, and printing inks. As the carrier solvent may influence the
stability, the adhesive properties, and the coating properties of the solvent-
based adhesives, it appears necessary to separate the formulating of the
uncoated and coated adhesive.
506 Chapter 8

In order to obtain adequate coating properties the solution polymer


molecular weight and concentration must be balanced [156]. High molecular
weight elastomers give viscous solutions. Therefore, the solids content of
solvent-based formulations is generally less than 25%. Higher solids content
is achieved for natural rubber-based (35%), SBS-based (20–50%), or SIS-
based (20–40%) formulations [147]. Technical problems arise when coating
highly viscous solutions and drying low solids formulations. The main
components of the adhesive formulation have to be dissolved before or
during blending. Mastication of natural rubber may facilitate the dissolu-
tion process. The choice of a suitable solvent is controlled by such factors
as solubility, viscosity, flammability, toxicity, and costs. The solubility
parameters provide the means to choose the proper solvents. So-called
green solvents are those not included in the VOC and TRI (Toxic Release
Inventory). The most commonly used solvents include ethyl acetate,
n-hexane or n-heptane, toluene, special boiling point petrol, acetone, and
isopropyl acetate. Their efficiency, drying speed, and economical and
environmental properties are the main criteria for this choice. Their recovery
possibilities have also to be taken into account [157–159]. The overall
performance of solvent-based formulations in comparison with hot-melt
and water-based PSAs will be discussed in Section 2.14.
The most important coating characteristic of solvent-based formula-
tions remains the viscosity which depends on the base elastomers and on
the solvent used. For special cases other important parameters were
identified. The viscosity of PSA solutions may be reduced using dispersed
rubber (CSBR, neoprene, or polybutadiene) together with dissolved
rubber (nitrile rubber, NC; butyl rubber; polyisobutylene; ethylene
butylene diene rubber, EBR; polyisoprene) in an isoparaffinic solvent.
Viscosity adjustment for solvent-based rubber-resin adhesives is achieved
with toluene/hexane, where toluene is a solvent for polystyrene blocks in
SBS and hexane is a solvent for the elastomeric domain. Other solvents
(i.e., heptane, octane, petrol) are similar to n-hexane; cyclohexane and
methyl isobutyl ketone (MIBK) are like toluene [17]. For natural rubber-
based formulations heptane, hexane, blends of hexane/toluene, and
hagasoline were suggested; for SBR-based formulations toluene, heptane,
heptane/ethanol blends, or gasoline are preferred. Butyl rubber is solved
in toluene and heptane; for SBCs toluene, toluene/heptane, or petrol/
toluene were suggested [147,160]. The viscosity of rubber-resin solutions
decreases with the increase of the resin content [156]. Pressure-sensitive
adhesives using the less photochemically active ethyl acetate instead of
toluene or xylene are recommended for clear film labels. The main
solvents used for PSAs based on different elastomers are discussed in
detail in [160].
Manufacture of PSAs 507

Generally, the manufacture of solvent-based PSAs (i.e., manufactur-


ing the adhesive solutions) is a process of dissolving the main adhesive
components (rubber and resin) and adding other solid or liquid
components. This is a simple procedure, where dissolving the elastomer
takes more time and needs, in some cases, special dissolvers (homogeniza-
tion and heat transfer) and a programmed feed-in of the materials. Swelling
and dissolution of the elastomer are accelerated if the rubber-solvent blend
is processed in a sigma blade mixer. Rubber can be cut and rolled in a jar
mill also. Formulations which contain fillers are first milled to disperse the
solid material, sheeted off, and cut before solving. Thermoplastic elastomers
are manufactured as pellets, therefore they do not need previous cutting.
Low and medium viscosity rubber-resin solutions are manufactured in a
vertical high speed stirrer.

Hot-Melt Adhesives
Hot-melt adhesives are 100%-solid materials which do not contain or
require any solvents. For such adhesives low molecular weight migrating
components (e.g., surfactants) or detackifying components (e.g., surfactants,
defoamers, etc.) are not necessary. They are solid materials at room
temperature, but upon application of heat melt to a liquid (fluid state) in
which form they are applied to a web. Upon cooling the adhesive regains its
solid form, and gains its cohesive strength. In this regard hot-melt adhesives
differ from other types of adhesives which achieve the solid state through
evaporation or removal of carrier liquids (solvents, water) or by
polymerization. Hot-melt PSAs are carrier-free systems where formulating
is necessary in order to improve the adhesive, coating, and end-use
properties, mostly by changing the viscoelastic properties of the pure, bulk
components. Normally no coating additives are used.
Hot-melt adhesives were developed in the 1940s. In the 1960s block
copolymers for hot-melt were introduced [34]. This group of adhesives
has developed particularly rapidly since the early 1970s because of their
nonpolluting character thanks to the absence of solvents and the low energy
input required to manufacture them. In addition thicker adhesive layers
of up to 100 g/m2 can be applied by hot-melt coating instead of using
solvent-based PSAs [161,162]. Much attention is given to the proper
selection of the polymer backbone component of hot-melt PSAs. While the
backbone polymer determines cohesive strength, flexibility, viscosity, and
heat resistance for an adhesive system, the tackifier determines the color of
the hot-melt PSA, as well as its wettability and thermal stability. Tackifiers
also determine the entire system functionality, greatly influencing overall
performance. The suitability of a block copolymer for removable PSAs is
508 Chapter 8

heavily dependent on other ingredients used in the adhesive formulation,


such as tackifying resins and processing oils (see Chaps. 5 and 6).
A detailed description of the characteristics of the raw materials for
hot-melt adhesives was given by Fricke [163], King [164], and Jordan [165].
The cohesion determining components are the base polymers; for hot-melt
PSAs particularly, these are importantly the SBS and SIS block copolymers
and polyisobutylene, but also the rubber-like polymers such as natural
rubber, CSBR rubber, acrylonitrile copolymers, polychloroprene, etc. To
increase tackiness, different tackifiers are added as adhesion determining
components. The resins used are either natural resins such as colophonium
or synthetic resins such as coumarone resins, hydrocarbon resins of
different grades and others, some of which are thermoplastic and have
softening points of about 70–120 C. Additional adhesion promoting
substances are plasticizers, which are used (particularly for permanent
adhesive formulations) with rubber-type base polymers. In order to obtain
exactly defined mechanical, physical, and chemical properties in the adhesive
layer, fillers are added. The addition of stabilizers protects the polymers from
degradation during processing and from damage caused by heat and
oxidation.
Styrene-isoprene block copolymers provide a long open time for hot-
melt formulations, while SBS is preferred for better cohesion. New triblock
copolymers PS-PEB-PS contain a soft segment of PEB in a polystyrene
matrix (30% PEB and 70% PS), whereas S-EB-S copolymers display better
compatibility with paraffinic oils [166]. Generally, a hot-melt PSA recipe
consists of a thermoplastic rubber, an endblock compatible resin, a midblock
compatible resin, oil, and an antioxidant. With the choice of the resin for
hot-melt PSAs, one has to separate the influence of endblock and midblock
compatible resins. The main functions of midblock resins comprise the tack,
improving the adhesion, and their function as a processing aid. New EVAc
segment polymers were developed for hot-melt PSAs [64]. These products
display better environmental stability. Their recipes do not contain waxes
because of their mutual incompatibility. These formulations have a viscosity
of 9,000–56,000 mPa.s (at 150 C). Developments in the synthesis of styrene
block copolymers, which allowed the manufacture of products with various
sequence distributions (diblock/triblock), with different styrene derivatives as
endblocks, with various structures (linear, radial-branched, or multiarm)
and with functionalized endblocks, strongly enlarged the versatility of such
raw materials for HMPSAs. The main thermoplastic elastomers used for
HMPSAs are discussed in detail in [167] and [168].

Tackification of Hot-Melt PSAs. To gain useful PSA properties,


thermoplastic rubber, which has almost no intrinsic adhesive character, is
Manufacture of PSAs 509

compounded with tackifying resins and sometimes plasticizers. These are


low molecular weight materials, often unsaturated and with an aliphatic
character, such as rosin esters, polyterpene resins, and C5 hydrocarbon
resins. These resins tend to concentrate in the continuous elastomeric phase
of the thermoplastic rubber, thereby swelling and softening the rubber. Hot-
melt PSAs based on styrene/butadiene multipolymers were designed using
25% rubber, 34–44% resin, and 30% processing oil [99] (for the suggested
tackifier level as a function of the elastomer nature see Chap. 6, Section 3.2
also). The addition of high melting point polystyrene associating resins
improves the heat resistance of hot-melt PSAs based on multiblock
butadiene/styrene elastomers with no or slight decrease in tack [78].
Plasticizer oils should lower the modulus, and improve the tack and deep
freeze properties of hot-melt PSAs (see Section 2.7 also). They also lower the
melt viscosity. Ideally, a plasticizer oil is not soluble in the polystyrene
domains, is fully soluble in the rubber segments, has low volatility, low
specific gravity, and high storage stability [162]. Adequate plasticizer oils
contain less than 2% aromatic fractions. Soft resins or elastomers may be
used as plasticizers also.
A screening recipe (formulation) for a hot-melt PSA designed on the
basis of the literature should contain [169–171]: 100 parts SIS, 100 parts
aliphatic hydrocarbon resin, and 5 parts antioxidant. In order to improve
the tack and to lower the viscosity, a plasticizer oil should be added to the
recipe: 100 parts SIS, 100 parts resin, 40 parts plasticizer oil, and 5 parts
antioxidant. In order to reinforce the polystyrene domains, a high melting
point coumarone-indene resin should be added to the recipe as well: 100
parts SIS, 100 parts resin, 40 parts plasticizer oil, 60 parts high melting point
resin, and 5 parts antioxidant.
With a polybutene used as plasticizer and a high melting point linear
homopolymer of -methyl styrene as reinforcing resin, the formulation
becomes [172]: 100 parts thermoplastic elastomer, 150 parts tackifier, 2 parts
antioxidant, 125 parts polybutene, and 40 parts reinforcing resin.
It remains difficult to formulate removable hot-melt PSAs. Current
removable adhesives supplied for label stock are acrylic latexes and solvent-
based formulations. Both of these materials have high molecular weight
polymers that reduce flow on a surface to prevent build-up of adhesion. In
contrast, hot-melt PSAs are based on materials having lower molecular
weight components that make reduced flow or wetting on a surface very
difficult. Certain formulations containing a SIS block copolymer, a low
softening point highly aliphatic resin, and a metal salt of a fatty acid exhibit
a good balance of properties and are used as removable PSAs [120].
Screening formulations of removable HMPSAs based on SIS, APP, or
EVAc are presented in [173].
510 Chapter 8

Stabilizers in Hot-Melt PSAs. Adhesive formulations based on


unsaturated synthetic elastomers and tackifiers are highly sensitive to
oxidation. These adhesives thus require the addition of stabilizers to protect
them from oxidative degradation during preparation, storage, and end-use
(see ‘‘Formulation for Special Uses’’ also). While the use of a nitrogen
blanket can offer some protection during mixing, further significant
protection against oxidative degradation of hot-melt PSAs can be obtained
with stabilizers. A hot-melt PSA formulation may contain several
components, including a base rubber, tackifier resin, and oil. Oil is the
most stable of the viscous components used in the adhesion formulation
[174]. The presence of stabilizers in the rubber and in the tackifier can
significantly improve the performance of the materials over the unstabilized
formulation. Antioxidants such as hydroxycinnamate derivatives and
hydroxybenzylbenzene derivatives were suggested.
A blend of rubber with tackifier resin (without oil) responds well to the
stabilizers present in the rubber as well as in the tackifier resin. However, the
stability of the adhesive formulation is significantly lower than that of its
components [175]. Thermoplastic rubbers used in hot-melt PSA formu-
lations, like the other ingredients in the adhesive formulation, must be
stabilized against oxidation attack. Phenol derivatives as antioxidants are
used for hot-melt PSAs; zinc dibutyl-dithiocarbamate also may be used [70].
Antioxidants for hot-melt PSAs are Tenol BHT (Eastman), Antioxidant 80
(ICI), Santonox R (Monsanto), or dilaurylthiopropionate and n-butyl-
hydroxy toluene. Other derivatives such as substituted phenols, phenotria-
zines, and carbamide or ortho-toluylbiguanidine are suggested [169]. For
electron beam-cured hot-melt PSAs (e.g., Kraton D 1320 X) Irganox 1076
is used as antioxidant [176].

Viscosity of Hot-Melt PSAs. A hot-melt adhesive must possess a


sufficiently low viscosity to enable flow at melt temperatures. Traditionally
the adhesive property that suffers the most in this instance is shear. Curing
of the hot-melt adhesives greatly improves shear properties. While thermal
cure or crosslinking is feasible, radiation curing by electron beam or UV
light is preferred.
Hot-melt PSAs possess viscosities of 20,000–100,000 mPa.s [177],
respectively 50,000–250,000 mPa.s [178]. Hot-melt PSAs exhibit a viscosity
of 11,000–16,000 mPa.s at 160 C [179]. Low viscosity hot-melt acrylic PSAs
possess viscosities of less than 50,000 mPa.s at 178 C [24]. Viscosities for
hot-melt PSAs based on styrene/butadiene midblock copolymers are
situated between 4000 and 17,800 mPa.s at 120 C and 600–1100 mPa.s
at 180 C. For high quality hot-melt PSAs a low processing viscosity at
150–175 C and high peel strength at room temperature are required [180].
Manufacture of PSAs 511

Hot-melt PSAs based on thermoplastic rubbers use temperature and


not solvents to reduce the application viscosity; the addition of plasticizers
such as oils or low molecular weight resins is necessary to lower the viscosity
of the blend to practical levels. Polybutenes also have been used successfully
as tackifiers/plasticizers in hot-melt PSAs. The choice of an adequate
resin as tackifier for hot-melt PSAs also may influence the viscosity.
A noncrystallizing rosin as low viscosity tackifier in hot-melt adhesives
(providing it is stabilized against crystallization and oxidation) offers good
adhesive performance characteristics [179].

Single Component Hot-Melt PSAs. The proper selection of como-


nomers and synthesis additives has led to the development of olefinic
copolymers that are PSAs when unformulated [181]. Because these hot-melt
PSAs are pure olefin copolymers they offer several advantages over typical
styrene block copolymer/oil/tackifier blends. Those include the lack of oils
which can bleed into face materials, less possibility of skin sensitivity
problems (which is generally associated with the tackifiers in conventional
hot-melt PSAs), a good thermal stability, and a low color and odor. Other
advantages are the ability to be chemically modified by reaction with
-olefin polymers, the ability to be blended with most olefinic compatible
adhesive ingredients for general applications, and a lower density than
conventional rubber-based hot-melt PSAs, resulting in increased durability.
Developments in acrylic polymer technology have led to a new family
of phase-separated acrylic block copolymers. The thermoplastic elastomer
characteristics exhibited by these block copolymers were utilized for
hot-melt PSAs applications [24]. Acrylic hot-melt PSAs possess excellent
thermal, oxidative, and photooxidation stability relative to the competitive
hot-melt PSAs and solvent-based products. A range of modifiers (rosin,
terpene, phenols, hydrogenated rosin esters, etc.) also may be used [182].

Manufacture of Hot-Melt PSAs. The entire operation of compound-


ing hot-melt PSAs from the raw materials to the coating of the label or tape
is divided into two heat-treatment stages: compounding and coating. In
order to limit the excessive thermal history and oxidation of the materials
there are three possibilities: reducing the residence time during compound-
ing, running the coating in-line with the compounding operation, or keeping
all polymers from the compounding stage to the coating line in a closed
system or under inert atmosphere to avoid oxygen contact at the lowest
possible temperature [162,183,184].
With the sigma blade mixer, which is the preferred mixer type for
small to medium sized plants, a semicontinuous operation is possible by
using the mixer with a discharge screw and a process control system. For
512 Chapter 8

medium to large batch sizes, however, compounding on an intermeshing


co-rotating twin-screw mixer is the compounding method which leads to the
best quality products. Some of those plants have been in operation for
several years and are capable of compounding at rates of over 1000 kg/hr.
This technology is similar to that used for compounding plastics, but
the feeding is more complicated on account of the many components
required. The requirements for compounding can be summarized as
follows [162]:
mastication or plastification of the polymers through shearing
and heat without degradation through oxidation;
melting of the tackifiers and homogeneous incorporation into the
polymer matrix;
homogeneous incorporation of fillers with good dispersion of the
filler particles;
homogeneous incorporation of plasticizers, waxes, antioxidants,
and stabilizers;
facilities for creating vacuum to ensure freedom from pores and
remove remaining moisture from the fillers;
constant temperature of the finished adhesive compound, so that
the required viscosities can be adjusted precisely and reproducibly.
In the mixing cycle the overriding principle is that of obtaining a
completely homogeneous mix in the shortest possible time at the lowest
possible temperature (i.e., with a minimum thermal loading history). For
continuous mixers the thermal history remains very short due to the short
residence time, and no evidence of polymer degradation is usually observed.
When high shear batch mixers are used in open contact with air, degradation
occurs rapidly, but this can be minimized by nitrogen blanketing.
Precautions must also be taken to avoid degradation by oxidative attack
during the application of the hot-melt PSAs. In roller melt mill coaters when
the adhesive is transferred from a heated roller the adhesive is exposed to air,
while in die coaters the adhesive is extruded onto the substrate in a closed
system and air exposure remains much more limited.
During the manufacture of hot-melt PSAs a part of the resin and the
antioxidant are fed in first, then the block copolymer, and last the rest of the
resin. Mixing in Z-blade mixers needs 30–35 min. This method is slower
than using a durable (twin) screw extruder. The most important parts of a
fully continuous production line for compounding adhesives are shown in
Fig. 8.17 [162].
Elastomers (either pellets or base product), additives, and some of
the resins are premixed and gravimetrically fed into the premixer. When
formulating with fillers, these also are fed into the premixer. The basic
Manufacture of PSAs 513

Figure 8.17 Production line for continous compounding of hot-melt PSA.

polymers are masticated or, if they are thermoplastics, plasticized by special


kneading elements and mixed with some of the resin. The second feed part
for the rest of the resin is in another premixer and venting takes place there
also. In the subsequent mixing zone the resin is melted and incorporated
together with the liquid components and has to be well homogenized. The
entire compound is very efficiently degassed. After a residence time of only
3–5 min the adhesive compound is ready to be transferred to the buffer
vessel. From there filtering and coating are carried out via a gear pump. The
main component of the compounding plant is a co-rotating twin-screw
compounder. The screw profiles scrape each other as well as the S-shaped
barrel base. This brings about self-cleaning of the screw profile, avoids dead
corners, and results in a relatively narrow residence time distribution. The
screw profile is characterized by a diameter ratio described by Friedrich
[162]. At typical shear rates of 1000–2000 s1, the screw is in the range of
high shear mixers.
Although in principle all heatable mixers can be used to make hot-melt
formulations, the extent varies to which degradation of the base polymers
and polymer resin components occurs. The thermal history of the product,
which is the result of thermal stress and its duration, is to be kept as short as
possible. This can be done by using suitable continuous compounding lines,
if possible in on-line operation with the adhesive coating equipment. This is
a good solution to the problem for medium and large sized batches. For
small and medium sized batches, however, contact between the melt and
514 Chapter 8

oxygen should be avoided and the possibilities of using batch mixers in a


semicontinuous process with the coating line should be exploited. Generally,
high degrees of degradation with open mixing systems and low degrees of
degradation with closed mixing systems are observed. The possibility to
carry out the entire mixing operation in a sigma blade semicontinuous mixer
(medium shear mixer with a shear rate of about 100–200 s1) was
investigated [162]. The mixing machine with sigma-shaped blades remains
the classical mixing system for the production of hot-melt PSAs [185]; it is
being used for mixing volumes of 0.5 l to several cubic meters. Batch times
last from 45 min up to approximately 2 hr, depending on the size of the
mixing chamber.
Hot-melt PSAs do not meet all performance requirements. The melt
viscosity of hot-melt PSAs may change during storage of the molten hot-
melt PSAs [186]. The major benefits of hot-melt PSAs are aggressive tack
and peel adhesion, bonding to rough surfaces, good water resistance, and
economical.
Their limitations are heat resistance, aging, cohesion, and plasticizer
resistance. Since hot-melt PSAs are 100%-solid adhesives it is easy to apply
heavy coat weights at a reasonable speed. Viscosity can be modified within
limits by changing the coating temperature. No drying is required, only
cooling of the adhesive. Therefore coating equipment for hot-melt PSAs
requires a relatively low investment. In die-cutting and slitting operations
the adhesive must be at room temperature since otherwise the risk of
stringing may appear. Hot-melt PSAs need a much lower energy input than
solvent-based or water-based adhesives [187], namely only 0.05 kWh/kg to
0.15 kWh/kg. The energy consumption for hot-melt coating is about 5–10%
of the energy consumption for solvent-based or water-based coating
(i.e., 0.0109 kWh/m2). Furthermore hot-melt PSA coating machines run at
200–300 m/min, compared to 100 m/min for solvent-based or water-based
coating [188]. The thermoplastic character of adhesives based on thermo-
plastic rubbers is a definite advantage in that it makes it possible to apply
the adhesive as a solvent-free hot-melt. However, there are two major
potential disadvantages: the limited maximum service temperature at which
the adhesive will perform satisfactorily and the limited resistance to attack
by common solvents and to PVC plasticizer migration. Conventional
heat-initiated crosslinking systems as used in solvent-applied adhesives are
not suitable for hot-melt processing as premature gel formation can be
expected [176].
Therefore, for hot-melt PSAs radiation curing by exposure to electron
beams or ultraviolet (UV) light immediately after application would be a
more practical approach. Initial irradiation trials on linear SIS block
copolymers showed that high radiation dosages were required and tack was
Manufacture of PSAs 515

lost, especially when polyacrylate curing agents were included. Extensive


research has led to the development of new radiation crosslinkable
polymers. These development polymers have a multibranched SIS structure
(about 10% styrene content). The polymer does not contain functionalized
groups and crosslinking takes place in the isoprene phase; at the same time
the low polystyrene level facilitates good tack even after cure. Other work
also is underway which includes the effect of modifying the endblock and the
introduction of functionalized groups grafted onto the rubber midblock.
Radiation-cured PSAs were discussed by Hinterwaldner [189]. UV radiation
sources produce IR radiation also, therefore a thermal treatment in parallel
to UV light absorption occurs. On the other hand, from both electron beam-
curing methods—linear cathode and scanner method—only the scanner
method ensures a low level of the dosage. Generally, electron beam-curing
generates higher crosslinking than UV light-induced curing.
In radiation-curing applications liquid PSAs or hot-melt PSAs are
used. The liquid PSA formulations are blends of acrylic monomers and
oligomers or macromers. The monomers are crosslinkable and work like
reactive diluents. Although for such hot-melts radiation curable classical
elastomers can be used too, hot-melt PSAs based on warm-melts, i.e., on
relatively low molecular weight polyacrylates are more important because
they are monomer free and curable with UV light. They can be formulated
or not. They may be coated at lower temperatures than normal hot-melt
PSAs. Such an acrylic hot-melt is prepared by solution polymerization of
common monomers along with small amounts of UV reactive comonomers
[190]. Special, electron beam-cured hot-melt PSAs are based on elastomers
such as thermoplastic elastomers (TPE). UV crosslinkable acrylic hot-melt
PSAs were described in a detailed manner in [191].

Water-Based Adhesives
The formulation of water-based PSAs requires special attention because of
the water sensitivity and the water incompatibility of some formulating
agents, and because of the fact that some additives to a water-based
formulation are left in the coated adhesive and thereby modify the chemical
composition of the bulk adhesive. Such additives (used mainly in order to
ensure the fluid, dispersed character of the adhesive) primarily include the
surface-active agents, surfactants, and defoamers. Generally, there are two
categories of additives used in water-based systems: additives used for the
improvement of the adhesive system and those used for the improvement of
the dispersion system. Additives aimed at improving the adhesive system
influence the adhesion/converting and end-use properties; additives used for
the improvement of the emulsion system influence mostly the converting
516 Chapter 8

properties. The most important additives used for improvement of the


adhesive system comprise tackifiers, antioxidants, crosslinking agents, and
peel modifiers. Additives used for improvement of the emulsion system
include surface-active agents, such as wetting agents, stabilizing agents,
solubilizing agents, defoamers, and viscosity modifiers. Special additives
which improve the water resistance also are used. One should note that
additives used for the improvement of the adhesive properties of the water-
based systems are similar to and chemically identical to additives used for
solvent-based systems. However, their physical state may differ.

Improvement of Adhesive Properties of Water-Based Adhesives. The


most important features of tackifiers were discussed in Sections 2.3–2.6.
The same tackifier used for solvent-based systems also can be proposed
for water-based PSAs. Both micromolecular tackifiers (plasticizers) and
macromolecular tackifiers (tackifier resins and other polymers) are
employed. Suitable water-based tackifiers display superior adhesive
performance properties after aging, peel adhesion and tack without major
loss of shear, exhibit resistance to humidity, and specific adhesion to
polyolefin films.
Resins preferably are added as aqueous dispersions in order to
produce solvent-free adhesives. However, the use of resins as solutions in
toluene and white spirit is equally possible and, in some cases, liquid resins
or molten resins may be added also. There are many suppliers of tackifier
dispersions with a different resin base. A detailed description of the
problems concerning the use of water-based resin dispersions as tackifier
was given by Dunckley [55] and Piaseczinsky [192]. Jordan [193] pointed out
that it also is possible to tackify (using mostly hydrocarbon-based resin
dispersions) natural latex, but the most important application of water-
based resin dispersions remains the tackification of acrylic PSA dispersions.
The most important theoretical and practical problem of tackification of
acrylic PSAs, CSBR and EVAc dispersions was discussed in Section 2.3 in a
detailed manner. Trends in water-based acrylic PSAs and the effects of
tackification on properties and performance were described by Wood [46]
and Di Stefano et al. [194]. A quite different tackifying response is suggested
in recent developments of CSBR and EVAc dispersions; CSBR emulsions
typically require 50–60% tackifier to develop optimum properties. New
ethylene multipolymers are less cohesive and accept less tackifier than
acrylic PSAs.
Tackifying with aqueous dispersions differs from the solvent-based
tackification process. Because of the composition of the dispersed systems,
nonadhesive (i.e., technological) additives may play an important role.
As seen when tackifying solvent-based systems, the softening point and
Manufacture of PSAs 517

molecular weight of the resin significantly influences the performance


characteristics of the adhesive. For water-based systems there are some
parameters which override the effect of softening point and molecular
weight of the base polymer. The tackifier supplier’s particular formulation
(surfactant type and level and other components) appears to have the largest
impact on shear resistance. The effects of resin incompatibility on the
viscoelastic, morphological, and performance properties of tackified acrylic
adhesives are qualitatively very similar to those observed for rubber-based
solvent-based adhesives [50]. Predicting the adhesive performance char-
acteristics on the basis of the technological characteristics of the raw
materials remains difficult for water-based systems. Factors such as pH, and
surfactant nature and level play an important role; some resins are acid,
other are esters (neutral). Some base dispersions are acid (CSBR), others
have to be alkaline (neoprene). On the other hand, some surfactants (like
sulfosuccinates) are stable in acid or neutral media only; some thickeners
work in alkaline media only. The surfactant in a tackifier dispersion may be
incompatible with the polymer, causing opacity, even though the tackifier
itself is miscible with the polymer. In this situation it is possible to provide
some guidelines for formulations with a particular tackifier and different
base elastomer dispersions [195]. Blending acrylic dispersions requires a test
of compatibility, the adjustment of the pH, and after addition of solvent,
plasticizers or fillers, etc., a storage time [196]. The compounding of latex
has been discussed in a detailed manner [197,198].

Tackifiers. A wide range of technically equivalent (i.e., yielding the


same adhesive properties) tackifier dispersions are available. Differences
exist concerning their dispersion properties (machinability), logistical and
economical benefits. Most suppliers provide technical brochures containing
data about the chemical composition (resin type), rheological behavior of
the bulk resin (softening point and Tg of the base resin), physical
characteristics of the base resin (e.g., color), and applications. The adhesive
properties with different base elastomers (acrylic PSAs, CSBR, natural
rubber, etc.) also are evaluated and physical properties and the stability of
resin emulsions are described. In Chap. 5 the various chemical bases of
tackifiers were discussed in a detailed manner. The selection of a supplier
also is determined by logistical and economical considerations. The main
criteria are the existence of a convenient production site, the range of
tackifiers on the market (e.g., pine/tall oil/hydrocarbon-based resins), the
availability of raw materials, knowledge of tackification and emulsification,
development of new resin types and improvement of resins (over the last
decade), number of successful grades, knowledge of PSAs, estimated
innovative capacity, sufficient production capability, supplier flexibility,
518 Chapter 8

technical assistance in converting, and last but not least price. Tackifier resin
dispersions, their formulation, manufacture, and stability are described in
detail in [167].
Plasticizers. The plasticizers used for solvent-based PSAs also are
used, but at a lower level, in water-based PSA formulations. The use of
plasticizers is more difficult in the case of water-based systems because of the
solvent role of the liquid plasticizer, influencing the rheology and wetting
properties of the aqueous system. Water-based partially soluble plasticizers
may increase the viscosity. Migration of the plasticizer (strike-through) is
more dangerous with water-based systems; therefore the plasticizer level
used for water-based systems remains lower than that generally accepted for
solvent-based ones. The technology of addition is very important. In making
up the coating mix for a peelable PSA using a hard latex with a plasticizer,
there are advantages to making up the mix by first mixing the emulsion
or latex with the plasticizer, which is usually supplied as a liquid. This
incorporates the plasticizer into the hard (reinforcing) dispersion before
adding the main adhesive polymer. This procedure appears to reduce any
tendency of the plasticizer to cause undesirable migration in the finished
product. Polybutenes are first emulsified to form an oil-in-water emulsion,
using either an anionic or nonionic surfactant. An alternative method of
addition is to use a solution of the polybutene (e.g., 50% solution in
toluene). However, while this may be convenient, it can only be used in
plants where solvents are already being used and recovered. The polybutene
emulsion or solution in solvent is blended with the acrylic latex to give the
desired level using a high speed, low shear mixer [54].
Suitable antioxidants do not influence the printing properties of the
laminate. Generally, antioxidants used for PSAs should be tested (before
evaluation of their antioxidative effect) for their interaction with printing
inks. In general, the same antioxidants used for solvent-based PSAs also
are adequate for water-based PSAs. In this case they are supplied as
aqueous dispersions (Bayer) or may be dissolved in alcohol (AS-14 Bayer).
Table 8.25 lists some common protective agents used for PSAs.
Peel Modifiers. Tackifiers (plasticizers and resins) act as peel
modifiers. There are other additives with a tackifying effect that exert a
more pronounced effect on the peel, like polybutenes (e.g., Hyvis). These
additives added to the acrylic latex yield a more pronounced peel increase
than their aqueous dispersions. Thus, not only their amount, but their form
(state) appears important. The compatibility of macromolecular additives
should also be taken into account. Polybutenes show incompatibility with
acrylonitrile (AN) copolymers, causing a marginal decrease in tack. Another
class of peel modifiers are the silicone derivatives (e.g., Goldschmidt). These
Manufacture of PSAs 519

Table 8.25 Common Protective Agents Used for PSAs

Application Field Protection Against

Water-based Solvent-based Heat, oxygen,


Product Code Supplier PSA PSA UV light aging

Vulkanox Bayer — x — x
BKF Bayer — x — x
NKF Bayer — x — x
SKF Bayer — x — x
AS13 Bayer x x x —
Naugard 445 Lehmann — x — x

are liquid compounds added in 0.5–2.0% wet weight to the aqueous


dispersion in order to reduce the peel. Unfortunately they only act at room
temperature. High temperature aging of removable formulations on the
basis of these additives yields no reduction of the peel adhesion.

Crosslinking Agents. Crosslinking agents interact with functional


groups from the bulk polymer. Crosslinkers in the form of water-soluble
agents have only a restricted interaction with the bulk polymer (from the
dispersion particles) but an unlimited interaction with the aqueous medium
(dispersion additives) and the moisture-sensitive liner. The use of postadded
crosslinking agents for water-based systems should be limited. On the other
hand, some organic crosslinking agents are micromolecular, with a delayed
crosslinking action up to the drying step, thus having enough time to give a
penetration in the uncrosslinked state (e.g., Bayer, aqueous dispersions of
PUR oligomers). External crosslinking agents include polyaziridines,
epoxies, aminoplasts (such as melamine formaldehyde resins), and metal
salts or oxides such as zinc oxide or zirconium ammonium carbonate, etc.;
N,N0 -bis-1,2-propenyl-isophthalimide can be used also. Amine-formalde-
hyde condensates (0.8–10%) may be used together with substituted
trihalomethyltriazine as additional crosslinking agent. The most important
class of crosslinking agents used are the polyfunctional aziridines.
Polyaziridines are suggested for use in crosslinkable water-based PSAs.
The most well-known commercial product used is Neocryl-CXl00. The
reproducibility of the crosslinking process is low. In order to increase
cohesion and water resistivity, Hercules Polycup resins were evaluated.
These are water-soluble, polyamide epichlorohydrine (EPC)-type materials
effective as crosslinking agents for carboxylated acrylic PSAs or CSBR
latexes, carboxymethyl cellulose (CMC), PVA, etc. It was proposed to use
0.5% solids of the latex-based crosslinker in order to increase the modulus.
520 Chapter 8

A hexamethylol melamine resin, together with 0.5 mole p-toluene sulfonic


acid catalyst also may be used. Such fluid, heat-reactive resin was proposed
as a tackifier for neoprene latex, yielding very high bond strength upon the
application of heat. Polyvinyl alcohol (PVA)-stabilized dispersions may be
crosslinked with Quilon (Hercules).

Improvement of Emulsion Properties


Hot-melt PSAs and solvent-based systems are stable systems with unlimited
shelf life. Dispersed systems and aqueous or solvent-based dispersions may
coagulate as a result of shear forces, temperature, or chemical interactions.
Therefore stabilizing agents (mostly surface-active agents) are added in
order to improve their mechanical and/or shelf stability. These additives
ensure the dispersion of the PSA before and during coating, its coatability
in the optimum case having only a minor effect on the end-use properties
of the products. As dispersion additives to improve the stability of PSA
dispersions surface-active agents (emulsifiers) or protective colloids may
be added. The polymerization is carried out with emulsifiers, but post-
stabilizing surface-active agents can be added as well. Sedimentation
depends on the particle diameter, the density difference between solvent
and solute, and the electrical potential of the particles. According to [199]
methods to improve the stability of tackified acrylic emulsion may include
proper selection of the tackifier dispersion or using protective colloids.
Stabilizing Agents. Stabilizing agents are emulsifiers (surface-active
agents) used during the polymerization or thereafter in order to ensure shelf
stability under static and dynamical conditions (storage, formulation, coat-
ing). Generally, stabilizing agents are recipe specific and it is not the task of
formulators to select them and to specify their level. Carboxylated latexes
have very good inherent colloidal stability and need no additional stabilizers.
Many emulsifiers are commercially available for emulsion polymerization
and may be classified as anionic, cationic, or nonionic. Generally, anionic
emulsifiers are preferred, but nonionic emulsifiers also are efficient [135].
The common anionic surfactants are alkyl or alkyl ether sulfates such
as sodium lauryl sulfate, sodium lauryl ether sulfate, alkyl oxyl sulfonates
(e.g., sodium dodecyl benzene sulfonates), and alkyl sulfosuccinates (e.g.,
sodium hexyl sulfosuccinates) (2–3%). Surface-active agents, like surfac-
tants, are used in dispersions for stabilizing and for wettability. A detailed
discussion of their effect on the properties of the coated adhesive will be
presented in Section 3.1 of Chap. 10.
Protective Colloids. Protective colloids are necessary for dispersions
with particle sizes larger than 1 pm (because of the large particle surface)
Manufacture of PSAs 521

[200]. The use of protective colloids allows better mechanical, electrolytic,


and storage stability, as well as neutral pH, higher particle size, narrower
molecular weight distribution, and higher surface tension [201]. Using a
polymeric colloid which is a low molecular weight alkali-soluble polymer,
surfactant-free adhesives may be prepared [107]. On the other hand, a high
viscosity, a lower tack, and higher water sensitivity are characteristics of
protective colloid-stabilized latexes. In general, only a few PSA dispersions,
mostly based on EVAc or neoprene, contain PVA as a stabilizer.

Wetting Agents. Inherent problems common to water-based systems


include substrate wetting, foaming, foam and leveling deficiencies, as well as
a possible reduction in adhesive properties. The first requirement for the
bonding process is that the adhesive must establish intimate contact with the
adherent; it must be capable of wetting it [202]. For virtually all uses of
PSAs this is the case; the surface energy of the adhesive is much lower than
that of the adherend, which is a condition for good wetting. The wetting
process is far from being the instantaneous phenomenon one would like to
believe it to be. Using radioactive tracers, Toyama, Ito, and Nakatsuma
[203] showed that the adhesive was present in discrete parts. The quantity
percentage depends on the dwell time of the adhesive.
In order to improve the coating properties (i.e., wet-out), surfactants
are added to the formulation. Surfactants are defined as substances which
significantly reduce the surface tension of liquids at very low concentrations.
The degree to which these molecules accomplish this depends on the balance
of their hydrophilic and hydrophobic regions. When properly balanced,
surfactants will concentrate at the liquid/air interface, causing a compressive
force to act on the surface [204]. For a given solid, liquid, or fluid medium, a
unique value for the contact angle of a triplet line would be expected, but in
practice a given system often displays a range of contact angles between lower
levels (receding angle) and an upper limit (advancing angle); this difference is
often referred to as hysteresis. While the static reduction of the surface tension
can give the formulator a quick idea of the efficiency of the surfactant, many
industrial applications never reach equilibrium. Therefore it is important in
processes where surfaces are generated at a rapid rate, that the surfactants
migrate rapidly to the interface so as to prevent film retraction and other
surface defects; hence the dynamic surface tension is a very important
parameter. Good dynamic surface tension means that the surfactant migrates
rapidly to the interface so as to prevent film retraction. The formulator has to
employ the wetting agent(s) leading to good wetting on the most difficult
surfaces at the lowest viscosity level and necessary defoamer level, and at the
lowest wetting agent concentration. Surface tension reduction and dynamic
surface tension characterize the activity of the wetting agent, leaving only a
522 Chapter 8

limited range of available wetting agents (independent from the economic


aspects). On the other hand, water sensitivity and migration, and the latter’s
influence on the peel and removability, limit the level of the wetting agent to
be used. In order to select the surfactants, their chemical nature, their surface
tension reducing effect, and their influence on the coated adhesive layer
should be determined. The choice of the surfactants is not so simple because
the required surface tension depends on the characteristics of the web to be
coated (face stock or release liner) as well as the coating parameters (coating
device). Aqueous dispersions usually have a built-in level of surfactants
that are required for dispersion stability during and after the polymerization
or for handling purposes.
These dispersion inherent surfactants may be polymerization related or
postadditives. Postadditives may be micromolecular (surfactants) or macro-
molecular (protective colloids). Their level depends on the end-use of the
PSAs [205]. Quite different surfactants are used for primary dispersions
(elastomer dispersions: acrylic PSAs, CSBR, EVAc) or secondary ones (resins
or other tackifiers, antioxidants, etc.). On the other hand, the different
chemical base of the elastomer dispersions requires different emulsifiers.
Generally, surfactants (not protective colloids) are used for PSAs (as
compared to viscosity, water sensitivity), but polyvinyl alcohol-stabilized
carboxylated neoprene latexes make an excellent base for PSAs [12]. The
choice of emulsifiers used for secondary dispersions depends on the nature of
the additive, but its molecular weight is important also. During the emulsi-
fication of polybutenes, surfactants with different hydrophilic-lipophilic
balances (HLB) are used as a function of the molecular weight of polybu-
tenes; the optimum HLB for polybutenes increases with increasing molecular
weight. Regarding the choice of surfactants for primary dispersions, it
depends on the nature (original surfactant level and surface tension level) of
the elastomer dispersion and of its formulation (tackified or not).
The surface tension of most neoprene latexes and their compounds
is relatively low (usually around 30–40 mN/m as compared to water at
72 mN/m); for example, Neoprene Latex 102 has a surface tension of
58 mN/m at 20 C [201]. A polyvinylacetate latex has a surface tension
of 38 mN/m, whereas styrene-butadiene latexes possess a surface tension of
45–60 mN/m [206]. Dispersions of the same composition may have quite
different surface tensions (Table 8.26); moreover the same dilution of a
dispersion may lead to very different surface tensions (Fig. 8.18).
The use of wetting agents has been reviewed [22]. The choice of the
wetting agent depends on the efficiency and the influence on the adhesive
end-use performance (i.e., the activity to improve wetting and coatability).
Their influence on the adhesive performance characteristics concerns the
change of the adhesive properties, optical properties, chemical resistance,
Manufacture of PSAs 523

Table 8.26 Surface Tension Values


of Water-Based PSAs

Batch Surface Tension


Number Value (mN/m)

01 42.4
02 41.9
03 41.6
04 40.0
05 40.8
06 40.2
07 40.8
08 43.2
09 39.2
10 44.0
11 40.8
12 39.0

Figure 8.18 Dilution of water-based PSA. Interdependence of viscosity and


surface tension.
524 Chapter 8

and aging of the coating. Their versatility to be used (ease of compounding,


influence on the viscosity, drying speed) has to be considered. For example,
some sulfosuccinates have only a limited solubility in cold water; in order
to improve their solubility alcohol should be added. Some sulfosuccinates
are sensitive to pH changes and thicken with the increase of the pH.
Sulfosuccinates are sometimes hygroscopic and therefore cause bleeding and
migration; they can reduce the drying speed, or cause adhesive residues
upon removal of the label from a substrate (i.e., they are not recommended
for removable PSAs). The most common wetting agents are sulfosuccinate
derivatives [10,207]; dioctylsulfosuccinates display better properties than
other esters (Table 8.27).
The selection criteria for wetting agents include the ability to provide
good coverage on difficult-to-wet surfaces, a reduced foam level, no adverse
effect on adhesive properties, and a reduced water sensitivity. The best
ability to provide good coverage on difficult-to-wet surfaces is displayed by
sulfosuccinates (Table 8.28). Fluorinated surfactants are superior but not

Table 8.27 Common Sulfosuccinate Derivatives Used as Wetting Agents for


Water-Based PSAs

Solids
Content
Code Supplier (%) Chemical Basis

1 Aerosol OT Cyanamid — Natrium, dioctylsulfosuccinate


2 Aerosol 18 Cyanamid 35 Di-natrium, N-octadecyl sulfo-
succinate
3 Aerosol A 102 Cyanamid 30 Di-natrium, fatty alcohol,
sulfosuccinate
4 Aerosol MA Cyanamid — Natrium, dihexyl sulfosuccinate
5 Aerosol AY Cyanamid — Natrium, diamil sulfosuccinate
6 Disponil SUS-65 Henkel 40 Natrium, fatty alcohol, ethylene
7 Fenopon GAF — —
8 Humifen W GAF — Natrium, alkyl sulfosuccinate
9 Lankropol GR-2-3 Münzing — —
10 Lankropol KO2 Münzing — Natrium, di-isooctyl sulfosuccinate
11 Lankropol KMA Münzing — Natrium, dihexyl sulfosuccinate
12 Lumithen IRA BASF — Natrium, dioctyl sulfosuccinate
13 Rewopol SBFA 50 Rewo 30 Di-natrium, fatty alcohol,
polyglycolether sulfosuccinate
14 Rewopol SB-MB-80 Rewo 80 Di-isohexylsulfosuccinate
15 Rewopol SBFA 30 Rewo 40 Di-natrium, fatty alcohol,
polyglycolether sulfosuccinate
16 Triton GR-5 Rohm 60 Natrium, dioctyl sulfosuccinate
Manufacture of PSAs 525

Table 8.28 Wet-Out Improvement, Using Acetylenic Diol Surfactants in a Blend


with Ethoxylates and Sulfosuccinates

Surface Active Agent Characteristics

Formulated surface active agent


Global surface
Code Genapol X80 Surfynol 61 Lumithen IRA active agent Wet-out
Supplier Hoechst Air Products BASF level (wt%) index

50 50 — 2 4
50 50 — 4 4–5
33 66 — 3 4
66 33 — 3 3
25 75 — 4 2
— 50 50 2 2
— 66 33 1.5 2
— 80 20 1.5 2

Table 8.29 The Influence of Fluorinated Surfactants on the


Surface Tension of a PSA Formulation

Fluorinated Surface Concentration Surface Tension


Active Agent (%) (mN/m)

Zonyl, FSK 0 35.7


0.2 36.8
0.5 30.7
Zonyl PT 448 0 40.8
0.2 41.7
0.5 42.8
Acrylic PSA: Acronal V 205.

economical [208] (see Table 8.29 also). In some cases better results are
obtained with acetylenic diol surfactants alone or in mixtures [204].
The presence of surfactants can lead to foaming. The evaluation of
the foam level is discussed in Chap. 10. The lack of adverse effect on
the adhesive or end-use properties remains very critical. In order to limit
or avoid this, it is necessary to determine the influence (mechanisms) of
surfactants on the PSA layer.
Surfactants exert a chemical and a physical influence on PSAs.
Surfactants interact mainly with the bulk polymer, with components used as
reactants for the bulk polymer, or with other dispersion additives. Similar
526 Chapter 8

interactions are chemically possible between the surfactants and the other
components of the laminate (face stock or liner). Physical processes
occurring with the aid of the surfactants may influence the properties of
the PSA laminate. A surfactant in an adhesive tends to migrate to the
surface, where it often reduces the bond strength [10,107,209]. Surfactants
also may react with chemicals added in the formulation in order to interact
with the reactive groups of the bulk polymer; an interference with certain
crosslinking agents used as modifiers also may occur [10].
The amount of nonadhesive additives in water-based dispersions
influences the adhesive properties of the PSAs; thus the effect of surfactants
should be taken into account. The adhesive performance is influenced to a
larger degree than expected, by the surfactants and other additives in the
tackifier dispersion [63]. Midgley [14] found that surfactants reduce the tack,
but with the exception of a slight change in rolling ball tack, there is little
difference between the same water- or solvent-based adhesive. Emulsion
polymers contain surface-active agents which cause a decrease in tack and
adhesion. For CSBR different, optimum tack values are achieved, when one
reference latex is used with different emulsification systems. Low levels of
potassium rosin acid soap may even improve the tack and peel adhesion of
CSBR latexes. Similar effects are obtained by using ethoxylated tetramethyl-
decyne-diol (Surfynol 465), and the nonionic soap, octyl-phenoxy-poly-
ethoxy-ethanol (Igepal CA 630, GAF) [210]. Lauroyl sulfates migrate
rapidly to the surface and reduce tack. Ethoxylated octylphenol affects
(reduces) the peel after aging [211]. Water-based PSAs contain surfactants
which complicate the formulating process because of the interference with
certain crosslinking agents [10]. Furthermore a surfactant in an adhesive
tends to migrate to the surface where it often weakens bond strength. The
presence of the surfactant decreases the wet bond strength of the laminate
[107]. Furthermore, surfactants reduce the peel of PSAs [212]. According
to Delgado [213,214] factors that have been implicated in causing lower
adhesive performance in emulsion films were surfactant migration and
incomplete film formation. For special applications an emulsifier is added
into the adhesive formulation in order to avoid the adhesive acting like a
rubber at low temperatures (i.e., in this case the emulsifier is acting like
a plasticizer). The influence of the surfactant on the peel also has been
summarized by Makati [215].

Interaction of Surfactants with Other Layers in the Laminate. In the


particular case where the adhesive is to be used to form self-adhesive
laminates and where at least one of the surfaces is a printed surface, the
presence of any residual surfactant can lead to discoloration and bleeding
of the ink [14]. This is a recognized problem in applications such as
Manufacture of PSAs 527

over-laminated books or printed labels when the purpose of the film is to


preserve the integrity of the printed surface.
Generally, the hardness of a clear film obtained by evaporation of a
latex is largely dependent on the polymer composition and the temperature.
The surfactants and colloids have the same effect, with regard to the
humidity. The presence of the surfactant reduces the water resistance of
the laminate [107]. Water absorption in acrylic latex films is influenced by the
storage conditions/time or thermal coalescence, which reduce water absorp-
tion [216]. Addition of emulsifier for latex stabilization increases the starting
rate of water dissolution in acrylic films; an increased amount of surfactant
reduces the drying rate under given conditions. Ethoxylated nonylphenol
derivatives bring a high viscosity increase, but poor cold water solubility [98].
Anionics are strongly hydrophilic and are particularly undesirable in wet
environments [211]. The water resistance of the surfactants depends on their
chemical composition. The resistance to rewetting is good for fatty acid
derivatives, poor for anionic, and fair for nonionic materials.
Generally, polymers have different surface tensions. Normally ready-
to-use systems are blends of several batches. The same amount of wetting
agent is not sufficient in order to ensure the same surface tension for PSAs
from different sources (i.e., the required surfactant level depends on the base
polymer) (Fig. 8.19).
Generally, the nature and the level of the surfactant used depend on
the base polymer of the PSAs. For ethylene vinyl acetate-dioctyl maleinate
copolymers the amount of the emulsifying agent varies by a factor of 1 to 10,
preferably about 2–8% of the amount of monomers used [110]. Water-based
acrylic PSAs possess a lower surface tension than common EVAc or CSBR
dispersions. Current surface tension values are 34–40 mN/m while EVAc
dispersions show more than 45 mN/m; rubber latexes have a surface tension
of 45–60 mN/m [217]. Water-based acrylic PSAs generate more coagulum
(grit) than CSBR dispersions and exhibit a better mechanical stability
than CSBR.
Generally, PSA properties improve as the anionic surfactant con-
centration increases. All major properties can be altered by changes in
surfactant nature and level. The hydrophilic-hydrophobic balance of a
surfactant is another of the physical parameters available to enable an easier
choice of the emulsifiers. Relating the water-soluble and water-insoluble
part of the surfactant is a useful guide to the solubility of the product itself
and to the solubilizing ability which it possesses.
Surface-active agents were used in synthetic latexes in amounts of
1–6% [218]. Natural latexes contain 1% surface-active agents. Some natural
latexes include acids as an emulsifying agent [121]. In practice 0.75–1.5%
(wet/wet) sulfosuccinates and 1 to 3% ethoxylates are used; the surfactant
528 Chapter 8

Figure 8.19 Dependence of the required surface-active agent level on the nature of
a PSA dispersion. A, B, C, and D are different water-based acrylic PSAs.

level also depends on the pH. The level of the water-sensitive wetting agents
may be lowered using the special fluor-based emulsifiers or the volatile
ones (Surfynol S 61), but actually they remain too expensive. Table 8.29
illustrates the influence of the concentration of fluorinated surfactants on
the surface tension. The viscosity of several dispersions depends on the pH.
As wetting-out is a function of the viscosity, the level of the wetting agent
necessary in order to achieve a good wet-out depends on the pH (Fig. 8.20).
As seen in Fig. 8.20, the amount of the sulfosuccinate active agent used for
wet-out depends on the pH; a minimum level is required in the high acid
domain.

Adjustment of Viscosity. The wet-out is a function of viscosity [219],


surface tension of the water-based adhesive, and foaming. The viscosity of
the formulation depends on its basic components. Elastomeric dispersions
exhibit a different viscosity, depending on their chemical composition (the
presence of water-soluble monomers) dispersion characteristics (which
depend on the surfactants used during their synthesis), and solids content.
Generally, water-based acrylic PSAs possess a higher solids content and a
higher solids content/viscosity ratio than common CSBR/EVAc dispersions
as well as a better dilution response (Table 8.30). This is very important when
the viscosity has to be lowered for coating (machinability) purposes. In some
Manufacture of PSAs 529

Figure 8.20 Dependence of the surface-active agent level required for good wet-
out on the pH of water-based acrylic PSAs.

cases, the running viscosity has to be increased again (before or after


dilution); there are certain special applications where reducing the viscosity
appears required.
With regard to the influence of the viscosity and the surface tension
on the wetting behavior, Hansmann [220] stated that the decrease of
the viscosity influences the wetting more than the surface tension. The
rheological shortcomings of latex PSAs with conventional surfactants arise
at least in part from the need to incorporate thickeners, such as hydroxyalkyl
cellulose or polyacrylic acid. Without thickening the conventional low
viscosity latex will not completely wet a silicone-coated release liner. This is a
necessity for transfer coating which is preferred for heat sensitive face stocks
[107]. Addition of thickeners (alginates) may avoid stripes [221]. Migration
through face stock paper may be avoided using thickeners as well [222]. To
improve surface coverage the viscosity can be increased with the addition of
cellulosic or polyvinyl alcohol (PVA) thickeners. However, problems with
machinability, air entrainment, and moisture sensitivity may arise. Thixo-
tropy can be introduced by adding different components such as silica,
ricinus oil, titanium, or zirconium chelates [86]. As thickeners methyl
cellulose, PVA, starch, polyacrylate salts, alginates, dextrine, and ethoxyl
cellulose also can be used [223,224]. Copolymers of methylvinyl ether and
530

Table 8.30 Solids Content and Viscosity of Water-Based Acrylic PSAs Versus EVAc and CSBR PSAs

Water-Based PSA Emulsion Characteristics

Viscosity Shear rate Total


Brand name Code Supplier Base Method (mPa.s) (s1) solids (%)

Dilexo N600 Condea Acrylic DIN 53019, 20 C 400–800 57 55


Dilexo AK610 Condea Acrylic DIN 53019, 20 C 800–1500 57 55
Latex 3703 Polysar CSBR Brookfield LVF, S2, 30 rpm 250 — 50
Latex 3958 Polysar CSBR Brookfield LVF S2, 30 rpm 500 — 53
Mowilith 130 Hoechst EVAc Brookfield RVT S5, 20 rpm 5000–11000 — 50
Mowilith 131 Hoechst EVAc Brookfield RVT S5, 20 rpm 5000–13000 — 50
Mowilith 132 Hoechst EVAc Brookfield RVT S5, 20 rpm 4000–1000 — 60
Acronal 4D BASF Acrylic DIN 53019, 23 C 15–40 250 50
Acronal 80 D BASF Acrylic DIN 53019, 23 C 80–150 250 50
Acronal V 205 BASF Acrylic DIN 53019, 23 C 800–1600 250 69
Nacor 314 National Acrylic Brookfield > 200 — 52
Nacor 525 National Acrylic Brookfield > 200 — 53
LVT, known Brookfield viscosity method; RVT, known Brookfield viscosity method.
Chapter 8
Manufacture of PSAs 531

maleic anhydride are used as thickeners for natural rubber and synthetic
latexes [225].
According to their activity as thickeners, these agents may be classified
as pH regulators, surfactants, protective colloids/polyelectrolytes, water-
soluble polymers, fillers, crosslinking agents, or solvents. It is evident that
most PSA emulsions containing polar groups incorporated in the polymer
backbone should modify their viscosity through a pH change. For this
purpose, alkaline additives (mostly fugitive ones) are used. In some cases
discoloration (by ammonia or alkaline hydroxides) was observed and
organic amines are preferred. The change of pH is required for some
polyelectrolyte thickeners (polyacrylic acids) (i.e., these act mostly in the
alkaline domain).
Aqueous solutions of carboxylated polymers have low viscosities. The
polymer is believed to be tightly coiled and only slightly ionized; as the pH is
raised, more carboxyl groups become ionized, the polymer chain is further
uncoiled, and, as a result, the viscosity increases [101]. Changes in the flow
curves of acrylic polymer dispersions incorporating acrylic acid during a
treatment with ammonium hydroxide were described with the Cross theory
of aggregation of dispersed particles [226]. If the surface tension value of the
material to be coated is too high and the formula is not thickened properly,
defects in the film will appear. Cratering, pinholing, and crawling are all
evidence of poor substrate wetting, of a too high surface tension value,
and/or of an improperly thickened system. Viscosity adjusters are agents
used to lower or increase the viscosity of the water-based systems.
High speed rotogravure machines run with 100–500 mPa.s dilute
dispersions. On the other hand, good wetting requires a higher viscosity or/
and wetting agents. Some wetting agents increase the viscosity also, thereby
thickening the dispersion. The high running speeds assume higher solids
contents associated in several cases with an inadequate rheology of the
semidry coated layer, resulting in an unsmooth coating. In this case dilution
of the base dispersion and rethickening it is necessary. Summarizing, two
kinds of viscosity adjustments are needed: thickening and lowering of the
viscosity; pH regulating agents are not really thickeners, but they are added
more in order to increase the dispersion stability or to change certain
adhesive properties. Table 8.31 lists some common thickeners used for
water-based PSAs.
Thickeners are manufactured by build-in of hydrophilic or salt-
forming reactive groups on polymers. Soft hydrophobic and hydrophilic
comonomers (butyl acrylate/methacrylic acid, BuAc/MAA) produce thick-
ening latices (their viscosity increases more than 1000% in the alkaline
region) which makes the penetration of the alkali (natrium ions) into the
polymer particle possible, and thus changes the polymer volume and
532 Chapter 8

Table 8.31 Common Thickeners Used for WB PSA Formulations

Usual
Concentration Comments,
Code Supplier Basis (wt%) Ref.

Borchigel L75 Borchers Polyurethane < 1.5 —


Viscoatex 46 Coatex Acrylic <2 —
Coatex BR 100 Coatex Polyurethane < 1.5 —
Cyanamid Cyanamid Modified <2 FDA approval
A-370 polyacrylamide
Kelzan M Lancro Polysaccharide, 0.2–0.5 Tested for
natural acrylics
sugar derivative with good results
Viscalex V430 Allied Acrylic 0.1–0.5 Current
Colloids industrial use
Polysar Polysar Acrylic <2 Proposed
Latex 6100 for CSBR
Rohagit SD15 Rohm & Haas Acrylic < 1.5 [227]
Latekoll D BASF Acrylic < 1.5 Current
industrial use
Jaguar CHMP — Natural <1 Proposed for
polysaccharide neoprene latex

Kelset Langer Natrium calcium <2 [228]


alginate
Collacral VL BASF Acrylic <2 Recommended for
film coatings

structure. The influence of the pH regulating agent was discussed in Section


2.2 of Chap. 2.
Some surfactants also act as thickeners. This behavior is very
pronounced for certain sulfosuccinates and may differ for the ‘‘same’’
compound from different suppliers. Care should be taken whether
surfactants should be used as thickeners. PSA dispersions possess a certain
surfactant level as a result of the polymerization process; some surfactant
amount can come from the face stock material. All these may interfere
with the adhesive and end-use properties (see Section 1.1 of Chap. 6). The
hydrophobic segment of a soap can migrate to the relatively nonpolar
rubber adhesive and alter release, tack, and readhesion properties.
Contamination can come not only from the migration of certain species
in the release coating, but also from species in the face stock such as
plasticizers from fibers or the soaps in paper saturants. Calendered PVC
(emulsion PVC, EPVC) made from 0.1 m diameter particles, has a thin
Manufacture of PSAs 533

layer of emulsifier on its surface; this layer behaves like an adhesive [229].
Thickeners can increase water sensitivity and can act as nutrients for
microorganisms. Protective polyacid-based colloids are preferred, together
with polyamides, because they are less sensitive to microbial attack and
cause less decrease of the tack than natural derivatives.
Polymer thickeners for emulsions are copolymers of (meth)acrylic
acid (25–45%) and alkyl (meth)acrylates (25–65%), esters, ethylene glycol
derivatives (1–40%), polyethylene unsaturated monomers (0–11%), and
hydrophilic ethylene monomers (< 10%) [230]. Water-based polymers (e.g.,
polyvinyl ethers) improve the adhesive characteristics also, but display water
and aging sensitivity. Polyvinylpyrrolidone in PSAs imparts a high initial
tack, strength, and hardness; water-remoistenable PSAs based on PVP, cast
on paper, reduce the usual tendency to curl too [231]. N-Vinylpyrrolidone-
ethyl methacrylate (EMA) copolymers may be used as thickeners also
[232]. Fillers may act as thickeners too. They generally decrease the adhesive
properties and therefore have to be used in concentrations below 5%.
Crosslinking agents (e.g., metallic salts) interact with the polar monomers
from the bulk polymer or from the protective colloid and thus may produce
thickening also. They improve the water resistance, but decrease the adhesive
properties. Cohesion and heat resistance can be increased further through
crosslinking. The divalent metal ion/acid functionality approach used with
acrylate PSAs also applies to new systems such as ethylene/vinyl acetate/
maleinate copolymers. Solvents may be added to water-based adhesives and
can improve their anchorage by 7–30% [224]. Depending on their solubility
in water, solvents increase the viscosity of water-based PSAs. The choice of
an adequate thickener depends mainly on the raw dispersion and the end-use
requirements. The thickening power on polymer latices is influenced by the
type of emulsifier, particle size (i.e., surface area), and presence of mineral
fillers. High molecular weight water-soluble polyelectrolytes are readily
absorbed by polymers or mineral particles. The degree of absorption depends
on the type of latex, emulsifiers, thickeners, or mineral filler employed. For a
poorly stabilized latex, the absorption is the greatest and the thickening
power the most pronounced [233]. A more recent development is the range of
the so-called associative thickeners. Such products include hydrophobically
modified ethylene-urethane block copolymers. Associative thickeners consist
of molecules with hydrophilic and hydrophobic groups. Their effect is based
on the interaction of hydrophobic segments of the molecule with other
components of the coating.
In some dispersions the viscosity may be too high, even though no
conventional thickener is present. The easiest way to reduce the viscosity
is to dilute the emulsion with water, if a reduction in solids content can be
tolerated. If a high solids content must be maintained, the addition of alkaline
534 Chapter 8

(e.g., sodium or potassium hydroxide, sodium polyphosphate, ethylene dia-


mine, diethylamine, etc.) may reduce the viscosity. Urea (up to 6% wet weight)
may be used to decrease the viscosity of natural latexes [234]. The viscosity
may be decreased by special additives like polyvinylpyrrolidone [235].
Table 8.32 lists some viscosity reducing agents; Fig. 8.21 illustrates
the influence of carbamide addition on the viscosity of a common
water-based PSA.

Table 8.32 Agents for Lowering the Viscosity of PSAs

Water-Based PSA Viscosity

Viscosity Reducing Agent Code Supplier Before After Refs.

Dicyandiamide 1360 Hoechst 5000 2400 [236]


Carbamide V 205 BASF 5000 3600 —
Polyvinyl pyrrolidone Natural latex — — — —

Figure 8.21 Viscosity decrease through formulating. The use of carbamide as


a viscosity reducing agent. The Brookfield viscosity of the formulation after: 1) a
storage time of 2 days; 2) instantaneously.
Manufacture of PSAs 535

Coatability and Foaming


By virtue of lowering the surface tension, any surfactant reduces the energy
needed to create foam; however, what really dictates the foam volume
achieved is the stability of the foam. Surfactants form a dense, ordered,
viscous film on the surface of aqueous solutions in which bubbles of air
can be encapsulated. The mechanism apparently involves the hydrophilic
polyethylene oxide chains forming closely packed hydrated layers that
are crosslinked by interstitial water molecules. In contrast, defoaming
copolymer surfactants form low viscosity, mobile, open surface films that
form areas of weakness in the surface of a bubble, eventually leading
to bubble rupture [237]. There is no generally valid theory concerning the
mechanism of defoamers [238].
Generally, if more surfactants are used there is a competitive
adsorbtion of the nonionic and anionic surfactants on the latex particles.
The addition of one surfactant may cause the desorbtion of another one.
In practice, this can cause problems such as foaming [239]. Foaming also
depends on the surface quality; on porous surfaces more foam is left [238].
In stirred systems foam builds up if the strain  is greater than the capillary
pressure p [240]. The strain depends on the particle diameter  according to
the following equation:

 ¼ 1:9 ð"  Þ2=3 ð8:4Þ

where denotes the density of the particles. The particle diameter is a


function of the surface tension and therefore foaming is a function of the
surface tension.
Foaming on the coating device depends on the depth of the immersed
coating cylinder. Deeply immersed coating cylinders cannot carry much air
into the liquid, therefore low foaming results. The rotation direction of the
coating cylinder influences foaming also. Reverse cylinder sense (against
the web direction) is suggested [241]. Temperature (related to the choice of
the surface-active agent) also may influence foaming [211]. Typical low
foaming nonionic surfactants foam less at temperatures above their cloud
points because they become insoluble.
Defoamers are often used because the polymers are present as
emulsions (or latexes) in water and the hydrophobic nature of the polymers
implies that substantial amounts of surfactants are necessary in order to
maintain the stability of the dispersion. Conventional defoamers (e.g.,
mineral oils and waxes) or silicones can be typically used in an amount of
about 0.25 wt% of the wet coating mix. Defoamers are added into the
dispersion in order to reduce foam formation. Their use is well known from
536 Chapter 8

the synthesis of water-based adhesives, where defoamers are necessary to aid


free monomer reduction when heavier reflux develops. Different kinds of
defoamers (e.g., silicones, fatty acid derivatives, or alcohols) in an amount
of about 0.15% are suggested [239]. Defoamers absorb into the polymer,
especially at elevated temperatures, and thus have a short life and require
frequent replenishment. When foaming during the handling or coating is
encountered, regular additions of defoamer should be made. Many additives
were found to minimize or eliminate the formation of foam, and a large
number of proprietary defoaming agents are on the market. None is
completely satisfactory for all compounds; conditions of use are such that
only by trial and error can one identify the most efficient agent or
combination of agents for a given compound and process. With any
antifoaming agent the minimum amount required should be used since such
agents tend to cause localized coagulation, poor wetting, and/or ‘‘fish eyes’’
in the film. Fish eye formation can sometimes be prevented by adding a
small amount of a chelating agent.
Besides the many antifoam agents, there are other readily available
materials which function as defoamers. Higher alcohols (such as 2–ethylhex-
anol), water-soluble oils, and tributylphosphates also have been used for this
purpose [242]. Equipment and devices used for testing foam formation are
discussed in Chap. 10; as defoamers, silicones, fatty acid derivatives, and
caprylalcohol are used. A blend of tributylphosphate/pine oil (3/1) also may
be used as a defoamer [223]. There are many antifoam agents and they
are usually classified as silicones or nonsilicones [207]. Silicone fluids and
emulsions are common defoaming agents. The silicone fluids most commonly
used in the printing ink industry are based on polydimethylsiloxane where the
chain length may vary from 1 to 1000. Antifoam agents must be used with
care, since the resulting adhesive performance characteristics are concentra-
tion dependent. An excess can cause pinholing and poor roll-to-roll ink
transfer. Silicone derivatives are used in a concentration of 0.05–0.2% [223].
Defoamers based on mineral oil may contain emulsifiers and other
defoaming agents like fatty acid esters. It is often preferable to use a
nonsilicone type to avoid film scission. Polyethylene/polypropylene glycol
ether derivatives of amines and alcohols were suggested. Propylene glycol,
diethylene glycol, and carbamide also have been used [241]. Defoamers for
labels should not influence the printing properties of the face material. They
must also meet FDA and BGA requirements. Foaming is a problem mostly
encountered with medium level viscosity formulations.
Solvents (used as defoamer) may swell the dispersion particles in the
coating, giving a more continuous dry film. Hoechst stresses the role of
solvents to improve water resistivity, especially with regard to water-
whitening. Table 8.33 lists some commonly used defoamers.
Manufacture of PSAs 537

Table 8.33 Common Defoaming Agents

Name Code Supplier Base Comments

Polymekon 1488 Goldschmidt Silicone —


derivative
— EOPK-53 Goldschmidt Silicone Not BGA-FDA
derivative approved
NOPCO 8034 Diamond — One of the most
Shamrock recommended
defoamers
BYK 040 BYK — —
Bevaloid 581 B Erbslöh — —
Dehydran G8 Henkel — —
Drewplus T-4211 Drew — One of the most
Ameroid common defoamers
used for PSA.
FDA approved
Lanco-Foamex VP-345 Langer —
Lumithene ES BASF —
Surfynol 104 Air Products Acetylenic Wetting agent
alcohol
Polymekon 81 Goldschmidt Silicone —
derivative

Adjustment of pH
Generally, water-based dispersions are stable above a pH of 2; acrylic
dispersions are usually delivered with a pH of 5–6, vinyl acetate copolymers
at a pH of 4–5, and natural rubbers and neoprene dispersions in the alkaline
pH range. Dispersions of polymers bearing carboxylic groups (CSBR or
rosin acids) are acid; their neutralization may change their stability,
adhesive, and end-use properties (e.g., tack, shear, water sensitivity, aging
resistance). In order to avoid an ionic shock, formulation components have
to display the same pH range (i.e., they need neutralization agents).
Nonvolatile neutralization agents may change the water absorption/
desorption of the polymer particles (i.e., drying properties) or may interact
with other components of the laminate; they also may change the viscosity
of the dispersion or destroy ester-based surfactants (e.g., sulfosuccinates).
The modulus, peel adhesion, and water resistance of the adhesive
depend on the nature of the neutralization agent used [243]. For neoprene
latices diethanol amine (DEA) is recommended [244]. Polychloroprene
latices are supplied with a pH of about 7. It is advisable to adjust the pH
538 Chapter 8

upwards to obtain a long storage stability and to ensure good cohesive


strength and film aging. The adjustment of the pH in order to modify the
balance between adhesive tack and cohesive strength is just one of several
formulation variables the adhesive manufacturer has available with the
carboxylated chloroprene latexes. For CSBR it is desirable to adjust the pH
upwards in order to obtain a long storage stability and good film aging
[201]. In some cases the use of thickeners makes the change of pH necessary.

Solvents
Solvents are added to adhesives in order to improve the anchorage on a
substrate [245] or to plasticize them. The drying of latex adhesives is
particularly susceptible to environmental conditions (temperature, humid-
ity). What is not so obvious is that these also affect the surface of the dry
film. Therefore coalescence aids can be added to counteract this problem.
Generally, up to 20% solvents may be added. The acceleration of the
bonding process (green strength) that occurs with polyvinyl acetate
adhesives cannot be observed for PSAs [57]. In the case of water-based
adhesives, the addition of coalescing solvents, plasticizers, etc., depends to
a large extent on their water solubility. Additives which are water soluble
should be added at 20 C, preferably diluted with an equal quantity of water.
If added at too high a temperature or concentration, they can cause
destabilization of the latex by removal of water. If the additive is insoluble in
water, it can be added while hot. The higher the addition temperature, the
faster the absorption of the solvent into the polymer particles, and the less
the effect on the particle size.
Addition of toluene to an aqueous PSA produces a swelling of
polymer particles; therefore thickening occurs and no other thickeners are
needed [122]. Solvent addition can lead to a better adhesion/anchorage of
the adhesive to special surfaces. The addition of 10–20% solvent to latex
under adequate stirring is possible [246]. Like pure solvents, rubber or
adhesive solutions (20%) may be added as well in order to get a higher solids
content. The reverse procedure—the addition of rubber latexes in adhesive
solutions—has been practiced also [247]. Solvents may improve the
coalescence and water resistance of the coated adhesive. Generally, care
should be taken with the use of water-soluble solvents.

Special Additives
Special additives are necessary in order to impart some specific end-use
properties to the dispersion. The most important of these products are the
waterproofing agents and the wet-adhesion agents. Special labels for wine
Manufacture of PSAs 539

bottles (mostly champagne) need prolonged water stability (in cold water).
Normal water-based formulations do not meet this requirement. Because of
the tension (strain) induced in the web, hot-melt or solvent-based PSAs do
not completely meet these requirements and there is a need to improve the
water resistance of the water-based adhesive. On the other hand, labeling of
wine-cellar stored, recycled, and freshly washed bottles, and deep freeze
labeling require wet adhesion (on water condensation or a water layer).
Special adhesives with a noninflammable character need special
flameproof additives. Generally, these are the same as used for solvent-
based PSAs, but their formulation is more difficult because of the low
viscosity of the aqueous PSAs (e.g., Genomoll P, stibium oxide, Al silicate).
The role of solvents as special solubilizing agents for the wetting agent
should be noted. As an example, diluted Acronal V 205 does not wet-out
solvent-based silicone release liners at a concentration of 1.2% (wet/wet)
Lumithen IRA; 5% ethyl alcohol improves the wet-out.

Solubilizing Agents
Solubilizing agents ensure solubility of the water-based adhesive layer after
storage at room temperature or elevated temperatures, and in cold or hot
water (see Section 2.11 also). These are special additives used for wash-off
labels, used mainly for recycled bottles.
Theoretically, solubilizing agents are micromolecular (wetting agents)
or macromolecular (protective colloids). From the class of protective
colloids used, there are only a few that provide good water solubility
without a loss of the tack. The most important ones are polyvinyl ether
(Lutonal M, Gantrez NM), polyvinyl pyrolidone copolymers (Antarez), and
polyacrylic acid copolymers. Their use always depends on the balance of
water solubility/tack. All these products are sensitive to bleeding, thus care
should be taken when selecting these products (solubility/bleeding); for
example, PVP-K-15 gives migration (formulated in Vinamul 4512) while
PVP-K-90 does not cause migration.
Polyvinyl ethers may be used as aqueous solutions or dispersions. They
were originally employed as tackifiers, especially for film coatings on soft PVC
because of their good plasticizer resistance. They are water soluble (Lutonal
M) and thus they impart good water solubility to the formula. Unfortunately,
they lead to strike-through (bleedthrough) and have a low aging resistance.
Solubilizers are discussed in detail in [93,248]. Such additives play a special
role in the formulation of adhesives for medical and splicing tapes.
Fungicides. Fungicides/bacteriocides can be added to the emulsion to
prevent biodeterioration. Antimicrobical agents such as p-hydroxyheptan-
phenol or p-hydroxy-cyclo-hexylphenone may be used [249].
540 Chapter 8

Compounding of Water-Based PSAs


The compounding of water-based adhesives should be carried out in a tank
at low stirrer speeds to minimize foam formation. The solids content of both
latexes and resin emulsions should be closely monitored to prevent
deviations from the intended polymer/resin ratio and to ensure consistency
of adhesive properties. Generally, the dispersion additives (surfactants,
defoamers) are fed in at the end of the formulation process. Because of the
absorption phenomena of defoamers, they should be added stepwise. In
some cases a separate preparation of the surfactant solution, antioxidant
solution (dispersion), or filler slurry is needed. Equipment for blending of
water-based PSAs is discussed in detail in [93]. The main suppliers of
elastomer/tackifier dispersions provide technical information concerning
the storage/use of their products [250,251]. The main characteristics of the
equipment used are given below.

Tanks. Either vertical or horizontal designs are acceptable. Filament


wound, glass fiber-reinforced polyester tanks are recommended for their
relatively low cost, easy installation, and excellent resistance to chemical
attack. Other construction materials also are acceptable. Stainless steel
tanks are more expensive than other types but have proven durability and
inertness to polymer emulsions and are easily cleaned. Aluminum tanks may
be used for some emulsions, but are darkened by others. Some acrylic
emulsions (mostly tackified ones) must be agitated continuously to ensure
homogeneity. A mixer entering on the side (300–500 rpm) or a vertical shaft
agitator (with a peripheral velocity below 66 m/min) may be used. The space
above the emulsion must be humidified to prevent the formation of skin
over the surface of the liquid as well as on the walls of the tank. In some
cases steam is injected for 3–5 min every 8 hr [252].

Pumps. Low shear pumps should be used. Diaphragm pumps exert


the lowest shearing force. They are particularly suited to transfer highly
loaded formulated emulsions containing resins or other additives. Pumps
should always be operated under conditions of flooded suction. Generally,
minimum agitation, turbulence, or pulsation is recommended.

Pipelines and Filters. Pipelines of steel or rigid PVC may be used. All
pipes should be easily accessible for cleaning. Outside lines should be insu-
lated and heated. When polymer emulsions are withdrawn from storage
tanks, skin and foreign matter must be removed before they pass through
pumps, meters, and other equipment. A basket strainer having about
0.09 m2 of filtering area provides adequate capacity for this task. Removable
inserts are convenient for cleaning purposes. Stainless steel screens
Manufacture of PSAs 541

(14–20 mesh) suffice for coarse straining. For still finer filtration, cartridge
filters with disposable inserts are employed. Generally, filters with filter bags
(like nylon) should be used. Working pressure for 200 mesh filters is about
140–180 Pa (to start).

2.13 Technological Considerations


Formulating solvent-based or water-based PSAs needs no special technolo-
gical equipment. Formulating water-based dispersions should be carried out
taking into account the shear sensitivity of the dispersed system. When for-
mulating hot-melt PSAs, one should consider the temperature and aging
sensitivity of these materials.
There are different technical possibilities to compound the tackifier
with the raw PSA materials. The tackifying technology depends on the
tackifier, PSA, and adhesive end-use properties of the laminate. Hot-melt
PSAs require solid and/or liquid tackifiers, as do solvent-based PSAs; water-
based PSAs are usually formulated in combination with water-based
tackifiers. Each of the current tackifying technologies show some advantages
and disadvantages.

Tackifying with Water-Based Tackifiers


From a technical point of view this remains the most simple technology.
Demixing problems during feed-in can be avoided. No special knowledge
for the converter is needed. The coating properties (wet-out, mechanical
stability) are the best; no flammable agents are used. Current dispersion
coating machines may be used without special precautions.
On the other hand, aqueous resin-tackifier dispersions show some weak-
nesses also. These can affect the properties of the final coating in areas such as
a lower water resistivity of the coating, a lower anchorage of the adhesive, a
higher penetration or bleeding of the adhesive, lower shear, and poorer die-
cutting. The water-based resin dispersions are secondary dispersions made
with surface-active agents (2–14%). This additional amount of surface-active
agent in the final coating increases its solubility and hygroscopy. Concerning
the influence of the emulsion character of the tackifier on the adhesive
properties, there are divergent data in the literature. It is assumed that the
adhesive properties of the emulsion-tackified PSAs are inferior to those of
solvent-based resin-tackified PSAs. This hypothesis is illustrated by the
limited amount of tackified water-based PSAs for film coating. In this case the
loading of the tackifier remains limited (<10%). Therefore solvent-based
tackifiers are used more often. On the other hand, it was shown that with the
542 Chapter 8

exception of a slight change in rolling ball tack, there is little difference


between the same adhesives coated with water and solvent carriers.

Tackifying with Solvent-Based Tackifiers


Solvent-based tackification is an old technology enjoying some renaiss-
ance because of the superior water resistivity and (perhaps) tack of the
solvent-based resin-tackified coatings. The anchorage of the directly coated
solvent-based resin-tackified coatings also appears superior. Unfortunately
there are some technological disadvantages associated with solvent-based
resin tackifying. First, the manufacture of the concentrated (70–80%) resin
solutions causes difficulties for a converter without chemical knowledge.
Furthermore the formulation (feed-in) technology of the resin solution in
the aqueous dispersion requires special knowledge and equipment. Because
of the low level of surface-active agents the mechanical stability of the
solvent-based resin-tackified dispersion is low; the wet-out of the dispersion
is inferior and generally does not allow transfer coating on solventless
silicones. Theoretically, no special converting machines (e.g., explosion
proof) are required for solvent-based resin-tackified formulations; practi-
cally, some precautions are needed. Except for the water resistivity, no
improvement in the peel/tack/shear balance for solvent-based resin-tackified
water-based acrylic PSAs is observed.

Tackifying with Liquid Resins


Tackifying with liquid, unformulated resins is the most simple method to
feed-in the tackifier (no pH or viscosity regulation necessary). Unfortunately
the amount of the liquid tackifier which can be added remains limited
because of the strike-through of the tackifier. Loading more than 10% liquid
tackifier causes migration or bleeding because of the pronounced decrease
of the cohesion. Different liquid resins were tested (Resamin, Cymel,
polyethyleneimine, Hercolyn, etc.); the best results were obtained with a level
of 5–10% hydrocarbon-based resins, mainly with regard to the tack.

Tackifying with Molten Resins


There are only a few aqueous dispersions with sufficient heat stability (at the
boiling temperature); some copolymers of chlorine-butadiene (e.g., DuPont
CR-115) meet these requirements. Thus it is possible to add the solid high
softening point tackifier as a (molten) liquid into these dispersions; these
formulations display a higher shear strength. Unfortunately these special
polymers are very expensive. On the other hand, special equipment is
required for correctly dispersing the molten resin. Experience shows that
Manufacture of PSAs 543

Table 8.34 Peel Improvement After Longer Storage


Time of the Formulated Adhesive

Storage Time

Adhesive Characteristics 1 hr 1 day 1 week

Coating weight (g/m2) 24 28 25


Peel (N/25 mm), glass 25 PT 26 PT 20 PT
Peel (N/25 mm), PE 18 25 PT 20 PT
Shear resistance (min) 45 46 31
Formulation: Acronal V 205/Snowtack CF 52; face stock: paper;
room-temperature shear measured on stainless steel.

more than 30–40% loading (by weight) of Neoprene CR-115 (Dow) is


required for current acrylic PSAs in order to raise the cohesive strength.

Special Features
The age of the formulated (tackified) water-based acrylic PSAs influences
the adhesion/cohesion balance. Polyethylene peel increases as a function of
dwell time (Table 8.34). A dwell time of several hours (optimally 24 hr) is
suggested.

2.14 Comparison Between Solvent-Based, Water-Based,


and Hot-Melt PSAs
In Chap. 6 the properties and performance characteristics of rubber-resin-
based and acrylic-based PSAs were compared. Next, a comparative study of
the PSAs with a different physical state (i.e., solvent-based, water-based, and
hot-melt PSAs) evaluates the formulating features of different raw materials,
taking into account their adhesive and other properties.

Solvent-Based Formulations
Solvent-based acrylic PSAs have the following strengths [253]: a high water
resistance, an improved clarity for polyester overlays, and an extremely high
temperature resistance. The drawbacks of solvent-based acrylic PSAs
include the following:
lack of compliance with environmental regulations;
need for incineration or solvent recovery with accompanying
capital expenditure;
fire hazards/static electricity;
544 Chapter 8

higher insurance rates;


future availability of solvents.
Solvent-based adhesives generally have been the preferred choice of
converters because of their ease of application and their desirable balance
of PSA performance properties. However, there remain significant problems
associated with these solvent systems. These include high costs, environ-
mental and toxicity concerns, poor shelf life, and very real fire hazards
because of the solvents. Both hot-melt PSAs and water-based PSAs were
developed to overcome these concerns.
Of the two alternatives, hot-melt PSAs (particularly rubber-based)
have generally been plagued by relatively poor oxidation stability, marginal
PSA performance (especially at elevated temperatures), and high initial
application replacement costs. In contrast, emulsions are water-based
systems which present no major environmental or toxicity concerns; they
also pose no fire hazard associated with the volatiles present. Emulsion
systems possess a further advantage in their ability to be coated on coating
equipment previously used for solvent-based products. Because of the fire
and explosion hazards, there is a need for the use of special ventilating and
explosion-proof equipment when coating solvent-based PSAs. In the large
scale use of solvent-based adhesives the removal of organic solvent during
drying requires special solvent recovery equipment to avoid problems
of air pollution. Solvent emission restrictions, increasing energy costs, the
potential decreased availability of energy, rising solvent prices, and more
stringent health and safety regulations make the availability of safer, less
energy intensive alternatives urgent [50].
Solvent-based adhesives generally display good adhesion to a broader
range of film substrates than water-based coatings [254]. Solvent-based
rubber-resin PSAs have the following benefits: as fully commercial products
they possess aggressive adhesion, good cohesion, and water resistance. The
limitations include poor aging, a low plasticizer resistance, low temperature
adhesion, and solvent emission. Solvent-based rubber PSAs coat easily and
fast at lower coating weights. Owing to the usually low solid content, high
coating weights require a reduction of the coating speed. The viscosity of
solvent systems cannot easily be changed which implies that the coating
head needs to be adjusted to the actual viscosity. Explosion hazards remain
real and a solvent recovery installation is required. Wetting is not a problem,
fish eyes seldom occur, and the resulting coating is smooth. Die-cutting and
slitting may be critical since the adhesive is often somewhat elastic, soft, and
exhibits stringing.
Solvent-based acrylic PSAs wet the web easily, do not form fish eyes,
and lead to smooth, transparent coatings. They allow fast machine speeds,
Manufacture of PSAs 545

but at higher coat weights (as in tapes) the speed should be reduced in order
to prevent explosion; explosion-proof equipment and solvent recovery are
required. The finished products are relatively hard, so slitting or die-cutting
normally present no problems. Solvent acrylic PSAs have excellent aging
characteristics and resistance to elevated temperatures and plasticizers; they
also offer the best balance of adhesion and cohesion, as well as an excellent
water resistance. Acrylic PSAs are harder than rubbers; this can be seen in
the less aggressive tack and slower build-up of peel strength. Lower
adhesion to nonpolar polyolefins is caused by the polar chemistry of acrylic
PSAs. Solutions of acrylic-based PSAs have the following advantages:
no degradation in air, ozone, sunlight, heat;
elevated temperature strength;
low temperature flexibility retention and bonding;
resistance to oil, fuel, and chemicals;
resistance to plasticizer migration (from PVC);
good adhesion to metals, plastics, and other substrates.
The weaknesses include rather low quick stick values and softening by
oxygenated solvents (glycols, alcohols).
In general, solvent coatings are free of migrating species of hydrophilic
components which can cause humidity-induced variations in release
properties. Solvent coatings also dry faster because of the low boiling
point or the low heat of evaporation of the solvent used. Solvent rubber-
based adhesives may cause problems when printing and die-cutting labels
due to legging [87]. Normally, solvent coatings contain 40% or less solids
because of viscosity constraints. Equipment clean-up requires the use of
solvents. Contrary to conventional wisdom, the drying costs are greater for
solvent coatings. While the heat of evaporation of hydrocarbon solvents is
much less than that of water, a greater amount of heated air must be passed
through the oven in order to operate below the lower explosion limit.
Table 8.35 summarizes comparatively the most important properties
of solvent-based, water-based, and hot-melt PSAs [135]. In Table 8.36 the
advantages and disadvantages of water-based and solvent-based adhesive
systems are listed. The comparison of the technical and economical
characteristics of PSA labels as a function of the physical state of the
adhesive is summarized in Table 8.37 [255].

Water-Based PSAs
The replacement of solvent-based adhesives with aqueous latexes is a conti-
nuing trend in the industry, for reasons of air quality, safety, and economics.
However, the transition has not been simple and is by no means complete.
546 Chapter 8

Table 8.35 Comparison of the Properties of Water-Based, Solvent-Based, and


Hot-Melt PSAs

PSA Label Adhesive Technology

Water- Solvent- Hot-


Application field Properties based based melt

Permanent Label converting E G E


Adhesion E G E
Application temperature range E E G
Formulation flexibility G E F
Removable Label converting G F F
Adhesion G G F
Application temperature range G E F
Formulation flexibility F E F
Cold temperature Label converting E F F
Adhesion G F F
Application temperature range G G F
Formulation flexibility E G F
Clear film Label converting G E F
Adhesion G E F
Application temperature range G G F
Formulation flexibility F G F
E ¼ Excellent; F ¼ Fair; G ¼ Good.

Table 8.36 General Comparison of Solvent Acrylic and Emulsion Acrylic PSA

Acrylic PSA

Properties Solvent-based Emulsion

PSA performance Good to excellent Good to excellent


Aging properties Excellent Excellent
UV stability Excellent Excellent
Thermal stability Excellent Excellent
Oxidation stability Excellent Excellent
Bleeding/migration Good Fair
Edge ooze Good Fair
Solvent resistance Good Good
Water resistance Good Fair to poor
Crosslinkability Excellent Fair
Clarity Excellent Good
Coating properties
Foaming Not applicable High
Wet-out Excellent Poor
Manufacture of PSAs 547

Table 8.37 Technical and Economical Performance of PSA Labels as a Function


of the Physical State of the Adhesive

Physical State of the PSA

Properties Solution Emulsion Hot-melt

Adhesive properties Tack Excellent Good Excellent


Peel Excellent Excellent Good
Shear Excellent Good Poor
End-use properties Aging Excellent Good to fair Poor
Versatility Excellent Good Poor
Bleeding Good Fair Fair to poor
Converting properties Die-cutting Good Fair Poor
Printing Excellent Fair Good
Cost Very high Moderate Low

Aqueous adhesives have found limited acceptance, being regarded as


generally inferior to organic solution adhesives either in performance or
coating rheology or both [107].
The deficiencies of common aqueous PSAs may be summarized as
follows [256–258]: the inferior balance of tack, peel, and shear properties,
the low water and humidity resistance, the bad plasticizer migration
resistance, the limited heat resistance, and concerns about machinability
and/or coatability. The poor coating rheology typical of conventional latex
adhesives is manifested in several forms: first, there is a tendency in roll
application to form ridges parallel to the coating direction. Ridges adversely
affect adhesive performance by reducing the effective area of contact with
the substrate, but more substantially, destroy the optical appearance
(transparency), particularly when a clear film is used. Other disadvantages
include the shear rate and water sensitivity. High relative humidity may
cause liquid components to migrate through label papers, so that staining
appears, a common characteristic of several water-soluble adhesives [156].
Special emulsion polymerization techniques allow properties which
cannot be acquired through other systems (e.g., solvent-based or hot-melt).
The core/shell microstructure of the polymer particles (core/shell structured
particle blends) produces a latex which shows evidence of the presence of a
copolymer in its films, although the particles do not contain any copolymer
in the emulsified state [259].
As far as adhesive properties of the water-based PSAs are concerned,
they do not offer any advantages versus solvent-based PSAs. The typical
room temperature shear values from the solvent-borne adhesives are
significantly higher than the typical values for water-borne adhesives
548 Chapter 8

[260]. The shear resistance of solvent-based CSBR PSAs is higher than that
of water-based ones, theoretically, water-based systems can yield superior
shear performance, relative to solvent-based systems [6]. According to [59]
for gel free low-Tg acrylics there was no significant difference between
emulsion versus solvent-borne PSA film adhesive performances.
The emulsion polymerization does not impose viscosity/molecular
weight restrictions, because the polymers are formed as discrete particles
suspended in the medium. The high molecular weight achieved during
emulsion polymerization provides suitable performance characteristics for a
wide variety of PSA applications without necessitating further modification
[138]. An inherent advantage of the emulsion polymerization method is that
crosslinking can be built into the polymer at the time of polymerization.
A number of difunctional crosslinking monomers may be used such as
1,6-hexanediol diacrylate, ethylene glycol diacrylate, butanediol dimeth-
acrylate, N-methylol acrylamide, and triethylene glycol dimethacrylate. In
practice, the performance of water-based PSAs is reduced by the composite
structure.
Performance disadvantages of water-based acrylic PSA technology
include the need to improve the machine coatability, the lower adhesion to
plastic substrates, the lack of clarity of the adhesive film, and the relatively
low water resistance. The advantages of water-based coatings are evident;
lower dry weight costs, easy clean-up, and environmental acceptance.
Furthermore, they cause no pollution, require reduced energy for drying,
cost less than solvent types, and hence they displace some rubber-based
polymers. Water-based systems may be dried at high speeds when the web
allows high temperatures (e.g., release paper); explosive hazards do not
exist. However, wetting of the web is often a tricky problem; much work
is being carried out in the area of wetting agents to improve the wet-out
and minimize effects on performance. Most adhesives are fairly hard and
cause no problems during slitting or die-cutting, unless they are highly
tackified. An advantage of the water-based adhesives is their high
solids content. Rubber solutions generally possess a minimum of 15–20%
solids, water-based latexes have a solids content of at least 35%, while
solution polymers generally have a narrower MWD than emulsion
polymers.
The difficulties involved in the transition to water-based coating
include [259]: wetting, coalescence, and technological problems, as well as
concerns about the properties of the coating. The surface tension of water is
72 mN/m, while a typical solvent has a surface tension of 25–30 mN/m. For
many webs having a surface energy of 30–40 mN/m, wetting becomes
problematic. Film forming from aqueous media requires coalescence. A
switch from solvent-based products to water-borne products often results
Manufacture of PSAs 549

in longer ovens and the corrosion of piping, pumps, reservoirs, etc.


Furthermore, the dried films have inherent water sensitivity; also, the high
surface tension of water results in higher power consumption, increased
machine wear, and waste streams which cannot be incinerated or dumped
into a sewage system.
Therefore the following general requirements towards aqueous
dispersions [260] can be stated:
Dispersing high molecular weight polymers can achieve a higher
solids content than solution adhesives, thus maximizing the cohesive
strength.
Dispersions should have the viscosity and general flow properties
necessary to allow efficient film coating with minimal striations on
conventional coating equipment.
There should be adequate wetting and no crosslinking of the film
between the coating head and the drying tunnel.
There should be low viscosity and maximum flow coupled with the
highest possible solids content for optimum drying efficiency.
High mechanical stability of the adhesive dispersion is an absolute
necessity for efficient and trouble-free machine coating.
The concerns include the tendency to foam and the presence of
migrating components. Care must be taken in the selection of defoamers
which can exert a negative effect on adhesive and/or release properties.

Hot-Melt PSAs
Hot-melt PSAs offer all (or most) of the advantages of solution PSAs (i.e.,
no vapor hazard nor fire hazard); they require a low processing energy and
no special storage conditions, and allow unlimited coating possibilities
(speed, thickness). However, they also possess some weaknesses, such as the
need for specially designed equipment, an elevated application temperature
with higher processing costs, and process sensitivity; they remain inferior in
heat-sensitive coating applications and display the lowest heat resistance.

REFERENCES

1. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,


Marcel Dekker, New York, 1997, Chap. 6.
2. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chap. 3.
3. E. Djagarowa, W. Rainow and W. L. Dimitrow, Plaste u. Kaut., (2) 100 (1970);
in Adhäsion, (12) 363 (1970).
4. Coating, (6) 175 (1972).
550 Chapter 8

5. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 375.


6. J. Kendall, F. Foley and S. G. Chu, Adhes. Age, (9) 26 (1986).
7. A. Zawilinski, Adhes. Age, (9) 29 (1984).
8. Coating, (7) 184 (1984).
9. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 2.
10. J.P. Keally and R.E. Zenk (Minnesota Mining and Manuf. Co., St.Paul,
MN,USA); Canad. Pat., 1224.678/19.07.82 (USP. 399350).
11. J.A. Schlademan, Coating, (1) 12 (1986).
12. J.C. Fitch and A. M. Snow, Adhes. Age, (10) 24 (1977).
13. Die Herstellung von Haftklebstoffen, T 1.2.2; 15d, BASF, Ludwigshafen,
Nov.,1979.
14. A. Midgley, Adhes. Age, (9) 17 (1986).
15. C.W. Koch, J. Polymer Sci., (C3) 139 (1963).
16. P.J. Counsell and R.S. Whitehouse, in Development in Adhesives (W.C. Wake,
Ed.), Vol. 1, Applied Science Publishers, London, 1977, p. 99.
17. A. Bull, Shell Bulletin, TB, RBX/73/8/6.
18. R. E. Downey, Adhes. Age, (3) 35 (1974).
19. Coating, (1) 16 (1984).
20. A. W. Bamborough, 16th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1991, p. 96.
21. Coating, (1) 13 (1984).
22. Coating, (18) 240 (1972).
23. G. Ruckel, Coating, (1) 18 (1984).
24. P.A. Mancinelli, New Development in Acrylic HMPSA Technology, p. 165.
25. R. Mudge, ‘‘Ethylene-Vinylacetate-Based, Water-Based PSA,’’ in TECH 12,
Advances in Pressure Sensitive Tape Technology, Technical Seminar Proc.,
Itasca, IL, May 1989.
26. E. G. Ewing and J. C. Erickson, Tappi J., (16) 158 (1988).
27. Adhäsion, (4) 24 (1985).
28. Adhäsion, (9) 352 (1966).
29. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 1.
30. J. Class and S.G. Chu, Coat. Appl. Polymer Sci., Proc., (48)126 (1989).
31. R. Sattelmeyer, Adhäsion, (10) 278 (1976).
32. C. A. Dahlquist, in Handbook of Pressure Sensitive Technology (D. Satas, Ed.),
Van Nostrand Rheinhold Co., New York, 1982, p. 82.
33. Coating, (7) 186 (1984).
34. J. Villa, Adhäsion, (10) 284 (1977).
35. G. W. Drechsler, Coating, (3) 98 (1987).
36. F. H. Wetzel, Rubber Age, (82) 291 (1957).
37. C. W. Koch and A. N. Abbott, Rubber Age, (82) 471 (1957).
38. H. Hultsch, Farbe u. Lack, (77) 11 (1971).
39. E. De Walt, Adhes. Age, (3) 38 (1970).
40. J.A. Schlademan and J.G. Bryson, Proc. of the 21th Annual Meeting of the Adh.
Soc., Savannah, GA, February, 1998, in [40].
Manufacture of PSAs 551

41. J.A. Schlademan, ‘‘The Role of Tackifier Compatibility and Molecular Weight
on Bulk Properties of PSAs,’’ Proc. of the 22nd Annual Meeting of the Adh.
Soc., Panama City Beach, FL, Febr. 21, 1999, p. 75.
42. Adhes. Age, (12) 35 (1987).
43. K.H. Schumacher and T. Sanborn, ‘‘UV-Curable Acrylic Hot-Melt for Pressure
Sensitive Adhesives-Raising Hotmelts to a New Level of Performance,’’ Proc.
24th Annual Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg, VA, p. 165.
44. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 193.
45. G. W. Drechsler, Coating, (2) 52 (1987).
46. Adhäsion, (5) 162 (1971).
47. R. Houwink and G. Salomon, in Adhesion and Adhesives, Vol. 2, Elsevier Publ.
Co., Amsterdam, 1967, Chapter 17.
48. M. Sheriff, R.W. Knibbs and P.G. Langley, J. Appl. Polymer Sci., 17, 3423
(1973).
49. G. Kraus and K. W. Rollman, J. Appl. Polymer Sci., (23) 3311 (1977).
50. T.G. Wood, Adhes. Age, (7) 19 (1987).
51. Coating, (7) 184 (1984).
52. S.W. Medina and F. W. Distefano, Adhes. Age, (2) 18 (1989).
53. P. Green, Labels and Labeling, (11/12) 38 (1985).
54. British Petrol, Hyvis, Technical Bulletin (1985).
55. P. Dunckley, Adhäsion, (11) 19 (1989).
56. Adhäsion, (12) 390 (1969).
57. Adhäsion, (12) 430 (1972).
58. M. Kajiyama, ‘‘Phase Structure of PSA Prepared from Solution and
Emulsion,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001,
Williamsburg, VA, p. 283.
59. S. Tobing et al., J. Appl. Polym. Sci., 76, 1965 (2000); in S. D. Tobing and
A. Klein, ‘‘Synthesis and Structure Property Studies in Acrylic Pressure-
Sensitive Adhesives,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25–28,
2001,Williamsburg, VA, p. 131.
60. T.G. Wood, Adhes. Age, (7) 19 (1987).
61. M.F. Tse and K.O. McElrath, European Tape and Label Conference, Exxon,
Brussels, April 1989, p. 91.
62. S.D. Tobing, O. Andrews, S. Caraway, J. Guo, A. Chen and Shelly Anna,
‘‘Shear Stability of Tackified Acrylic Emulsion PSAs,’’ Proc. 24th Annual
Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg, VA, p. 273.
63. Adhes. Age, (11) 40 (1988).
64. Allg. Papier Rundschau, (16) 462 (1986).
65. T. Yoshida, European Tape and Label Conference, Exxon, Brussels, April
1989, p. 69.
66. O. Cada and P. Peremsky, Adhäsion, (5) 19 (1986).
67. R. Schuman and B. Josephs (Dennison Manuf. Co., USA), PCT/US86/02304/
25.08.86.
68. J.M. Avenco, Resin Review, (3) (1972).
552 Chapter 8

69. J. Lin, W. Wen and B. Sun, Adhes. Age, (12) 22 (1985).


70. A. L. Bull, Shell Elastomers, Thermoplastic Rubbers, Technical Manual, TR
8.12, p. 13.
71. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 194.
72. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 227.
73. R.W. Rance, E. Lazarus and F. Wilkes, Labels and Labeling, (2) 16 (1988).
74. R. Jordan, Coating, (10) 278 (1982).
75. R. Dörpelkus, Allg.Papier Rundschau, (16) 456 (1986).
76. W. Druschke, ‘‘Adhesion and Tack of PSA,’’ AFERA Meeting, Edinburgh,
October 1986.
77. W. Blume, Adhäsion, (7/8) 20 (1983).
78. J.F. Kwiatek, Adhes. Age, (11) 28 (1988).
79. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter
4.1.3.
80. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 319.
81. L.C. Broggs, Adhes. Age, (5) 19 (1983).
82. M.E. Ahner, H.L. Evans, S.G. Hentges and M.R. Gerstenberger, Coating,
(10) 330 (1986).
83. Coating, (10) 304 (1969).
84. Coating, (2) 46 (1970).
85. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 197.
86. S. Pila, Defazet, (2) 54 (1974).
87. H. Muller, J. Türk and W. Druschke BASF, Ludwigshafen, EP.0.118.726/
11.02.1983.
88. Coating, (1) 14 (1987).
89. Mowilith DM 104, Kunstharze Hoechst, Hoe, KGM, 3070d., 1992.
90. B. Henzel, Plast-Europe, (3) 86 (1982).
91. Mowilith DM 56, Technical Data Sheet, Hoechst.
92. Adhes. Age, (10) 36 (1977).
93. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter
4.1.4.
94. L. Jacob, ‘‘New Development of Tackifiers for SBS Copolymers,’’ 19th
Munich Adhesive and Finishing Seminar, Munich, Germany, 1994, p. 107.
95. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 198.
96. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 221.
97. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 336.
98. Coating, (3) 79 (1974).
99. Firestone, Technical Service Report, 6110, August 1986.
100. L. Krutzel, Adhes. Age, (9) 21 (1987).
101. Converting and Packaging, (3) 24 (1986).
102. C. Massa, Coating, (7) 239 (1989).
103. C. Zang, US Pat. 3.532.652.
104. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 55.
105. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 7.
Manufacture of PSAs 553

106. H. Salmen and M. Htun, Allg. Papier Rundschau, (35) 28 (1986).


107. C.P. Iovine, S.J. Jer and F. Paulp (National Starch Chem. Co., Bridgewater,
USA), EP 0.212.358/04.03.87, p. 3.
108. C.T. Albright, D.S. Culp (Avery Internat. Co., USA) EP 0099087B1/
23.12.1987.
109. G.R. Hamed and C.H. Hsieh, J. Polymer Physics, 21,1415 (1983).
110. R.P. Mudge (National Starch Chem. Co., Bridgewater, USA), EP 0.225.541/
11.12.1985.
111. J. Lechat, ‘‘The Pressure Sensitive Labelstock Market in Western Europe,’’
Finat World Congress, Monaco, 1989.
112. M. Mazurek (Minnesota Mining and Manuf. Co., St.Paul,MN,USA), EP
4.693.935/15.09.1987.
113. J. Lin, W. Wen and B. Sun, Adhäsion, (12) 21 (1985).
114. L. Jacob, 8th Munich Adhesive and Finishing Seminar, Munich, Germany,
1983, p. 42.
115. Coating, (10) 366 (1988).
116. Vinnapas, Eigenschaften und Anwendung,7.1. Teil, Anwendung, Wacker,
München, 1976.
117. K. Goller, Adhäsion, (4) 101 (1974).
118. T.J. Bonk and S.J. Thomas (Minnesota Mining Manuf. Co., St. Paul, MN,
USA), EP 120.708.31/03.10.1984.
119. Adhes. Age, (5) 54 (1987).
120. I.J. Davis (National Starch, Chem.Co., Bridgewater, USA), U.S. Pat.
4.728.572 /01.03.1988.
121. DE-AS 2407484/EP 0118726.
122. Die Herstellung von Haftklebstoffen, T1.-2, 2-14d, Dec. 1979, BASF,
Ludwigshafen,Germany.
123. J.L. Wacker and P.B. Foreman, CAS, Colloids, (1) 5 (1988).
124. Silikone, Sitren, 532276, Technical Information, Th. Goldschmidt A.G.,
VK-Silikone, Essen, 1986.
125. Adhäsion, (7) 242 (1972).
126. Coating, (9)247 (1985).
127. J.R. Creasney, D.B. Russel and M.P. Wagner, Am. Chem. Soc., Rep., 35, 1988.
128. J. Hansmann, Adhäsion, (10) 360 (1970).
129. B. Mayer, Coating, (4) 111 (1969).
130. F. Hartmann, Coating, (1) 7 (1969).
131. Coating, (11) 338 (1973).
132. J. Young, Tappi J., (5) 78 (1988).
133. G. Grove, Adhes. Age, (6) 39 (1971).
134. S. Tadashi, M. Noboyuki, I. Yoshimide and T. Toyokichi, CAS, Crosslinking
Agents, 14,6 (1988), 108.222762n.
135. T.H. Haddock (Johnson & Johnson, USA), EP 0.130.080 Bl/02.01.1985.
136. P.J. Moles, Polym. Paint. Colour J., (178), 4209 (1988).
137. M. Hiroyasu, T. Kitazaki, Y. Truma, M. Tetsuaki and K. Yunichi, DE
354486861/18.12.1985.
554 Chapter 8

138. Eastman Kodak, US Pat. 359.943; in Adhäsion, (5) 328 (1967).


139. M. Hasegawa, US Pat. 4.460.634/17.07.84; in Adhes. Age, (3) 22 (87).
140. P. Tkaczuk, Adhes. Age, (8) 19 (1988).
141. F.W. Brown, W.St. Paul and L.E. Winslow, US Pat. 4.629.663; in Adhes. Age,
(5) 15 (1987).
142. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 5.
143. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.
144. Kimberley Clark Co., US Pat. 799.429; in Coating, (1) 9 (1970).
145. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p. 383.
146. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 426.
147. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 5.
148. Adhäsion, (9) 349 (1965).
149. Adhes. Age, (10) 16 (1977).
150. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p.325.
151. J. Harrison, J.F. Johnson and W. Rossyates, Polym. Eng. and Sci., (14)
865(1982).
152. S. Mitton and C. Mak, Adhe. Age, (1) 15 (1983).
153. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 202.
154. Offset-Technik, (8) 11 (1986).
155. Coating, (8) 240 (1972).
156. Adhes. Age, (11) 28 (1983).
157. W. Wittke, Coating, (9) 334 (1987).
158. W. Wittke, Coating, (7) 249 (1987).
159. W. Wittke, Coating, (8) 206 (1987).
160. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 367.
161. J. Paris, Papier u. Kunststoffverarb., (9) 53 (1988).
162. R. M. Friedrich, European Tape and Label Conference, Exxon, Brussels,
April, 1989, p. 181.
163. H.J. Fricke, Allgem. Papier Rundschau., (24/25) 720 (1983).
164. F. King, Coating, (2) 59 (1970).
165. R. Jordan, Coating, (2) 37 (1986).
166. D. Krüger, Kaut., Gummi, Kunstst., (6) 549 (1988).
167. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 4.1.
168. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.
169. Shell Elastomers, Technical Handbuch, TR 5.1.1. (G).
170. Shell Elastomers, Cariflex TR, Merkblatt TR 5.2.1. (G), p. 2.
171. E. Diani, A. Riva, A. Iacono and E. Agostinis, 16th Munich Adhesive and
Finishing Seminar, Munich, Germany, 1994, p. 77.
172. R. A. Fox, Adhes. Age, (10) 35 (1977).
173. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p. 429.
174. A. Patel and R. Thomas, Tappi J., (6) 166 (1988).
Manufacture of PSAs 555

175. H. Hadert, Coating, (7) 203 (1970).


176. D. De Jaeger, 11th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1986, p. 87.
177. Coating, (12) 344 (1984).
178. Adhes. Age, (8) 35 (1983).
179. Celanese Resins Systems, Technical Information, LHM, 991.
180. Coating, (3) 68 (1985).
181. C.N. Clubs and B.W. Foster, Adhes. Age, (11) 18 (1988).
182. European Adhes. Seal., (9) 3 (1987).
183. F.C. Jagisch, ‘‘Recent Developments in Styrene Block Copolymers for Tape
and Label PSAs,’’ European Tape and Label Conference, Exxon, Brussels, 1993.
184. F. Jagisch and L. Jacob, ‘‘Advances in Styrene Block Copolymer
Technology,’’ PSTC Technical Seminar, Tech. XIV, May 2, 1991.
185. D. De Jager and J.B. Bortwyck, Thermoplastischer Kautschuk,Technisches
Handbuch, TR 8.11 (G), Shell Res. B.V.
186. Papier u. Kunststoffverarb., (9) 48 (1990).
187. R. Hinterwaldner, Coating, (7) 176 (1980).
188. H. Röltgen, Coating, (11) 400 (1986).
189. R. Hinterwaldner, Coating, (7) 250 (1991).
190. ‘‘UV-Curable Acrylic Hot-Melts for PSA Application,’’ BASF Symposium,
2/15/96; in [59].
191. G. Auchter, J. Barwich, G. Remmer and H. Jager, Adhäsion, (1/2) 14 (1993).
192. S. J. Piaseczinsky, Adhesives & Sealants Council, 4/89,Reprint.
193. R. Jordan, Coating, (2) 27 (1984).
194. F.V. Di Stefano, S.W. Medina and B.R. Visayendran, Sampe J., (7/8) 27 (1988).
195. H. Yang, L. Jacob and L. Heymans, 15th Munich Adhesive and Finishing
Seminar, Munich, Germany, 1990, p. 92.
196. R. Hinterwaldner, Coating, (5) 198 (1991).
197. Technische Informationsblätter, Kautschuk Latices, Nr 11/2, Aug. 1977,
Bayer, Leverkusen,Germany.
198. M. Takuhiko, Setchaku, (9) 389 (1987); in CAS,7, 5 (1988).
199. ‘‘Shear Stability of Tackified Acrylic Emulsion PSAs,’’ Proc. 24th Annual
Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p. 273.
200. Coating, (4) 11 (1974).
201. Celanese Chemical Company, Acrylate Esters, 4M/5-71, Technical Bulletin.
202. J. Johnston, Adhes. Age, (12) 24 (1983).
203. H. Toyama, T. Ito and H. Nakatsuma, J. Appl. Polymer Sci., (17) 3495 (1973).
204. W.R. Dougherty, 15th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1990, p.70.
205. H. Saxen, Etiketten, (4) 9 (1994).
206. Coating, (5) 121 (1971).
207. Converting and Packaging, (3) 26 (1986).
208. Adhäsion, (11) 481 (1967).
209. C.L. Zao, Y. Holl, T. Pith and M. Gambia, Colloid Polym. Sci., (9) 823
(1987).
556 Chapter 8

210. Product and Properties Index, Polysar, Arnhem, 02/1985.


211. Surfynol Technical Bulletin, Air Products and Chemicals Inc. (1991).
212. M.H. Tanashi, K. Yasuaki, M. Tetsuaki and K. Yunichi (Nichiban Co,
Japan), OS/DE 3.544.868 Al/15.05.1985.
213. J.Delgado et al., Proc. Of the Int. PSA Technoforum, Tokyo, Japan, 1997,
p. 33.
214. J. Delgado et al., Proc.19th Adhesion Soc. Meeting, 1996, p.342; in S. D.
Tobing and A.Klein, ‘‘Synthesis and Structure Property Studies in Acrylic
Pressure-Sensitive Adhesives,’’ Proc.24th Annual Meeting of Adh. Soc., Febr.
25–28,2001,Williamsburg, VA, p. 131.
215. A.C. Makati, Tappi J., (6) 147 (1988).
216. J.R. Snuparek, A. Bidman, J. Hanus and B. Hajkova, J. Appl. Polymer Sci.,
28, 1421 (1983).
217. A. Zosel, Colloid Polymer Sci., 263, 541 (1985).
218. Adhes. Age, (9) 15 (1976).
219. L. A. Rutter, Adhäsion, (6) 180 (1971).
220. J. Hansmann, Adhäsion, (4) 21 (1985).
221. E. Bohmer, Norsk Skogindustrie, (8) 258 (1968).
222. Adhes. Age, (12) 36 (1986).
223. Coating, (4) 118 (1969).
224. R. Pfister, Coating, (6) 171 (1969).
225. Coating, (6) 177 (1969).
226. O. Quadrat, J. Appl. Polymer Sci., 35,1 (1988).
227. Coating, (7) 120 (1983).
228. Coating, (7) 20 (1984).
229. G. Meinel, Papier u. Kunststoffverarb., (10) 26 (1985).
230. Coating, (11)4 (1988).
231. L. Becher, D. Lorenz, H.L. Lowd, A.S. Wood and N. D. Wyman; in
Handbook of Water Soluble Gums and Resins (R.L. Davidson, Ed.), McGraw-
Hill, New York, 1980, Chapter 21.
232. Coating, (5) 122 (1974).
233. Viscalex, Polyelectrolyte Thickening Agents, TPD/60/10, Allied Colloids.
234. I.B. Portnaya, N.O. Kazacinskaya, O.A. Stenina and R.M. Panic, Kaut schuk
Rezina, (6) 25 (1986).
235. Adhäsion, (9) 266 (1965).
236. Coating, (2) 120 (1970).
237. P.H. Gamlen and R.M. Lane, ICI, Research and Technology, p. 9.
238. W. Heilen, H.F. Fink, O. Klocker and G. Keener, Coating, (9) 338 (1987).
239. P. Stenius, J. Kuortti and B. Kronberg, Tappi J., (5) 56 (1984).
240. R. Holinger, Chemie-Anlagen & Verfahren, (11) 19 (1983).
241. Coating, (2) 39 (1970).
242. Mowilith, Farbwerke Hoechst A.G., Frankfurt, 5 Aufl., 1970, p. 131.
243. Adhäsion, (10) 307 (1968).
244. Adhes. Age, (10) 24 (1977).
245. Coating, (6) 188 (1969).
Manufacture of PSAs 557

246. H. Hadery, Coating, (2) 17 (1970).


247. Coating, (12) 390 (1969).
248. Development and Manufacture of Pressure-Sensitive Products, Marcel Dekker,
New York, 1999, p.365.
249. Coating, (5) 122 (1974).
250. Kunstharze Hoechst, ‘‘Verpackung u. Lagerung van Kunststoff Dispersionen
und Kunstharzen,’’ Hoechst, Aug. 1982, GSA 0010.
251. Kunstharze Hoechst, ‘‘Verpackung u. Lagerung van Kunststoff Dispersionen
und Kunstharzen,’’ TR 1, Hoechst, Aug. 1982, GSA 008.
252. Emulsion Polymerization of Acrylic Monomers, Rohm and Haas, CM-104A/cf.
253. R. Lombardi, Paper, Film and Foil Conv., (3) 74 (1988).
254. R.D. Bafford and G.R. Faircloth, Adhes. Age, (12) 24 (1987).
255. R. Hinterwaldner, Coating, (5) 172 (1987).
256. K. Pathamanthan, J.J. Cavaille and G.A. Yohari, Polymer, 29, 311 (1988).
257. D.G. Pierson and J.J. Wilczynski, Adhes. Age, (8) 52 (1980).
258. Adhes. Age, (9) 22 (1986).
259. D.J. Zimmer and J. Smurphy, Tappi J., (12) 123 (1989).
260. Adhes. Age, (10) 26 (1977).
9
Manufacture of
Pressure-Sensitive Labels

The pressure-sensitive label is almost always manufactured and used as a


two-component laminate (i.e., pressure-sensitive adhesives are coated onto
face stock materials, the adhesive-coated face stock material will be covered
with a release liner); therefore manufacturing pressure-sensitive labels
includes a coating and laminating operation. In some special cases (electron
beam- or UV light-induced polymerization) PSAs are actually manufactured
during the coating (i.e., the polymerization of the special raw materials,
monomers or oligomers, is carried out after or during the coating).
A laminate is a composite structure that is obtained by assembling two
or more materials with the help of a bonding agent. Different lamination
processes include wet lamination or dry lamination. Lamination with PSAs
is a dry (adhesive) lamination process where the adhesive-coated web is
dried before meeting the other laminate component (i.e., the release liner).
Labeling also may be considered as laminating. In this case, ‘‘cold’’ PSA
lamination, as its name implies, involves the use of PSA-coated paper (or
film) with treated removable paper or film liner. The system includes a
laminator which unwinds the film, presses it into the substrate, and rewinds
the liner. In order to obtain a stable laminate there is a need for a bonding
agent (i.e., to laminate it is necessary to coat the web with adhesive).
In the present context, laminating can be taken to mean the production of
a layered material from two or more, ready made webs, using a bonding
medium. Laminating is always the second step of the manufacturing
process. In the first stage the bonding agents of the laminate (i.e., the
adhesive) are coated on one of the laminate components (i.e., on the face
stock or on the release liner).
The manufacture of various pressure-sensitive products was described
in detail in [1].

559
560 Chapter 9

1 COATING TECHNOLOGY

As discussed in [2] labels are the most sophisticated pressure-sensitive


products. In comparison with other PSPs such as tapes or protective films,
which are manufactured and applied as a continuous web, labels constitute
the temporary component of a laminated web and are applied as discrete,
discontinuous items. They have to carry information. Such requirements
impose the use of complex equipment for the manufacture of labels.
Generally the coating machine for labels performs the simultaneous coating
of various adhesive and nonadhesive components.
Pressure-sensitive adhesives are coated either directly or by transfer.
When direct coating, PSAs are coated directly on the final web (face stock)
material. When transfer coating, PSAs are deposited on a web, dried and/or
crosslinked, and finally transferred onto the face stock material. This
method is necessary in order to avoid destruction of or damage to the
face stock material during drying or to avoid the formation of air bubbles.
The choice between direct or transfer coating influences the penetration and
anchorage of the adhesive as well as the release force [3]. Figure 9.1 displays
a PSA-coating machine which is able to coat either directly or by the
transfer method (i.e., via a pre-siliconized paper web) [4].
In the case of transfer coating it is very important to adjust the
adhesive characteristics with respect to the silicone so that a perfect wet-out
will be obtained. The use of a precoated (siliconized) web in transfer coating
of PSAs led to the development of tandem systems where in-line coating

Figure 9.1 Coating machine for direct and transfer coating. A) Coating station; B)
secondary web; C) laminating; 1) unwinder; 2) automatic tension control of the web;
3) press roll; 4) coating device; 5) drying tunnel; 6) rewinder.
Manufacture of Pressure-Sensitive Labels 561

Figure 9.2 Classical tandem machine for the production of label material.
Siliconizing and PSA coating. 1) Coating device; 2) unwinder; 3) rewinder; 4) drying
channel; 5) turnbars.

with PSAs is carried out. Figure 9.2 represents a classical tandem machine
for the production of label stock.
The siliconized substrate is produced in the first part; subsequently, the
adhesive coating takes place. In the second laminating section, the adhesive
layer is finally transferred onto the paper or film carrier material. In some
cases, especially when removable adhesive-coated substrates are produced,
first the application of a primer onto the face stock will be necessary (see
Chap. 2 and 6); Fig. 9.3 shows such a combination.
In some cases the silicone-coating machine has to be equipped with
an additional flexoprinter for the printing of a logo on the back side of the
web [5]. Thus the coating/laminating machine is (or may be) a complex
laminating/printing device carrying out consecutive coating processes
involving an adhesive, an adhesive primer, and printing ink (Fig. 9.4). In
an ideal case, this process occurs on an in-line machine. This machine
integrates all necessary converting processes to produce label stock, as for
example:
back side printing;
silicone coating with curing;
self-adhesive coating with drying tunnel;
rehumidification of the siliconized and adhesive-coated label
material;
primer coating with dryers;
laminating;
cutting system.
562 Chapter 9

Figure 9.3 Coating machine for primer and adhesive coating. 1) Automatic
unwinder; 2) coating device for primer; 3) drying tunnel; 4) coating device for PSA;
5) drying tunnel; 6) unwinder; 7) rewinder.

Figure 9.4 In-line coating/laminating plant. 1) Automatic unwinder; 2) printing


unit; 3) drying tunnel; 4) primer coating; 5) drying tunnel; 6) cooling and
humidifying; 7) PSA coating device; 8) drying tunnel; 9) cooling, humidifying;
10) unwinder; 11) laminating unit; 12) rewinder.

A high performance coating and laminating machine is controlled by a


process control system [5]. The individual process parameters include drying
temperatures, air velocities in the dryers, chilling roll temperatures, web
tensions, roll pressures for press rolls, applicator roll speeds, and steam feed
of jet steamers. They are controlled by independently working analog
regulators which also can be adjusted manually [5,6].
Manufacture of Pressure-Sensitive Labels 563

According to this scheme the initial coating step is followed by the


laminating step. The coating technology may be an adhesive or nonadhesive
one. The adhesive formulations may be applied to the liner by any one of a
variety of conventional coating techniques (e.g., roll coating, spray coating,
curtain coating, etc.) by employing suitable conventional coating devices
designed for such coating methods.
In the previous chapter an overview of different adhesive manufacture
technologies for PSAs (i.e., a comparison of performance, coating proper-
ties, and applications) was presented. Today a wide range of adhesive
technologies is used in the production of PSAs. The first PSAs were based
on rubber-resin solutions, but newer developments have gained a significant
part of the market and have paved the way to new applications and end-
uses, such as the development of acrylic PSAs in solution and later in water.
In addition acrylic PSAs provide excellent aging and weathering resistance;
they are used as a pure polymer (mostly in solvent-based acrylic PSAs) or
with the addition of tackifiers (many water-based acrylic PSAs). Environ-
mental problems and energy considerations are the main reasons for the
worldwide boom in the use of water-based systems for very different
applications. Both adhesive systems (solvent-based and water-based) require
the evaporation of the carrier liquid (solvent or water) in a drying oven.
Hot-melt PSAs also have been developed for pressure-sensitive uses. They
are based on different thermoplastic elastomers with addition of tackifiers
and other ingredients, and provide ease of coating because of the lack of a
carrier liquid (i.e., no additional drying is needed).
UV light-cured acrylic macromer-based PSAs are 100% solids also,
displaying the same advantage. On the other hand, all these adhesive
systems depend on the adhesive state, require different coating viscosities,
and show a dependence of the viscosity on temperature and shear (i.e., they
require different coating geometries/machines). The ‘‘TA Luft’’ regulation
in Germany and similar legislation in other countries forces producers to
review emission levels of solvents and to take appropriate measures to limit
emission levels. Therefore, the drying systems for solvent-based or water-
based systems differ; on the other hand, similarities exist between coating
technologies for adhesive and nonadhesive systems.

2 COATING MACHINES

A short review of the coating machinery is presented next. Technical


developments in this area continue at a rapid pace. As early as 1971, 2.40 m
wide coating machines were being built [7]; after two decades, the running
speed of the new machines was already double that of the older machines.
564 Chapter 9

On the other hand, as a function of the nature of the adhesive medium and
of the desired coating weight (which has to be reduced) quite different
coating systems (devices) have to be used. For example, a primer with
a coating weight of 0.5 g/m2 may be deposited via gravure printing [8].
Dispersions with 50% solids, may be coated by the use of reverse gravure
coating, down to 4–6 g/m2. Migration may produce coating defects; they
may be avoided using high coating speeds and thickeners [9,10].
Different coating systems were designed for different kinds of PSAs.
For solvent-based PSAs reverse roll coating, and for water-based adhesives
an air knife system were proposed [11]. Tape coating requires a one- or
two-station machine whereas label stock coating lines may need up to four
coating stations. After coating the material is wound into jumbo rolls for
subsequent finishing on an off-line slitter-rewinder. Label stock is manufac-
tured for further converting and/or printing operations. A release paper
protects the adhesive and separates the layers or sheets. All label stock and
double-sided pressure-sensitive tapes use a release sheet (liner). Label stock
manufactured on a one-station coater might require two or three passes to
complete the manufacturing process. The first pass applies the primer, if
required; in the second pass the adhesive is applied with subsequent lami-
nation to the release liner. A separate pass is needed to apply the silicone
coating on the release paper. A two- or three-station line could perform all
of these steps in a single operation.
The number of components of a coating machine may be reduced
and, depending on the nature of the liquid laminate component to be
coated, the differences arise from the application of an adhesive, release,
or primer coating.

2.1 Adhesive Coating Machines


The main components for the three different technologies are described
next. A solvent-based PSA-coating machine includes the following
main parts:
automatic unwinder;
automatic web guide;
automatic dual infeed tension nip for prestretch and stabilization
of the web;
primer coating unit;
release coating unit;
automatic tension nip;
self-adhesive coating unit;
drying tunnel;
Manufacture of Pressure-Sensitive Labels 565

chill rolls;
rewinder.
A hot-melt PSA machine has the following sections:
unwinder;
primer unit;
tension nip;
slot-die or other coating head geometry;
chill roll;
rewinder.
A coating machine for water-based adhesives has the following parts [12]:
unwinding cylinders;
coating device (smoothing bar);
drying oven (IR heaters);
winding cylinders;
humidifier;
laminating station;
unwinding cylinders (secondary unwinder);
rewinding.
A common web width is 1500 mm; the coating weight can vary between 10
and 250 g/m2.
There are common parts used for coating/laminating machines for
different kinds of PSAs. Moreover, some parts like winders [13–17] for the
laminating station are well known from the general practice of wet or dry
adhesive lamination. On the other hand, remoisturization of the paper web
also is necessary in paper laminating or printing [18]. Web control techniques
for PSA coating are well known from the practice of laminating or extrusion
[19–24]. The most important part of the coating machine remains the coating
head. Its choice is influenced by the rheology of the adhesive (and vice versa),
so the formulation must be compatible with the coating device.

2.2 Coating Devices/Coating Systems


From a theoretical point of view, a coating device system is based on a knife
(or blade), a rotating cylinder, or a slot-die. Possible geometries of the coating
head are shown in Fig. 9.5. The following coating systems are possible [25]:
direct transfer—metering by knife over roll;
direct transfer—metering by slow, rotating roll knife (roll blade
coater);
kiss coating—premetering by knife or reverse roll;
566 Chapter 9

Figure 9.5 Coating device. Possible constructions: A) direct transfer, knife over
roll; B) direct transfer, knife over web on the roll; C) direct transfer, knife over the
web; D) kiss coating; E) direct transfer, knife on the web and air knife; F) direct
transfer with slot-die; G) multiroll coating in direct sense; H) reverse roll coating.

multiroll coating (in the direct sense)—premetering in direct or


reverse sense;
reverse roll coating—premetering by knife over roll or reverse roll.
A slot-die coating for dispersions also remains possible. This is used in
the case of air sensitive adhesives, or when strip coating (ungummed areas)
is required [12]. Reverse roll coating allows varying coating weights (i.e., an
easy adjustment of the coating weight). Unfortunately, there is a lower limit
of the coating weight for this procedure [26]. An air knife also can be used;
this older system has lost its importance over the past several years. In this
case a one-cylinder coating device picks up the adhesive and coats it onto
the web; the premetering occurs via another cylinder, but the final metering
occurs via the air knife [27]. There are many papers describing coating
devices [13,28–34]. Actually, the most common ones are cartridge style
coaters [33]. The most important coating systems are described next.

Meyer Bar
The main components include a pan, a chrome-plated cylinder with a
two-direction drive, and a Meyer bar that is adjustable both vertically and
Manufacture of Pressure-Sensitive Labels 567

horizontally, and is driven by a stepless speed charge gear motor. The


amount of coated adhesive is controlled by a wire wound rod.

Direct Gravure
The main components include a pan, an engraved roll, a doctor blade,
and an impression roll. The engraved roll is driven by a direct current
motor.

Direct Rotogravure/Fountainless
The coating is applied with an integral doctor blade and applicator system.
This system is especially designed for higher speed coating processes.

Offset Gravure
The main components comprise a pan, an offset roller, a back-up roll, and
a doctor blade. The engraved and offset rolls are equipped with a direct
current drive. The coating weight is controlled by the depth of the engraving
and by the roll speed.

Multiroll Coating Heads


The advantage of multiroll coating heads is that there is no wear out of
blades or engraved cylinders; all rolls are driven. The coating weight is
controlled by the nip pressure, rubber hardness, and roll speed. The
properties of the PSAs to be coated are a function of the coater; parameters
such as running speed, viscosity, antifoam requirements, shear stress, reflux
requirements, wet-out difficulty, and the general application influence the
usability of a coating device for a certain adhesive [35] (Fig. 9.6). The self-
adhesive coating unit differs as a function of the adhesive technology
(solvent-based, water-based, hot-melt).

Knife and Blade Coaters


Various types of doctor blade systems and roller systems for coating onto
flexible substrates such as textiles, synthetic leather, etc., and their systems
and technological uses were described by Patermann [36]. The schematic
construction of a metering device with knife over roll (static or rotating)
has been reviewed [37].
Knife and blade coaters obtain their nomenclature from their
application systems. These machines are employed to deposit thick coating
layers with at least a 50% solids content. A scraper is used to perform the final
metering, coating, and smoothing operations, since both methods are a form
568 Chapter 9

Figure 9.6 Parameters influencing the usability of the coating device.

of the excess coating system. Knife coaters employ a rigid blade that has sharp
or rounded edges, while the blade coater utilizes a thin spring-like doctor
blade. Both machines are similar but they have different coating reservoirs,
tension, and drying and wiping systems. They also are called applicator rods,
Meyer bars, equalizer bars, or coating rods. The web passes over the metering
rod, which may be stationary or which may be rotating slowly; the direction
of the rotation may be the same as the web or the opposite. In contrast to the
scraper (blade, knife) systems, roll application (rotating cylinder) or slot-die
coating devices do not use an excess of adhesive. In the case of rotating
cylinders the fine (gravure) surface structure of the roll regulates the adhesive
coating weight. In the case of a slot-die the opening of the slot-die regulates
the coating weight. Subsequent development of both systems resulted in more
complex metering devices. The use of many frictional cylinders or a knife
on the cylinder (or round knife, roll) allows a fine adjustment of the coating
weight. In a similar manner, the combination of a slot-die with a pump (gear-
in-the-die) assures a finer coating weight regulation. The coating of PSAs
onto a web may be carried out with different coating techniques for different
adhesives. For solvent-based systems a Meyer bar or reverse roll system is
proposed; water-based systems tend to use a Meyer bar, reverse roll coater,
roll bar, or air brush. Hot-melt PSAs are coated via roll systems or slot-die
systems, eventually with the aid of an extruder.
Manufacture of Pressure-Sensitive Labels 569

Generally, one should differentiate the knife/blade coaters according


to the location of the knife. Theoretically, it is possible to wipe off the excess
of the adhesive with a knife that is in contact with the web driving cylinder
(Fig. 9.7) or with the web (Fig. 9.8).
Knife Over Roll. This is a coating device with a knife on the cylinder
(doctor blade); this method allows the use of adhesives with viscosities up to
40,000 mPa s; commonly used viscosities average 5000–15,000 mPa s [38]. In
this case the web runs with the same speed as the driving cylinder and the
stationary knife wipes down the excess of PSAs from the top of the web. The
quality of the knife (shape, material, etc.) determines the coating weight and
appearance (Fig. 9.9). Other requirements for good quality include the
absence of blade damage and dried adhesive residues, a high viscosity (lower
limit above 2000 mPa s), and a high lower limit of the coating weight. The
method is sensitive towards web thickness and cylinder tolerances.
Generally, speeds of 220 m/min are achieved [39,40]. The following types
of knife are possible:
floating knife;
roller knife;
rubber blanketed knife.

Figure 9.7 Coating device with knife on the roll. 1) PSA; 2) knife; 3) cylinder;
4) web.
570 Chapter 9

Figure 9.8 Coating device with knife on the web. 1) PSA; 2) knife; 3) cylinder;
4) web.

Figure 9.9 Coating device with knife over roll. Parameters influencing the coating
weight and coating quality.
Manufacture of Pressure-Sensitive Labels 571

The thickness of the coated film is generally higher than the opening of
the blade, and depends on the pressure before and after the blade [41]. In
production the web travels through a wetting station, then to the metering
rod where a metered thickness of the coating is allowed to pass between the
wires, and the excess of liquid retrieved into the holding tank. The coated
liquid may be applied at the wetting station by several different methods.
The web can be immersed directly into a tank or an applicator can be
rotated in the reservoir to transfer the liquid to the web at the top of its
rotation. It is important to apply an excess of coating liquid at this station in
order for the metering rod to do its job. The types of knifes (blades) known
[42] are either smooth, wire wound, or machined. There are different
constructions for holding the knife, namely:
clip type holder;
magnetic holder;
champion type;
MylarTM insert holder;
heated rod holder;
water lubricated rod holder.
The blade or knife may be used as a polishing bar, inserted knife, knife
over roll, metering rod, etc. The web passes over the metering rod, which
may be stationary, or may be rotated slowly. A smoothing rod is a finely
polished metering rod without wire and is used in several different ways.
It can be installed in the web path after coating with a gravure cylinder to
eliminate the etched pattern and enhance the coated surface. It also can
be used after a wire wound metering rod in those cases where the viscosity
of the coating liquid prevents it from flattening out because of low surface
tension [43]. In applications where a large amount of coating liquid must
be metered off the web, two wire wound rods are used in tandem, spaced
together or apart. The first rod has a larger size wire for doctoring off most
of the excess liquid, and the second, smaller wire rod, controls the finished
coating thickness. A typical coating station may use one or two metering
rods, and also have a position for a smoothing rod.

Rod Coating (Meyer Bar), Wire Wound Rod Coating


Rod coating or Meyer rod coating has been in existence since the early
1900s. These machines used equalizer bars or doctor rods; the rods are made
of steel wound with different sizes of wire. In the past several years, rods
with threads instead of being wound with wire also have been developed.
The coater applies an excess of coating to the web with an applicator
roll. The applicator roll usually is supplied with edge wipers, which wipe the
572 Chapter 9

coating off the roll at the edges. The applicator roll is usually driven by a
variable speed drive which follows line speed at an adjustable ratio. The web
passes over the applicator roll where it picks up an excess of 3–10 times the
desired final coating weight; it then passes over a wire wound rod whose
wire size determines the final coating weight. The coating weight is
determined by the diameter of the wire and by the speed difference between
the web and the rod. Metering rods may be smooth, wire wound, or
machined. The new ISO box-rod coaters have an extremely wide coating
weight range, ease of operation, and precise coating weight regulation. For
classical wire wound rods the maximum viscosity is 400 mPa s (6 mm rod
diameter, 0.075–0.75 mm wire) [44].
The thickness of the coating is governed by cross-sectional areas of the
gravures between the wire coils of the rod. The geometry of this system
creates a wet film thickness which is directly proportional to the diameter
of the wire used. Wet film coating thicknesses accurate to 2–3 m can be
achieved easily using metering rods. Mathematically, the average thickness
of the area between the wires (i.e., the coating thickness) is 0.1073 times
the wire diameter. In production coating there are other physical factors
influencing the coating weight. The most important of these is the
phenomenon of shearing the liquid. In fact, not all of the coated material
passes through the gaps; some liquid adheres to the surface of the wire. The
impact of this is usually small, but can be significant when small wires
are used or when viscosities are high. The web speed, the web tension, and
adhesive penetration in the face material will also affect the coating
thickness. The sum of all these variations seldom implies a 20% difference
from the theoretical coating weight.
The coating weight depends on the diameter of the wire as follows;
for a wire diameter of 0.3, 0.5, 0.7, and 0.9 mm the coating weights are 10,
14, 20, and 30 g/m2, respectively [45]. The maximum coating weight which
can be achieved with this system amounts to 50 g/m2. In order to obtain a
constant coating weight, this system needs a lower dispersion viscosity, a
low foaming level, a speed as high as 150 m/min, and a regulated web
tension [46]. With a Meyer bar, the wire wound bar regulates the adhesive
layer thickness on the web. With wire diameters ranging from 0.1 to 0.9 mm,
coating weights of 15–100 g/m2 are possible. Adhesives with viscosities of
2000–25,000 mPa s may be coated with a 1 g/m2 tolerance. In order to
determine the choice of an adequate metering rod, tables or diagrams
describing the dependence of the coating weight on the rod size are used
(Fig. 9.10).
The knife over roll coater may generate mechanical stress and foam;
however, a price is paid not only in line speed but also in the difficulty to
achieve a good reflux [47]. When an applicator roller is used in a rod coating
Manufacture of Pressure-Sensitive Labels 573

Figure 9.10 Rod coating. Dependence of the coating weight on the rod size. 1) and
2) are different PSAs.

system, the speed of the applicator is not a critical factor. Other systems
using gravure or knife coating methods require precise roller speed
adjustments. In a rod coating system the machine operator can adjust the
applicator roller speed within a wider range and can do it while the machine
is running. In rod coating the actual thickness of the coating can be affected
by web speed, viscosity, and other factors. A gravure cylinder, although
expensive and limited in its range, provides an extremely accurate coating
(appearance, weight) almost independently from the web speed. Depending
on the application, rod coating speeds are usually limited to 300 m/min,
although some coaters claim web speeds up to 600 m/min. The critical factor
in the web speed of a rod coater is the time for the striations formed by
the rod to flatten out and become smooth; the web speed must be controlled
to allow time for flattening before the web is dried. The design of a rod
coating station should ensure that the web makes intimate contact with
the wires of the metering rod. ‘‘The wrap angle, ’’ the angle of the web
direction as it approaches the rod and its direction as it leaves the rod,
should be 15 for a heavy web tension, or up to 25 for a light web tension.
Web tension is a critical factor in the design of a rod coating station. With
574 Chapter 9

a wrap angle of 15–25 , the web must be tight enough to ensure intimate
contact with the metering rod, yet not so tight that the web is deformed by
the wires.
Adhesives can solidify between the wire windings of the rod whenever
the coater is stopped. Many coating machines have a ‘‘throw-off ’’ feature,
a mechanical means of separating the web from the rod automatically
whenever the machine is turned off. When a coating liquid is allowed
to extend to the edges of the web, it can spill over to the dry side and
contaminate the idle roller. There are several methods used to maintain a
dry edge, according to the coating process used. In gravure coating, the most
common method is to undercoat the backing roll at the edges; disadvantages
include the need for a wide assortment of back-up rollers. In roll coating
the spillover is usually controlled by specially constructed dams which are
placed in direct contact with the coating roll. In a rod coating system the dry
edge is controlled by wipers or deckle straps on each edge of the applicator
roll [47]. Because the wipers are easily moved their position can be adjusted
while the coater is operating, with no downtime at all. The next step in
the development of knife/blade scrapers is the use of ‘‘round blades’’ or
stationary rolls (Fig. 9.11).
In this case the blade has a ‘‘break-down’’ edge; a direct mass transfer
of the PSAs on the web occurs, as well as metering by the aid of a stationary
round blade. In a subsequent development the stationary blade is replaced
with a rotating cylinder; in this case one would have a metering roll (roller
coater).
The direction of rotation may be the same or the opposite to that of
the web. The most common procedure is to rotate the rod slowly in the
opposite direction to the movement of the web. The rotation works twofold:
it flushes the coating material between the wires, keeping the wire surfaces
wet, and it prevents setting up and hardening of some liquids. The rotation
will also distribute any abrasive wear evenly over the wires.

Figure 9.11 Coating device with stationary roll. 1) PSA; 2) stationary roll;
3) cylinder; 4) web.
Manufacture of Pressure-Sensitive Labels 575

Figure 9.12 Roll coating systems with direct/indirect mass transfer onto the web.
A) Direct gravure; B) direct gravure with reverse roll; C) offset gravure; D) kiss
coater; E) slot-die, duplex coater.

There are different roll coating systems for solvent-based or water-


based PSAs, with low or high viscosities; they differ according to the direct
or indirect character of the adhesive mass transfer onto the web (Fig. 9.12).

Roll Coating Devices/Rotogravure


In a manner different from the scraper (blade, knife) systems, roll
application (rotating cylinder) or slot-die coating devices do not use an
excess of adhesive. In the case of rotating cylinders the fine (gravure) surface
structure of the roll regulates the adhesive coating weight. In the case of
a slot-die the opening of the slot-die determines the coating weight.
Subsequent developments of both systems led to more complex metering
devices. The use of several frictional cylinders, or a knife on the cylinder (or
round knife on a roll), allows a fine adjustment of the coating weight. In a
similar manner the combination of the slot-die with a pump (gear with die)
ensures a finer regulation of coating weight. In order to define the choice
of an adequate coating device, classification charts were developed [48].
Figure 9.13 summarizes the most commonly used coating geometries.
576 Chapter 9

Figure 9.13 The main coating methods used for roll coating. A) Direct gravure;
B) reverse gravure; C) direct offset gravure; D) and E) three-roll reverse pan fed
gravure; F) three-roll reverse nip fed gravure.

Theoretically, it is possible to use a direct gravure when direct contact


between the engraved cylinder and the web occurs, and indirect gravure
when the adhesive is transferred from the engraved cylinder to the web via
intermediate cylinders. In principle, the engraved cylinder depth (gravure)
may be zero (i.e., a smooth cylinder also can be used). Moreover, according
to the rotation of the engraved cylinder one can distinguish direct or reverse
roll coaters. In many cases reverse gravure is combined with reverse roll
coating. A reverse rotation of a Meyer bar also is used in order to regulate
the coating weight. Indirect gravure or offset gravure coating employs one
or more transfer rollers between the gravure cylinder and the substrate.
Indirect coating is offset from the gravure cylinder to the transfer roller and
then applied to the web. A coating is offset when the adhesive is transferred
to a dry surface and then applied to another dry surface. The intermediate
rollers improve the wet coating characteristics, reduce or remove the
adhesive pattern, and enable the unit to apply a uniform and thinner coating
film. There are two basic offset gravure configurations (i.e., with vertical and
angular orientations) (Fig. 9.14).
An offset gravure coater is an extension of the direct gravure system,
where the rubber pressure roll becomes a driven offset on the transfer roll,
and a third roll (steel) is added. In operation the coating is first transferred
Manufacture of Pressure-Sensitive Labels 577

Figure 9.14 Roll coating. The basic offset gravure configurations. A) Vertical,
two-roll offset gravure, direct; B) vertical, two-roll offset gravure, reverse; C) angular
three-roll offset gravure.

to the intermediate offset roll before being applied to the web. The
advantages of offset gravure include the following:
The process is suitable for running with rough web surfaces.
The engraving pattern of coatings with poor flow is eliminated.
A more complete removal of the coating from the cells is possible
(suction effect of rubber).
A virtually self-cleaning system can be obtained; it is recommended for
100% solids. More shearing of the coating is obtained and pattern coating
is possible. A variation of this method can be used for two-sided coating.
The essential distinction between differential offset gravure and normal
offset gravure is that coating weight variations may be obtained by
independently altering the speeds of the compression roll as well as the offset
roll in their relation to the speed of the coating roll or the web itself.
The vertical offset gravure configuration is best suited for liquids with
slow to medium evaporation rates. The transferred wet coating remains on
the rubber-covered offset roller through an angle of rotation of 180 before
it is deposited. The angle of rotation through which the coating remains
on the wetted cylinder before being transferred or deposited is normally
referred to as the wet angle. The angular offset gravure configuration
578 Chapter 9

(Fig. 9.14) reduces the wet angle by almost one half. The angular design can
be used with faster evaporating diluents and when more than two transfers
are required. The coating is applied to the rubber-covered transfer roller by
the gravure process, after which the coating pattern has a tendency to flow
together, since the rubber covering repels the liquid. The coating then is
transferred by the film splitting process to the chromed steel application
roller, which in turn deposits the coating onto the web.
Two-roll gravure coaters may have a reverse direction of the engraved
roll. In reverse gravure the engraved cylinder rotates in the opposite
direction to the web being coated. In this case the wiping action of the
opposite rotation of the etched cylinder tends to obscure the cell pattern
that sometimes remains on the surface of the applied coating. This is
advantageous for coatings with poor flow-out properties. Because the web is
not nipped against the etched cylinder, as in direct gravure, the speed of the
etched cylinder can be varied; this offers the ability to vary the coating
weight. The doctor blade can be relocated to the opposite side of the coater,
which is not easy in direct gravure coaters. An alternative procedure would
be to keep the blade mechanism in its original location and adapt a flooded
blade technique. In reverse gravure it is not possible for the etched cylinder
and the impression cylinder to be pressed tightly together to form a tension
contact nip. Here, the tensile strength and elasticity of the web are important
factors for predicting the need for a driven back-up roll. A negative factor
on reverse gravure is the increase in the rate of gravure cylinder wear.
Another design of reverse gravure is able to vary the speed of the etched
cylinder. The relation between the speed of the etched roll and the speed
of the web running in the opposite direction has a decisive influence on the
overall quality of the coating. The etched roll speed must be at least as high
as that of the web or even higher. The indirect reverse gravure may
be used for webs having quite different thicknesses [49]. In gravure coating
the most common method is to undercut the backing roll at the edges. The
disadvantage is the need of an assortment of back-up rollers, one for each
width of web the user handles. Changing the backing roller for each width
also requires more time.

Coating Weight
Generally, when coating with cylinders, the adhesive is taken from an
adhesive pan (kiss coating). The coating weight depends on the pressure
of the web on the cylinder, the viscosity of the fluid, web speed, and cylinder
speed [46]; coating weights of 2–50 g/m2 are possible. Viscosities of 20 DIN
sec to 12,000 mPa s are used, with coating weight tolerances of 1 g/m2.
Generally, the web width is limited to 1600 mm; the system is open and tends
Manufacture of Pressure-Sensitive Labels 579

to dry. Continuous pumping (recirculation) of the material content is


necessary. Generally, for gravure cylinders, the coating weight depends on
the geometry of the gravure. The factors influencing the ink hold-out for
printing also influence the coating weight of the adhesive. These factors
depend on the engraved cylinder rolls, on the machine parameters, and on the
adhesive. Figure 9.15 summarizes the parameters influencing the coating
weight (e.g., line screen number, angle, depth, velocity, viscosity).
One of the most important factors influencing the hold-out is the line
screen number (Fig. 9.16). The adhesive hold-out decreases with increasing
line number. It also depends on the depth and angle of the engraving
(Fig. 9.17). The adhesive hold-out also depends on the running speed
(Fig. 9.18) and increases with increasing viscosity (Fig. 9.19).
There is a (theoretical) choice between gravure depth and angle
(Fig. 9.20). The relative importance of the different parameters influencing
the adhesive hold-out may be combined into a formula:

H ¼ f ½ð9C1:3V=1:1SÞðeNÞ ð9:1Þ

where H is the hold-out, C is the coating cylinder, V is the viscosity, S is the


speed, N is the nature of the adhesive, and e is a coefficient that is dependent
on V; the measured coefficients (9, 1.3, 1.1, or e) may differ according to the
experimental conditions. As can be seen from the above equation, the most
important parameter influencing the coating weight is the cylinder. The

Figure 9.15 Machine and running parameters influencing the coating weight.
Gravure coating.
580 Chapter 9

Figure 9.16 Gravure coating. The influence of the line number on the adhesive
hold-out. Angle 80 .

Figure 9.17 Gravure coating. Dependence of the adhesive hold-out on the depth
and angle of the engraving. Line number 70; 1) 140 angle; 2) 130 angle; 3) 120 angle.
Manufacture of Pressure-Sensitive Labels 581

Figure 9.18 Gravure coating. Dependence of the adhesive hold-out on the running
speed. 1) Line number 70; 2) line number 80. Cell depth 37 m.

Figure 9.19 Gravure coating. Dependence of the adhesive hold-out on the


viscosity. 1) Cell depth 27 m; 2) cell depth 30 m.
582 Chapter 9

Figure 9.20 Gravure coating. The influence of the gravure cell depth/angle on the
coating weight. The choice of these parameters for the same coating weight; 1) line
number 80; 2) line number 70.

influence of the viscosity also remains important; the influence of the speed
is relatively small:

CV >S ð9:2Þ

The nature of the adhesive may cause changes of 200% in the hold-out
(coating weight). Concerning the cylinder characteristics, the most
important parameter remains the depth of the gravure (up to 500%
change). The line screen has a decisive influence too (about 50%). The hold-
out also increases with increasing engraving angle (Fig. 9.17). Figure 9.18
illustrates the dependence of the adhesive hold-out on the running speed
for various gravure cylinder characteristics; in Fig. 9.19 the effect of the
viscosity on the adhesive hold-out is presented for different cell depths. The
coating weight for gravure cylinders depends on the gravure number, form,
depth, and cavity ratio [50–52]. Figure 9.20 illustrates the influence on the
gravure cell depth/angle on the coating weight for different line numbers.
Gravure cylinders are used for coating adhesives with viscosities of
100–3000 mPa s, and an ensuing coating weight of 10–50 g/m2 [51]. Different
line screen numbers are recorded for different coating weights; the gravure
Manufacture of Pressure-Sensitive Labels 583

Figure 9.21 Roll coating, gravure coating, coating weight. Dependence on the
cylinder geometry. 1) Solid content 45%; 2) solid content 75%.

characteristics (quadra gravure, free-flow, etc.) depend on the adhesive


characteristics [53–55]. For example, a 40-line screen is recommended with a
70-m depth for a coating weight of 5–7 g/m2 [56]. The choice of a cylinder
is made easier using tables or diagrams concerning the coating weight
dependence on the cylinder geometry/line number (Fig. 9.21).

Roll Coaters with Doctor Blade/Mixed Systems


Reverse gravure with a closed chamber and a doctor blade [57] in the
coating station also is used. A coating weight of 0.2–30 g/m2 is possible;
viscosities from a low number of DIN seconds up to 5000 mPa s can be used.
Changes in coating weight are obtained by changing the cylinder. The blade
and the cylinder cause shear of the adhesive [46]. Changes in web width are
made via changes in web covering or by changing the engraved cylinder.
Conventional or reverse angle blades can be installed. In the reverse
angle doctor blade arrangement the excess coating is sheared from the surface
of the engraved roll rather than being squeezed from it. The system eliminates
the need to increase blade pressure for higher viscosities and reduces the
turbulence in the coating fountain. A fountainless applicator with a coating
chamber also may be used. Knurled gravure rolls in combination with a
584 Chapter 9

doctor blade may be used to coat primers with a low viscosity at 25 C with a
thickness of less than 1 m [58]. Among the possible coating systems (e.g.,
MEC roll, reverse roll, fixed doctor blade system, etc.), the most used
one appears to be the MEC roll [25]. This system is not only a simple
configuration, but it also is easy to set up in order to obtain a uniform coating
with the desired coating weight. The system is designed to overcome
a series of problems, for example:
correct ratio between coating cylinder diameter and doctor
(metering) roll diameter;
angular position of the doctor roll;
mechanical tolerances of the order of a few microns concerning
excentricity, surface uniformity;
the doctor roll speed has to be adjustable, and to run dependent on
the coating cylinder speed by an electronic master/slave system;
adjustable compensation for the doctor roll flexure;
fine adjustment of the gap between coating cylinder and doctor roll,
independently adjustable parallelism on operator and drive side.
Second-generation doctor blade chamber technology was developed.
The first generation of doctor blade chambers, the so-called ‘‘single chamber
systems’’ do not ensure a sufficient fill in of the chamber. The coating
agitates against the cells of the roll in a circular movement, creating an
inconsistent pressure. The patented InkJector doctor blade design consists
of two chambers separated by a small, narrow ridge. The two chambers are
connected to each other by a small gap (less than 4 mm). There is a forced
flow of the liquid which causes high localized pressure [59].
There are six common designs used for gravure coating systems [60].
Gravure systems are normally constructed in two-roll or three-roll configu-
rations. Two-roll constructions may be direct, reverse, and differential reverse.
Three-roll machines may be offset, reverse offset, and differential reverse
offset. For direct gravure in a two-roll design, the roll is partially submerged
in a coating pan and as it turns it brings the coating up. The excess coating is
wiped from the surface with a doctor blade leaving an amount predetermined
by the total carrying capacity of the etched or knurled cells. Therefore for
a selected cylinder the only way to change the coating weight is to change
the solids content of the coating itself. The doctor blade is usually of the
oscillating type and the upper web back-up roll is made essentially of an
elastomer-covered sleeve. The web runs between it and the engraved roll.
The advantages of direct gravure include:
widely used and accepted;
provides a consistent and uniform coating;
Manufacture of Pressure-Sensitive Labels 585

low to medium equipment cost;


independent from operator skills.
The disadvantages of the direct gravure include:
the pattern of the engraved roll may be evident if the coating does
not have adequate flow;
the doctoring of viscous coatings at high speeds is difficult;
high foam generation at higher speeds.
There are several ways to reduce the disadvantages of direct gravure.
They include heating of the coating or using a polishing bar. The latter can
be variable speed driven, in either direction and cooled or heated.

2.3 Choice of Coating Geometry


Generally the choice of a coating device depends on the following
parameters:
desired coating weight range;
viscosity;
changes in coating weight;
uniformity of the coating weight;
abrasion of the metering devices;
changes in coating (web) width;
changes in coating medium.
There are general requirements for coating devices, such as environmental
and energy considerations, short operating time, and ease of operation [8].
New cartridge-style coaters and quick changeover rubber sleeve rolls allow
a full change of the gravure roll and supply system in 5 min or less [61].
The coating of PSAs on a web may be carried out using different
coating techniques. For solvent-based systems Meyer bar or reverse roll
coater systems are proposed; water-based systems use Meyer bar, reverse
roll coater, roll-bar, or air brush systems [4]. According to [62] on pilot
coating lines for solvent- and water-based products over 45 different coating
methods can be tested. Hot-melt PSAs may be coated via roll systems or
slot-die systems, and eventually through extrusion.

Roll Coating of Solvent-Based PSAs


The cylinders used for PSA coating generally have a 180–400 mm diameter, a
regulated drive, and are explosion proof [63]. Coating machines for solvent-
based PSAs vary as a function of the viscosity of the adhesive. For low
viscosity systems, a driving cylinder brings the PSA from the pan and coats
586 Chapter 9

it on the web (Fig. 9.22). A reverse roll regulates the coating weight (metering
roll). For highly viscous systems, two metering rolls bring the adhesive on the
web (reverse rolls) (Fig. 9.23) and each is continuously regulated.
Common roller coaters with accugravure operate with solvent-based
PSAs having a solids content of 38–40%, at speeds of 300–400 m/min. Slot-
die coating devices may be used for solvent-based PSAs also [64]; viscosities

Figure 9.22 Roll coating for a solvent-based PSA. Coating systems for low
viscosity solvent-based PSA. 1) PSA; 2) driving cylinder; 3) web; 4) rotary doctor bar.

Figure 9.23 Roll coating for a solvent-based PSA. Coating system for high
viscosity solvent-based PSAs. 1) PSA; 2) driving cylinder; 3) web; 4) metering roll.
Manufacture of Pressure-Sensitive Labels 587

up to 1–2 million mPa s are possible and coating speeds of 150 m/min were
achieved. Solvent recovery systems remain an absolute necessity.

Coating Machines for Water-Based Systems


Whereas for solvent-based and hot-melt PSAs coating the operating princi-
ples, the coating systems, etc., are widely known, the same is not generally
valid about water-based systems. Generally, reverse gravure and kiss
coating have been used since the 1970s [65]. Film forming of water-based
adhesives was described [66]; a wide range of requirements must be fulfilled
when using a PSA emulsion coater.
To keep foaming at the lowest possible level the coater must have a
closed supply system (i.e., overflow should be avoided, which precludes an
open pan feed and roll gap feed system). Since the emulsion-based PSAs
were introduced for label products the coat weight range has varied between
20 and 24 g/m2 (dry). To keep the coat weight tolerance within a maximum
of 5%, the viscosity of the emulsion must be kept at a constant level (i.e.,
the temperature must remain under control). To keep the intervals between
cleaning as long as possible (coagulation of an emulsion cannot be avoided)
the coating head must be easily accessible for quick cleaning to get rid of
streaks; good access is necessary even during operation.
The following coating devices are used for water-based PSAs [45]:
reverse roll;
gravure with direct adhesive transfer;
knife over roll;
reverse gravure.
These methods use a premetering system with the aid of a multi-
cylinder system. On the other hand, there are some methods using an excess
of adhesive, namely roll blade, air brush, and knife over web. The systems
with knife over roll or knife over web may be used for high coating weights,
and medium to high viscosities. Manufacturing speeds higher than 100 m/min
are possible; a low foam level is ensured. The shape of the knife does
not have any influence when solvent-based adhesives are used, but for
dispersion-based ones there is a need for a ‘‘break-down edge,’’ as a station-
ary metering roll is not possible. In Europe today, two coating methods for
emulsion PSAs actually dominate their fields (i.e., the metering bar systems
for tapes and the reverse gravure system for labels). The coating machines
for PSA tapes using water-based PSAs have been described [57].
Figure 9.24 shows the main components of a reverse gravure system. A
closer look at the reverse gravure system reveals why it also is considered
more and more for coating tapes. The most outstanding advantages of this
588 Chapter 9

Figure 9.24 The main components of the reverse gravure system. 1) Supply tank;
2) pump; 3) filter; 4) feed box (pan); 5) gravure roll; 6) knife; 7) backing roll; 8) web.

coating method are its ease of operation and cleaning, as well as the very
smooth surface of the adhesive. The method is not sensitive towards circular
roller tolerances [6]. Unfortunately, changing the coating weight requires the
change of the coating cylinder since it is not possible to cover a complete
coating weight range with a single gravure size. Other disadvantages include
the difficulty in doing pattern coating and in controlling the coating weight
at low speeds (below 50 m/min).
Different kinds of reverse roll coatings [38] include: the nip-fed system,
the pan-fed system, and the Dahlgren system. The nip-fed and pan-fed
systems use a metal cylinder; the Dahlgren system requires a special hydro-
philic cylinder (rubber). Reverse roll coaters were developed for adhesives
with a high viscosity. Generally, the coating device is built up from the
following parts: coating cylinder, metering cylinder, transfer cylinder, and a
falling blade. The transfer cylinder runs at 100% of the maximal speed of the
machine. The coating cylinder runs in a reverse sense to the transfer cylinder,
at a speed of 15 to þ200% of the coating cylinder [52]. Metering cylinders
run in a reverse sense to the coating cylinder, with a rotation speed difference
of 10% with respect to the coating cylinder. In order to achieve a smooth
film, the cylinder speed should be 1. 3–1. 5 times higher than the velocity of
the web (i.e., for a production speed of 200 m/min, the cylinder should run at
300 m/min) in order to avoid adhesive defects [45].
Conventional roll coaters designed for solvent-based adhesives can
be used successfully including techniques such as gravure, gravure offset,
reverse kiss, direct roll kiss, air knife, etc. Because of the basic differences
between solution adhesives and polymer dispersions there are considerable
Manufacture of Pressure-Sensitive Labels 589

differences in rheology and it is necessary to take extra care when coating


dispersions in order to optimize coating efficiency. For example, a greater
clarity and transparency of the final film structure and smoothness of
adhesive coating are ensured if a smoothing roll is installed just after the
coating head. Suggested specifications of such a smoothing roll include
a 1.5–2.0-inch diameter, a polished or plated surface, a reverse direction
drive, a speed of rotation which is 4 times the linear speed, and at a distance
from the coating head of about 20 mm; the track angle is determined
through trials [66].
With a given gravure size it is possible to adjust the coat weight within
a range of approximately 10%. This can be achieved by changing the pump
pressure, the speed of the gravure roll, and the pressure of the doctor blade,
and by the appropriate choice of the type of doctor blade (thickness of the
polyester blade and the shape of the blade tip). High speed (short contact
time between web and gravure) implies a low coating weight, whereas a low
dispersion viscosity means high coating weight. For water-based PSAs on
a reverse gravure coater, it is ideal to have a high solids content with a low
viscosity. The range of the solids or the viscosity is dictated by the surface
tension of the dispersion; production speeds of up to 300 m/min with a
coating weight tolerance of  3% are possible. The capabilities of a reverse
gravure coater for emulsion PSAs are summarized in Table 9.1.
Reverse roll coaters coat PSAs with a viscosity of 12,000–50,000 mPa s
[52]; the coating weight ranges are 0.5–800 g/m2. The reverse roll coater was
developed for viscous media; the shear forces reduce the viscosity during the
coating. The coating weight is determined by the distance between coating
and metering cylinder, and by the relative speed of the two cylinders
respective to the web. Generally, there is a speed difference between the

Table 9.1 Performance of a Reverse Gravure Coater for


Water-Based PSAs

Coater Characteristics
Coating Characteristics,
Line number Speed (m/min) Coating Weight (g/m2)

10 50 36
20 200 22
40 50 15
50 50 8
70 50 6
80 50 4
Gravure angle: 45 ; viscosity range: 800–1500 mPa s.
590 Chapter 9

coating and metering cylinder of 10%. In order to avoid a discontinuous


adhesive coating and to obtain a fine, smooth adhesive surface, there are
some special requirements the PSAs need to fulfill [38]:
The metering roll has to be fixed or to run so rapidly that no drying
of the adhesive on the roll can occur.
The diameter of the metering roll has to be small in comparison to
that of the application roll.
A low viscosity is needed.
An adhesive without thixotropy is needed (i.e., the emulsion PSAs
must be designed for the reverse gravure method).
Given these characteristics, production speeds of up to 300 m/min and
coating weight tolerances of  3% may be achieved [67]. For reverse gravure
a speed of 200–300 m/min with a coating weight of 18–25 g/m2 is possible
[45,68]. Generally, 65–160 lines/cm, an angle of 40–90 , and a cell depth of
22–85 m are used for engraved cylinders with a diagonal screen and the
calotte shape of cells is 45 ; for water-based dispersions high line numbers
and less gravure depth are required.
Having designed water-based acrylic adhesives with a desirable
rheology, wet-out behavior, low foaming, and good drying characteristics,
adhesive suppliers find that almost every converter operates a different
coater [69]. This means further formulation refinements to tailor make
water-based PSAs for each type of coater configuration (Table 9.2). The
higher speed rotogravure coaters require emulsions with higher mechanical
stability [70]. They require low viscosity dispersions when wet-out is difficult;
therefore more surfactants are added in the formulation, which in turn leads

Table 9.2 Dependence of the Formulation of Water-Based PSAs on the Coater


Configuration

Viscosity Range

Value (mPa s) Method Coater type

100–1000 Brookfield, 20 rpm Wire wound rod (Meyer bar)


1400–3400 Brookfield, 12 rpm, spindle 3 Gravure
100–1500 Brookfield, 20 rpm Dahlgren
3000–10000 Brookfield, 20 rpm Knife over roll
50–500 Brookfield, 20 rpm Reverse gravure
1000–5000 Brookfield, 20 rpm Nip fed
200–3000 Brookfield, 20 rpm Pan fed
1000-10000 Brookfield, 20 rpm Slot-die
Manufacture of Pressure-Sensitive Labels 591

to higher foaming. The reverse roll coating is suggested for direct and
transfer coating (Table 9.3) as the choice of a coater depends on the basic
formulation.

2.4 Other Coating Devices

Air Knife
It appears easier to coat an adhesive when the web is flooded (air knife)
with PSAs before doctoring occurs, rather than receiving coverage via
a theoretically precise layer transfer from a metering roll. The air knife often
allows the passage of a particle of solid matter which could cause continuous
lines when caught under the blade. Spray from the continual blast of air
onto the wet coating can be a nuisance and foaming will only be made
worse, but the introduction of predoctoring, either by roller or wire wound
rod, can reduce air knife pressures and volumes, and hence eliminate
potential problems.
In the air brush method the coating weight is regulated via air flow.
Coating weights of 7–11 g/m2 are possible at a viscosity of about 200 mPa s,
with a tolerance of 1 g/m2 [46]. Air knife systems operate with a 1500-mm
water pressure. In the case of kiss roll coating with an air brush, the latex is
coated using a smooth cylinder, and the excess of the latex is blown away
by an air brush; a thin air stream with a pressure of 6.9  10 N/mm2 should
be used.

Slot-Die Coaters
Simple slot-dies may have coating weight tolerances of  1.5 g/m [71]. The
slot-die coater and its use for dispersions have been discussed in a detailed
manner [72]. Bolton-Emerson developed a gear-in-die slot-orifice die. This
has a liquid pumping section built into the die; the pump is designed to cover
the complete die length. Advantages of this new die coating method include
higher solids, better accuracy, no foaming, and low shear [77]. The original
and patented tube extrusion die for full and strip coatings was invented by
George Park; over the years it was substantially further developed [29]. The
die evolution went from the original simple tube die to the coathanger
configuration, from there to the gear-in-die (GID), and lately to today’s
Duplex coater. The Duplex system lends itself not only to aqueous, but also
to two-component crosslinking adhesives. The shear rate in the slot-die is
lower than in roll coaters, thus dispersions which need a high shear rate to
be coated (for film forming) are not adequate for slot-die coating [28].
A coating speed of 250 m/min is achieved for hot-melt PSA coatings [8].
Table 9.3 Dependence of the Choice of the Coater on the Basic Formulation
592

Formulation Characteristics Running Characteristics End-Use Characteristics

Antifoam Reflow Coating Face Stock


Physical Viscosity require- Speed Shear require- Wet-out General weight
2
Coater Type state (mPa s) ments (m/min) stress ment difficulty use (g/m ) Paper Film Others Refs.

Knife SB, WB 100–1000 M H M L H D/T 0.5–50 x x —


Meyer rod L
Chambered HM, WB M L M M L M D/T 5–50 x x —
Doctor
blade
Knife over SB, WB >3000 M L M H L T 0.5–0.35 x x —
roll L
Direct roll SB, WB M M M M M M D/T 20–200 x x x [35,
H 73–76]
Reverse roll HM, SB, WB 1000–5000 M H M M M D/T 2–50 x x x
300–500 L L H L
Rotogravure HM, SB, WB 50–500 H H M L H D 1–40 x x x
VL H
Slot-die HM, WB 1000–10000 L M M M H D/T 1–200 x x x
L
SB, solvent-based; WB, water-based; HM, hot-melt.
D, direct coating; T, transfer coating.
VL, very low; L, low; M, moderate; H, high.
Chapter 9
Manufacture of Pressure-Sensitive Labels 593

The GID coating device allows running speeds of 8–400 m/min as a function
of the viscosity (2000–500,000 cP) [31].
Film coatings may show ribbing patterns that are generated by a poor
adhesive system. Ribbing seen on the applicator roll will result in ribs in the
film. Foaming can be caused by the recirculation of the adhesive emulsion
from the pick-up roll into the pan. The main coating defects and the
methods to avoid them are listed in Table 9.4.

Table 9.4 Coating Defects and Methods to Avoid Them

Coating Defect Cause/Solution

1. Foaming Cause: Pumping, stirring speed too high; viscosity


too low; wetting agent level too high, inadequate.
Solution: Reduce speed; increase the viscosity; add
defoamer.
2. Wetting problems Cause: Viscosity too low; surface-active agent inade-
quate, surface-active agent level too low; release
liner inadequate; treatment level (films) too low.
Solution: Increase viscosity; add adequate surface-
active agent, increase the coating speed and shear;
increase the drying speed, change the release liner;
treat the film again.
3. Bubbles in the adhesive Cause: Air speed too high; drying speed too high.
layer Solution: Reduce air blow level, reduce drying tempe-
rature in the first drying step.
4. Migration of the Cause: Viscosity too low; drying speed too low; paper
adhesive in the face porosity too high; low molecular weight water solu-
stock material ble components in the formulation; low molecular
weight tackifier/plasticizer in the formulation; sur-
factant level too high.
Solution: Increase viscosity; decrease surfactant level;
increase the molecular weight of the low molecular
weight components in the formulation; use primer
coating, make transfer coating.
5. Dry adhesive layer is Cause: Adhesive is not dried enough; surfactant level
too soft, legging too high.
Solution: Improve drying; use higher temperature and
air volume (WB).
6. Adhesive layer is not Cause: Viscosity too low/high; grit; coating device is
smooth enough, shows not clean; mechanical stability of the PSA is not
lines sufficient; too high shear during transport.
Solution: Change the rheology of the product; filter the
PSA, change the shearing conditions, change the
coating device.
(continues)
594 Chapter 9

Table 9.4 Continued


7. Release too high/low Cause: Drying is not adequate; release liner is not
adequate; release liner is too fresh/old.
Solution: Improve drying conditions; change release
liner quality and age.
8. Coating weight too Cause: Viscosity is too low/high; solid content is too
high/low high/low; coating speed is too high/low.
Solution: Change the rheology of the formulation;
change the solid content of the formulation; change
the coating speed, coating geometry.

3 COATING OF HOT-MELT PSAs

The composite laminate is typically manufactured by coating the hot-melt


PSA in a molten state at a temperature above about 130 C, from a slot-die
or roll coater onto a release liner. The coated release liner then is laminated
to the face stock with a nip roll using pressure between a rubber and a steel
roll. The coating machines and devices used for hot-melt PSAs have been
described [78–80]. Coating devices used for hot-melt PSAs are similar to
coating devices used for solventless silicones and for other 100% solids
adhesives. Evidently, other coating systems such as slot-die roll and reverse
roll coating geometries were proposed [81]. Such 100% solids materials
including UV- and EB-curable products can be coated at speeds up to
1500 m/min [62].

3.1 Roll Coaters for Hot-Melt PSAs


Coating devices using engraved cylinders move the adhesive from the nip
with the aid of a cylinder. The adhesive layer is smoothened by a knife (pres-
sure and angle regulated). The molten adhesive is transferred directly to the
web or indirectly through the use of an elastic cylinder. The coating weight
depends on the line numbers and geometry of the cylinder, and the viscosity,
adhesion, and cohesion of the adhesive; the viscosity is limited to 5000 mPa s.
A reverse roll system permits the use of 150–20,000 mPa s. In this case, a
metallic cylinder removes the adhesive from the nip. The excess adhesive is
controlled by a reverse roll metering cylinder (cleaned by a knife) [78].
The advantages of roll coating systems for hot-melt PSAs [81] are
a low sensitivity towards dirt, low tolerances in coating weight (1 g/m2 at
60–350 m/min), rapid changes of hot-melt PSA formulations, and a rapid
switch on/off and stop mode is possible. The disadvantages of the roll
Manufacture of Pressure-Sensitive Labels 595

coating systems for hot-melt PSAs can be summarized as follows [81]: a


direct contact between cylinders should be avoided, a coating weight range
of only 18–100 g/m2 is possible, a fine ridging on the adhesive layer may
be observed, and the system is only half closed and thus it is not protected
from oxidation.
It may be concluded that depending on the desired coating weight,
either the slot-die (10–16 g/m2) or the roll coating (16–100 g/m2) system
should be selected. A roller coating system with four cylinders is able to
operate up to 5000 mPa s, and coats up to 250 g/m2 [82]. Thus the main
disadvantage or limiting factor in the use of roll coaters for the hot-melt
PSAs remains the viscosity. In the kiss coat procedure the web goes over
the application cylinder (one or two cylinders) and is coated in a normal or
reverse roll sense. The coating weight is strongly influenced by the contact
web/cylinder, web tension, and rotating speed of the cylinder. No viscosity
above 900 mPa s can be successfully coated. Continuous metering of the
adhesive is possible using a smooth cylinder coating device, where the web
travels between two or more cylinders. The speed of the web, the viscosity,
temperature, knife clearance, web angle, cylinder synchron, etc., influence
the coating weight. This system may be used for hot-melt viscosities of
300–500 mPa s [78]. In a manner similar to solvent-based or water-based
adhesives, diagrams or tables are used in order to select the adequate
engraved cylinder (Fig. 9.25).
For hot-melt coatings a running speed of 250 m/min can be
achieved [83].

3.2 Slot-Die Coating for Hot-Melt PSAs


The justification for purchasing hot-melt slot-die coaters is based on
economic considerations. The use of highly concentrated coating media
and the necessity to avoid solvents has created the demand for coating heads
which can handle highly viscous materials. The roller coating methods have
steered new development into different directions. Through the combination
of reverse roll coating with the knife over roll coating, a new coating method
was born which allows the handling of a wide range of viscosities and which
is applicable to existing equipment and machinery. Above a viscosity of
30,000 mPa s it is not possible to use a metering roll; in this case a slot-die is
required. Higher viscosities are used for special products like electron beam-
curing of hot-melt PSAs. In this case viscosities up to 1,000,000 mPa s are
possible, but only in the final stage [83]. Coating systems for hot-melt PSAs
with slot-die (1300–1500 mm) are able to coat 18–40 g/m2 at a manufacturing
speed of 15–270 m/min [84]. The gear-in-die system used for solvent-based
and hot-melt PSAs may operate in a temperature range of room temperature
596 Chapter 9

Figure 9.25 Gravure coating for hot-melt PSAs. The choice of a hot-melt coating
cylinder as a function of the coating weight on the line number. Working
temperature 120 C. 1) and 2) are two different cell depths.

to 250 C, with viscosities of 200–1,500,000 mPa s. The slot-die system for


hot-melt PSAs was tested in 1964 by George Park (Park Coater) [85]. The
advantages of the die system may be summarized as follows:
high viscosities are allowed;
it is a simple coating device;
coating weight adjustment is easy;
there is a low oxidation level (closed system) [86];
coating weights as low as 10 g/m2 are possible;
a wide web is possible [87].
The disadvantages of the system are the following:
no changes in the recipe are possible (without cleaning);
frequent cleaning of the die is necessary [86];
there is a sensitivity towards changes in the thickness of the
web [88];
there is a sensitivity towards fine adjustments of the die;
there may be lines in the adhesive layer.
Manufacture of Pressure-Sensitive Labels 597

For hot-melt PSA coatings where usually coating weight and viscosity are
rather high, the system is entirely heated by diathermic oil and consists of
a slot-die and rewinding bar, an interconnection hose and hot-melt PSA
supply gear pump that are electronically adjusted to line speed in order
to apply the set and targeted coating weight at all line speeds. This system
possesses good reliability, coating thickness and uniformity, and adjustment
and repeatability of the production conditions. If the coating device is fed
by a continuous extruder, the residence time of the molten PSA is not more
than 2–6 min (200–800 kg/h). Viscosities of 120 Pa s at 175 C are possible
and the energy input of the system is 7–12 kWh/kg [88].
Generally, cleaning of slot-die coating devices for hot melts is difficult.
There are formulations which are not mutually compatible therefore for
rapid changes of formulations slot-die is not recommended. Cleaning of the
machines which work with water-based PSAs can be carried out using water
(at least theoretically). However, sometimes the dried adhesive needs the use
of solvents.
Metering devices for GID and Duplex systems, used mainly for hot-
melt and high solid solvent-based coatings (mainly for tapes), show the
following advantages [46]:
lower costs for drying and solvent recovery;
higher productivity (30%) at a maximum of 420 m/min;
lower installation costs;
lower pollution;
simpler cleaning;
less space required;
higher accuracy;
higher solids;
lower shear;
no foaming.
The GID device possesses an interspace between the die and the
coating roll; the adhesive is fed into this interspace with or without the pump
(Duplex/GID). Advantages include the absence of turbulence, foam, and
residues on the die. The system was originally developed to be used for all
coating media (i.e., emulsion, hot-melt, and solvent-based PSAs) [89]. It has
an entirely new slot-orifice die, with a liquid pumping section built into the
die. The pump is designed to cover the complete die length. The coating
die is hinged for quick opening and easy cleaning; adhesive transfer to the
coating slot is given via a hydrodynamically acting rotor over the entire
width, with rotor speed adjustable according to media viscosity and required
coating weight, and no pressure builds up in the coating die (thus there is no
598 Chapter 9

overflow and build-up of adhesive on the edges). Curtain coaters are special
slot-die coaters where the coating nozzle is built up of small orifices.

4 DRYING OF THE COATING

After coating solvent-based or water-based adhesives the liquid adhesive


carrier has to be removed (i.e., a drying channel or tunnel is required).
Recent developments for drying tunnels have been discussed [90–98]. The
main component of a dryer is the adhesive drying tunnel, operating in most
cases with heated air. The web travels through the tunnel supported by a
conveyor belt, idlers, or air (contact free).

4.1 Adhesive Drying Tunnel


The design of a drying tunnel and its parameters depend on the nature of the
adhesive to be coated and dried. The choice of the conveyor belt and length
of the drying tunnel is of fundamental importance and is selected depending
on whether the application uses solvent-based or water-based adhesives, the
desired dry coat weight, the solids content, the operating speed, the type of
web, etc. The various sections with independent air flow rate, temperature,
and exhaust/recycle features can be selected and/or combined; among them:
standard rollers;
rollers with adjustable air over/underneath the web;
air flotation;
conveyor belt;
festoons.
Figure 9.26 shows the main types of drying tunnel.
The influence of the formulation on the drying was discussed in detail
in [99]. The choice of drying system for solvent-based or water-based
adhesives (or both) must take into account the relative evaporation curves
of the solvent or water through the various sections. In the case of solvents
an increased amount of solvent should be evaporated at the beginning of the
tunnel. For water-based coating the water evaporated is maximum in the
middle of the path. The drying conditions for water-based PSAs differ from
those of solvent-based ones. Water-based systems are more efficiently dried
by air volumes rather than air temperature. The temperature gradient
throughout the drying oven should also be different [100] (see also Fig. 7.2).
Energy requirements for solvent-based coatings are 170–200 cal/kg and for
water-based ones 400–500 cal/kg [101]. Low viscosity water-based coatings
can be dried at an evaporation rate of 10–20 kg/m2, typical solvent-based
Manufacture of Pressure-Sensitive Labels 599

Figure 9.26 The main constructive parts of the drying tunnel. A) Drying tunnel
with rollers, with air jets over the web; B) drying tunnel with rollers, with air jets over
and under the web; C) drying tunnel with rollers and conveyor belt; D) drying tunnel
with air flotation.

adhesives are dried with 20–50 kg/m2 per hour. On the other hand, safety
requirements require more air volume than necessary from a thermal point
of view. According to [102] when drying solvent-based adhesive the air
volume and speed must be large enough to avoid an explosive concentration
(s. lower explosion limit, LEL). As discussed in detail in [103] the air volume
necessary to dry a solvent-based coating depends on the lower explosion
limit which is a function of the solvent nature. Therefore it is necessary to
heat large volumes of air to keep solvent concentrations low. Water-based
adhesives do not need such air quantities, thus energy requirements for such
adhesives are lower. For example, drying a water-based dispersion requires
2.2  105 kcal/h compared to 7.61  105 kcal/h for a solvent-based one. A
common mistake is the application of too much heat [35]. This leads to the
‘‘flash off’’ of the upper water layer and skinning occurs. For aqueous PSAs
drying is generally more efficient with multizone ovens programmed for
gradually increasing temperatures. Typically, a first zone temperature of
160 F may be utilized with subsequent gradients up to 200 F. Such designs
have eliminated the skinning and coating defects [46]. If an air-flotation
oven is used, the distance between the coating surface and the nozzle is quite
critical. Adjustments of as little as 12 mm can create a totally new drying
600 Chapter 9

environment as this gap defines the actual drying temperature [35]. Drying
channels have slot-dies for the air (8–10 cm between the dies), situated
2–4 cm from the web; the air flow rate amounts to about 60–80 m/sec [104].
At the entrance of the channel, the temperature should be high enough in
order to evaporate large water quantities. After a decrease of the temperature
in the middle, the temperature should increase again at the end in order to
eliminate the remainder of the humidity from the adhesive.
Different drying methods are actually used (hot air, IR, radiofrequency,
UV, electron beam) [105,106]. Narrow web modular label presses include
UV drying, IR drying, and hot air drying [107]. Generally, the drying
time depends on the adhesive and on the drying conditions. The drying
parameters depend on the coating weight, and the temperature, volume,
and speed of the air. The air flow rate depends like other aerodynamical
parameters on the machine construction. For aqueous dispersions drying
also depends on the pH.
There are a lot of literature data concerning the drying of adhesives in
general and PSAs in particular [103,107,108]; as mentioned above, solvent-
based and water-based adhesives require quite different drying conditions.
Concerning the speed of the air, the blow off (blow away) of the coating
should be avoided through a limited air speed (i.e., 25–45 m/sec) [107,109].
The solidification of crosslinkable systems may be a complex problem too.
During their coating the removal of the solvent must be first carried out
at a lower temperature. For instance, in the case of silicone PSAs higher
temperatures can cause the peroxide to decompose prematurely and cross-
link the solvent into the adhesive [110].

Infrared Drying
Infrared heaters generally are used in plastics processing and coating
[111,112]. Wavelengths are selected as a function of the product to be dried
[10]. Infrared heaters used as preheating displace 10–30% of the humidity
from the coating. Infrared drying also may be used for paper coating
[113,114]. Infrared drying may avoid adhesive penetration [115]. Short
wavelength devices have a 50–300 kW/m2 output, 80% IR heat output, and
fast on/off switching. The wavelength of IR dryers for water-based adhesives
should be in the domain where maximum IR absorption by water occurs
[116]. Infrared drying and equipment have been described [12]; heating
element parameters, air knives, exhaust and cooling, and processing
considerations are reviewed and an example of an IR dryer calculation is
given. Infrared radiation with a wavelength of 0.76 m to 1 mm should be
used; the highest energy transfer is given at a wavelength shorter than 2 m
[117]. In IR drying the thickness of the water-based layer influences
Manufacture of Pressure-Sensitive Labels 601

the absorbtion wavelength. Thinner water films absorb more energy in the
shorter wavelength domain [118]. Such coating weight related considerations
may influence the adhesive formulation also. The advantages and dis-
advantages of IR/UV drying have been discussed [119]. Broad-band infrared
radiant driers also may be used. The main benefit claimed is that they greatly
speed up drying, cutting the residence time by at least half. Another drier,
the air knife (originally developed for use with water-based coatings), blows
thin curtains of very hot air across freshly coated web at speeds of up to
1400 m/min. Another system utilizes convection to achieve high heat transfer
rates without the web being lifted from the nozzle or disturbing the coated
surface. New developments offer cassette-based drying systems to provide
the best drier for each type of web with rapid changeover.

Flying Dryer (Flotation Dryers, Contact-Free Dryers)


The construction of a flying dryer has been described [42]. The distance
between web and dies is about 3 mm [120–122]. The air speed amounts to
about 50 m/sec, the temperature 250 C, and the dryer length 3–30 m. The
primary advantage of using a flotation drying system is the elimination of
damage to the web or coating because there are no idle rollers. The flotation
system allows full web coating coverage and the web must be run with much
lower tensions. Cleaning downtime also is reduced [16]. Flotation driers for
paper and films were discussed by Drechsler [42,123].

Radiofrequency Drying
Radiofrequency drying may be used also [124]. Radiofrequency (RF) drying
is used mostly for PSA-coated paper material [93,94]. It is a characteristic
known as the ‘‘dielectric loss factor’’ which determines how readily a given
material will respond to RF energy. Typically for tap water the loss factor is
100 at an RF value of 10 MHz (i.e., at a frequency of 10 million oscillations
per second); for paper it is less than 1. Thus, wet areas of a product are
selectively heated and dried while dry areas are unaffected [125].

Ultraviolet and Electron-Beam Dryers


The use of UV predominates when drying inks, coatings, and adhesives in
printing and converting operations [126]. Heat-sensitive substrates may
run at maximum speeds without damage or change of moisture content of
the web. UV and electron-beam dryers are much more energy efficient than
thermal dryers, as they do not depend on heat. The high capital cost of
electron-beam dryers realistically makes them only suitable for use as a final
cure system. In flexo-printing using an electron beam-curable ink on a PSA
602 Chapter 9

hot-melt laminate, the electron beam radiation provides at the same time
cure of the print and after-cure of the hot-melt PSAs. The advantages
of high speed electron-radiation curing include the absence of solvents or
monomers in the converted product, the improved web profile achieved
by ‘‘cold’’ running, short web paths, economical operation, easy cleaning,
quick changeover, and the absence of the explosion risk.

5 ENVIRONMENTAL CONSIDERATIONS

New regulations for the safeguard of the environment have recently been
enforced in Europe, imposing the recycling of packaging materials. The
majority of pressure-sensitive products are used for such materials. Recycling
comprises economical and environmental aspects. It includes waste manage-
ment also. The laws state that waste should be minimized or reutilized.
The most favored approach by industry for long term solutions to solid waste
management is source reduction. The second possibility is recycling [127].
Today postconsumer reclaim is a reality [128]. In a similar manner as in the
whole economy, one can speak about pre- and postconsumer recycling in
the converting industry also [129]. It should be emphasized that recycling
of composites having components of quite different chemical composition
is difficult. As stated by Münstedt and Wolter [130], thermal use is the best
possibility for reclaim of such materials. Recycling concerns the reuse of
the product components (e.g., carrier, adhesive, solvent, release, etc.) and of
the finished products too (e.g., labels, tapes, etc.).

Recycling of the Product Components


Recycling of the product components includes the constructive parts (e.g.,
carrier, adhesive, etc.) and the technological components (e.g., solvents,
dispersants, etc.) of the products.
Recycling of the Constructive Product Components. As discussed
above, the main product components are the solid state carrier material (face
stock and liner) and the adhesive. Their recycling before use (i.e., before
lamination) is a current economical problem of the production technology.
The recycling of the solid state laminate components (coated or
uncoated) which cannot be used for the manufacture of PSPs is carried
out by special firms according to the general technology of paper or film
(plastics recycling). Paper recycling is a classical technology today.
Recycling of plastic films is more complex because of the broad range of
the materials used and their incompatibility. Relatively easier to be recycled
are the polyolefins. As stated, a low level (5% w.) of PE in PP or vice versa
Manufacture of Pressure-Sensitive Labels 603

does not disturb the quality of the recycled plastic [131]. This technology
is used in the manufacture of the plastic film itself. The extrusion of plastic
films or coating produces scrap in the form of the edge trim, start up waste,
trim during slitting, and influences profitability [132]. The recycling of
plastic carrier wastes by refeed in the machine during processing may save
5–10% of raw material costs [133]. The most important approach is in-line
recycling of regranulated scrap, without reextruding. Problems appear if the
recycled amount exceeds 25–30%.
The recycling of the adhesive is carried out by its reformulation, i.e., its
use as an auxiliary component (at low level) in the formulation. Concerning
the recycling of adhesive waste in Germany the following code numbers are
valid for waste classification: 55901 for residues of glues and adhesives and
55903 for noncured residues and resins. A draft [134] differentiates between
noncured resins, noncured residues of adhesives, cured residues of
adhesives, and cured residues of resins.

Recycling of the Technological Product Components. Technological


components are materials used during manufacturing of the PSPs. Different
solid state or liquid components serve as such materials in the production of
PSPs. Most of them are used in other industrial domains also. For these
materials several recycling technologies have been developed.
For most of the coated PSAs (excepting 100% solids, including hot-
melts and radiation cured adhesives), a liquid status, i.e., the use of a solvent
or dispersant is required. These are technological aids only, they must be
eliminated after coating. Today, solvent-based adhesives are facing a
continuing attack from environmental regulations. The impact of current
and future clean air legislation, and the impact of the EU packaging, and
packaging waste directive on self-adhesive laminating regulate the necessity
and the ways of solvent capture and elimination [135]. Environmental
requirements have affected the industry since the early 1970s. The emission
protection laws have imposed severe restrictions on the emission of the
solvents into the air. Such restrictions and economical considerations have
imposed the development of solvent eliminating or recycling methods. In the
early 1970s environmental and energy considerations forced the adhesive
industry to introduce solvent-free systems. In 1977 and 1990 Clean Air
Amendments appeared [136]. In Germany printing and converting machines
have to reduce the solvent emission to 150 g/m3 by a mass current of more
than 3 kg/h. In the USA since 1982 converters have had to fulfill the
requirements of government regulation concerning volatile organic
compound emissions. In 1994 the European Parliament and the European
Council adopted Directive 94/62/EC on packaging and packaging waste
[137]. According to this directive 60% by weight of the packaging materials
604 Chapter 9

must be recycled [138]. The favored approaches by industry for solutions to


solid state waste management are in the following decreasing order of
priority: source reduction, recycling, incineration, and land filling. Thermal
recovery is not accepted as a form of recycling by the German Packaging
Ordinance. The latest amended EC directive accepts thermal recovery by
incineration as a possibility for recycling providing it does not exceed 30% of
the packaging waste stream. In addition to compliance orders and to meet
clean air act deadlines, converters introduced solvent recovery or elimination
(incineration) technologies. A conversion process and products with
minimum residual solvent content are desired [139]. Converters can meet
the requirements of the clean air acts through a variety of methods: the use of
incinerators, solvent recovery, solvent-free reactive systems, high solids
coatings, and WB coatings. Each of these options possesses advantages and
disadvantages. As is known, ‘‘pure’’ water-based systems may contain
solvents also. Wetting problems can be corrected through addition of
cosolvents. Water/solvent mixtures have lower surface tension. With water-
based formulations waste problems may appear also. A drum of aqueous
product may generate 159–300 L of waste from flushing equipment and
cleanup. The HMPSA technology produces only very little pollution to the
air and no pollution to water. It has been shown that HMPSA blends
evaporate at 200 C no measurable amounts of hydrocarbon cracking
products. Usually the working temperature with HMPSAs does not exceed
180 C. Bulk HMPSA waste can either be remelted and reused or disposed
with landfill, or incinerated.
During drying the liquid carrier (solvent or water) is evaporated;
recycling this is very important. Solvents may be incinerated or recycled. Air
cleaning for printing and coating needs the same technology [140]. As stated
by Nitschke [141] a conventional technology with solvents (for adhesives
and printing inks), can use one solvent or a mixture of different solvents.
Solvent recovery or elimination methods depend on this choice. Recyclability
of solvent for acrylic PSAs requires a low monomer content of the solute and
a simple mixture of hydrocarbon-based solvents which are insoluble in water.
Technologies based on one solvent are adequate for solvent recovery;
multisolvent technologies should use incineration. Unfortunately the use of
water and solvents may produce difficulties in solvent recovery also, because
of the mutual solubility of water in solvents and vice versa (azeotropes).
Solvent and energy recovery should be coupled. Toluene replacement in
solvent-borne pressure-sensitive adhesive formulations has been a major
requirement. A high solvency hydrocarbon solvent system with a high
content of naphthenic species has a low affinity for water; it may be an ideal
candidate for currently used solvent capture and recycling equipment. PSAs
based on synthethic rubber use special petrol fractions as solvent.
Manufacture of Pressure-Sensitive Labels 605

The recycling of solvents has been discussed in [140,142–149]. For


many years there were very satisfactory means of recovering solvents and
preventing emissions into the air; various alternatives are available:
Absorbtion by active carbon and dissolution by steam. This is the
most common method; this system is suitable for hydrocarbon
solvents, which covers most rubbers and some acrylic PSAs.
Recovery by condensation. This system is more expensive, but
suitable for all solvents and yields the solvent or solvent blend
without contamination [150].
Afterburning. As selling recovered solvents becomes more difficult
and complicated solvent blends become more prevalent in the
manufacture of high performance acrylic adhesives, many manu-
facturers are changing over to afterburners on the drying line.
When the solvent vapors are concentrated in nitrogen before
burning, the afterburning can yield an energy output that can
be used to heat the drying air [151].
Biological cleaning of the air is used also [152]. Solvents that are easily
recoverable are used routinely; products are also polymerized in toluene and
hexane. Generally, the following solvents are used: toluene, hexane,
ethyl acetate, and isopropyl alcohol. While taking the decision for one or
the other process the following factors must be considered: the number of
solvent components, the concentration of the solvent and solvent
components, and the chemical nature of the solvent components. The
applicability of the main recycling methods of adsorbtion onto activated
carbon, washing, or condensation of the solvents depends on the chemical
composition of the dispersing/solving medium used in the manufacture of
a pressure-sensitive product. For instance, water soluble solvents such as
alcohol, acetone, methyl ethyl ketone, and ethylacetate (partially) can be
washed out. The different solvent recovery systems and the maximum
emission concentrations are listed by Wittke [143].
The two most prominent processes in the recovery of solvents from the
process air are adsorption and condensation. Both procedures (adsorption
and condensation) can be combined to achieve a higher concentration
of solvents in the air. Adsorption is accomplished on a bed of activated
carbon, where the solvent is trapped by the carbon and is recovered by
steam stripping. For the solvent recovery systems based on adsorption,
principally the pressure swing adsorption (i.e., desorption by changing
of the pressure) and temperature swing adsorption (i.e., desorption by
changing of the temperature) are employed. In this case the humidity
content of the solvents may influence its efficacity, because of the com-
petitive adsorption of water on the carbon bed. Generally blends of water
606 Chapter 9

insoluble aliphatic and aromatic solvents are needed to be recycled using


such methods. Solvent recovery by adsorption to activated carbon exhibits
the following advantages: high flexibility regarding concentration and
composition of the used air, as well as high degree of recovery. Generally
recycling via adsorption on activated carbon is suggested for hydrocarbon
derivatives and thus, it limits the choice of the base polymers also. Tape,
plaster, and label manufacture use petrol with low toluene levels as solvent.
For such technology adsorption on a carbon bed is recommended. For
this procedure the solvent concentration in the air should be of about
10–20 mg/m3. Because of the high boiling point of the solvents used in
printing, adsorption on a carbon bed is not suggested for recycling.
Adsorption allows advanced elimination (99%) of hydrocarbon
solvents. Adsorption on active carbon is suitable for hydrocarbon solvents
in rubber-based adhesives and for some acrylic PSAs. For instance, the
pressure-sensitive adhesive Durotak 280-2366 (having only toluene and
hexane as solvent) is the recyclable grade of the Durotak 280-1151 (having
an ethylacetate-hexane blend as solvent). It possesses lower viscosity, solids
content, and molecular weight [153]. The presence of reactive materials (e.g.,
isocyanates, phenol, free monomers, and ketones such as methyl ethyl
ketone) can be troublesome. Ketones, which have a higher heat of adsorption
than many other solvents, are of particular concern and generally their
presence is forbidden in solvent-recoverable adhesives. The water insolubility
is a prerequisite for the versatility of this procedure. The adsorption method
using activated carbon gives a water-solvent mixture. If esters of chlorinated
solvents are present, corrosion problems may occur. Such a method was
suggested for more than 800 t solvent/year [154]. If hot inert gas is used for
the desorbtion of the solvents from the carbon bed, there is no need for the
distillation of water soluble solvents and the explosion danger is reduced too.
In condensation the solvent is recovered by lowering the temperature
of the process air to a point necessary to condense the solvent and lower
the saturation concentration of the process air. The basis loading of the air
is a function of the condensation temperature of various solvents. The most
used industrial solvents are mixtures of two components. The vapor
pressure of these mixtures may be calculated using Dalton’s law. For their
composition Raoult’s law may be applied. The allowed concentration of
common solvents has been 0.5–2.0 g/m3. Recycling by condensation does
not have important solvent imposed limits. The solvent recovery via
condensation using a closed cycle with nitrogen as carrier gas allows the free
choice of the solvent concentration [144,155]. It is suggested for high
solvent concentration in the drying air [156], but it may be used if the
exhaust air has low concentrations of contaminants (up to 3 g/m3) too [152].
Recycling via condensation offers a broader range of usable solvents than
Manufacture of Pressure-Sensitive Labels 607

adsorption. Solvent recovery by condensation is more expensive but suitable


for all solvents and does not produce contamination of the solvent. It
works at about 38 C and allows 70% retention of the organic solvents.
Condensation does not work with isopropanol [157]. In comparison to other
systems (e.g., adsorption and thermal postcombustion) condensation allows
an energy saving of 40%.
Adsorption of the solvents with molecular screens (e.g., natrium-
aluminum silicates) is recommended for flexo-printing. In this procedure
(because of dry regeneration of the adsorption bed with hot air) no water
is introduced in the solvent.
Air recycling is related to solvent reuse or incineration. Exhaust and
recycling has to be adjustable as a function of the drying of solvent
concentrations. Air containing toluene has been recycled by thermal
incineration equipment which allows the incineration of toluene rests
(20 g/m3 or 200 kg/h/m3). Afterburners on the drying line are used too.
For halogenated solvents incineration is not possible.
Recycling of the dispersant (solvent) of water-based formulations is a
complex but classical technology. It includes the separation of the solid state
components, their incineration, and the purification of the water. Water-
based formulations have to be neutralized and coagulated before mixing
with common waste waters. Waste water treatment for aqueous adhesive
dispersions is known and was described several years ago [158]. According
to such technology water soluble components are unsolubilized by
polyvalent components. As is known, ferrichloride and catalytic destabiliz-
ing agents in alkaline media precipitate water-based dispersions. Other
chemicals (e.g., aluminum sulfate) have been introduced also. Fluid, one-
component flocculation agents were developed in order to allow waste water
purification from dispersions. Their efficacity depends on the chemical
nature and formulation (surfactant and protective colloid) technology
of the water-based dispersions. Protective colloid-based dispersions (e.g.,
ethylene-vinylacetate) are processed more easy than surfactant stabilized
ones (e.g., acrylics). Therefore the formulator has to take into account the
surfactant nature and level in such dispersions by the design of the manu-
facturing technology. Generally the manufacturers of such dispersions
supply the waste management technology for their products also. In the
USA incineration or water-based techniques are the current choice to
comply with Federal Air Emission Requirements.

Recycling of the Pressure-Sensitive Products


Recycling of PSPs requires the technology to transform heterogeneous pro-
ducts, having chemically different components, in chemically homogeneous
608 Chapter 9

products. This is possible if their components are separated during recycling,


or if their level in the ‘‘new’’ products is limited. It should be emphasized that
product heterogeneity as a main problem of recycling is more eloquent in the
reuse of packaging materials, independent of whether they contain PSPs. As
discussed in a detailed manner by Maack [159], the development of packag-
ing technology leads to a broad range of heterogeneous materials with a
sophisticated construction. Puncture, impact, and tear resistant shrink and
stretch films (e.g., LLDPE, VLDPE, and LLDPE/LDPE blends and co-
extrudates) replace craft paper wrapping. Cling films (13 m to 15 m) as
wrap, household and catering film (LLDPE blends with EVAc or PP) replace
PVC. Heavy duty bags produced from co-extruded, blown LLDPE/HDPE/
LLDPE or similar co-extrudates replace multiply paper bags. Recycling
problems exist with melt index increases, oxidative color changes, and gel
content increases. The presence of certain fillers, pigments, and lubricants
may cause gel. Many hygienic plastic applications (diapers, garments, drapes,
sanitary towels etc.) utilize highly filled materials. Some companies develop
products with degradable layers to facilitate accelerated decay if disposal is
not by incineration. In medical packaging injection and co-injection molded
PET, PP/elastomer, and HDPE/elastomer blends find increasing use, also
co-extruded PP/HDPE, HDPE/LMDPE, HIPS/HDPE, etc. Various label
carrier materials may increase the product heterogeneity. Although the
recycling of such plastic materials must be solved by the processing of
plastics, the suppliers of PSPs have to use compatible carrier materials and
adhesives or composites with improved component separation ability.
Developments in the recycling of plastic materials allow their reuse as
film, fiber, or items. Wide mouth glass jars for coffee and milk powders have
been substituted by co-injected PET barrier containers [159]. PET recycling
of postuse containers is well established. PETP can be recycled by
decondensation. For current, nondepolymerizing plastic films, co-extrusion
encapsulation with scrap reclaim built into a middle layer is common.
Reusable textile materials as goretex, polyamide, polyether sulfone/cotton,
100% cotton, or other hydrophobized carrier materials can be used also;
they are suggested for medical PSPs (operating tapes, labels, and
bioelectrodes) [160]. Plastic carrier materials can be easier recycled. Top
coated oriented polypropylene (OPP) films (50 m and 60 m) exhibit
recyclability with plastic containers [161]. Siliconized OPP can be recycled
via extrusion into fresh film [162]. A level of 5% PP in PE or vice versa
is tolerable for recycling of such films [163]. Recycling of protective films
coated with rubber-resin PSAs is carried out without problems; acrylic
coated films impose the use of degassing extruders.
Generally there is a trend to use papers including 30% recycled
paper [164]. The use of recycled papers has to be taken into account by the
Manufacture of Pressure-Sensitive Labels 609

formulation of special packaging adhesives. For instance, Escorez 2184 is


suggested as tackifier for NR-based rubber-resin formulations for packaging
tapes, particularly for recycled cardboard surface. Polyethylene-based
polymers for hot-melts (e.g., Epolene from Eastman) separate easily from
paper fibers which is important for paper recycling [165]. Due to the various
coatings of the label material this waste cannot simply be recycled as paper.
Actually PSAs used for labels or pressure-sensitive envelopes are not
recyclable in the pulp and paper industry [166]. There is a need to develop
a technology that can process the ‘‘stickies’’ generated when materials
containing PSAs are used in the manufacture of recycled fiber pulp [167].
The USPS has issued an ever increasing number of nonlick self-adhesive
stamp products. The use of self-adhesive stamps grew from 0% in 1990
to more than 93% in 2000 [168]. Their use can cause problems during
recycling. The PSAs can ‘‘gum’’ up the works for recyclers. When they are
introduced into the paper, recycling stream they break down into small
particles called ‘‘stickies.’’ The stickies adhere to wires and felts in the
equipment causing sparks or tears and downtime. Paper recyclers are
reluctant to accept waste paper that has a high PSA content. The current
practice is to attempt to remove material containing PSA from the waste
stream by sorting it out prior to repulping, or by removing stickies from the
pulp using mechanical screens and dispersing techniques. Because of their
insolubility the waste resulting from the production of tapes or labels cannot
be easily reclaimed; on the other side it is difficult to remove the labels
and their residues from high quality goods (e.g., glass or porcelain). An
experimental procedure for evaluation of the recyclability of pressure-
sensitive labels has been developed by FINAT [169]. Release papers can be
recycled only if they do not contain Sn catalyst or emulsion silicones [170].
Recycling of the face stock materials is facilitated if separation of the
PSA from the carrier is possible. This may be achieved using water soluble
adhesives or carrier materials. As discussed earlier (see Chap. 5) there are
several formulation possibilities to manufacture such adhesives or carrier
materials. Some of them can contain as base polymer a water insoluble
common plastic material. A special method to assure recyclability of tapes
concerns their water solubility or dispersability. Both the adhesive and
the carrier can be formulated as water dispersible products. Fully water
soluble plastics can be used as raw materials for films also. Their industrial
use has been hindered mostly by economical considerations (price level
of 1000–300% in comparison to common plastics). A paper-based tape,
recyclable as paper is manufactured using a water soluble adhesive based on
alkali soluble acrylic acid esters and polyethylene or polypropylene waxes
and alkali dispersible plasticizer. Repulpable splicing tape especially adapted
for splicing of carbonless paper uses a water dispersible PSA based on
610 Chapter 9

acrylate/acrylic acid copolymer, alkali and ethoxylated plasticizers. A


polyamide-epichlorohydrine crosslinker may also be included. Polyacrylic
acid blended with polypropylene glycol has been used for splicing tape
formulation in order to allow water solubility. Polyacrylate tackified with
water soluble PVE has been proposed also. Hot-melt PSAs which can be
applied as hot-melt but which allow the repulping of label stock waste and
the easy removal of labels in water would be attractive also. A hot-melt
which is water dispersible and biodegradable, for paper bonding packaging
application was developed using modified starch as base polymer [171]. For
adhesive use, high amylase starches were modified by acid hydrolysis and
by chemical substitution of the pendant hydroxyls (to improve compatibility
with additives). Such adhesive displays 70% biodegradability in 30 days.
The first totally water soluble (carrier and adhesive) labels were marketed
in Europe in 1999 [172].
Biodegradable polymers capable of being processed in films were
developed and are under development. Other recycling technologies are
based on thermoplastic starch compositions. Water resistant, fully
biodegradable, starch-based plastics were produced [173]. Biodegradable
polyester or compounds of starch with common polymers are known [174].
Additives for acceleration of the biological destruction of plastics were
introduced. Master batches for the biological degradation of polyolefins are
supplied too [175]. The properties of such degradable polyethylene films
are described in a detailed manner in [176]. A special plasticizer can be
incorporated into cellulose acetate to allow its decomposition in 6–24
months [175]. Additives based on starch and fatty acid derivatives have been
developed. A level of 25% of such additives allows biological degradation
of polyethylene in some months [177]. Such materials suffer two technical
limitations in their use: low processability (narrow processability window)
and low impact resistance. Biodegradable polymers processable on existing
extrusion-coating machines are under development, fully biodegradable
and compostable materials (e.g., fermented polyester) have been developed
also. Films made from copolymers of PVA with more than 50% starch,
modified thermoplastic starch, cellulose, and PVA have been tested as
biologically degradable; other films made from special polyester and
polylactides are compostable [178]. Biodegradable cellulose acetate is
made using a special plasticizer. There are degradable PE films (photo-
degradable, biodegradable). Formulations of PE with degradability
additives can be produced [177]. These are mixtures of starch, nonsaturated
fats and oils respectively as additives for build-up of peroxides. At least 20%
of the ingredient is added in order to achieve a self-destroying ability of
5 years. Different steps of self-destroying occur: biological, chemical-
physical, and again biological. Polypropylene film used as carrier material
Manufacture of Pressure-Sensitive Labels 611

may be destroyed biologically also [179]. The results concerning the


practical use of these films are very controversial. Tests with starch-filled
biologically degradable films have shown that the biodegradability of the
film does not fulfill the practical requirements [180]. Biodegradable polymers
have been synthesized also. Polyhydroxybutyrate, polycaprolactone, poly-
lactides, PVA-containing compositions, and cellulose diacetate are supplied
as biodegradable plastics. Some of these materials, e.g., ‘‘Biopol’’ (fermented
polyester) of ICI/Zeneca can be processed into film. Biopol is an aliphatic
polyester, i.e., polyhydroxybutyrate (PHB). Its physical properties, melting
point, glass transition temperature, and crystallinity are comparable to
those of polypropylene. Copolymers with polyhydroxyvalerate (PHV) can
be synthesized also, and allow the regulation of the mechanical charac-
teristics. Such polymers can be mixed with nonbiodegradable polymers
to allow their biodegradation. Poly-D-()(3-hydroxybutyrate) is biodegrad-
able and highly biocompatible, but it possesses low impact resistance and
a narrow processability window. Polyhydroxybutyrate and polylactides do
not have BGVV approval [181]. Another biodegradable material ‘‘Biocell
163’’ is a clear, amorphous polymer, based on cellulose diacetate. It exibits
thermal stability up to 60 C and a tension yield like that of polyethylene
[182]. Such materials are more expensive than common plastic grades. For
instance PHB was first introduced industrially in 1990. In 1990 its price was
eight times higher than that of similar materials, in 1993 its price was still
about five times higher than the price of common plastics [183].
Hot-melts can be manufactured on a starch basis also. Polyesteramide
(a partially crystalline material with a melting point of 125 C) has been
developed also for hot-melts. It possesses an unlimited storage stability
under usual conditions. It will be depolymerized in the soil only [184].
Aqueous dispersions of vinyl ester polymers which are biologically
degradable can by prepared by free radical polymerization in the presence
of 10% to 50% bw. of biologically degradable plasticizer and of 0.5–15%
biologically degradable emulsifier or protective colloid [185]. As stated
in [186] almost two decades ago, because of their limited performance
characteristics and excessive costs, ‘‘the scope of completely biodegradable
and compostable plastics (and adhesives, N.A.) is still limited in marginal
areas.’’ Therefore the recycling of adhesives during production is
recommended. Advances in the synthesis and formulation of biologically
degradable raw materials for adhesives are discussed in [187,188].
Research was carried out in order to develop bio-based adhesives,
using renewable raw materials also. Such PSAs were derived from plant and
animal oil triglycerides and monoglycerides with chemical groups (epoxy,
carboxyl, hydroxyl, vinyl, amine, etc.) that render them polymerizable [189].
The monomers for PSAs are derived from chemically functionalized
612 Chapter 9

triglycerides, diglycerides, and monoglycerides with controlled fatty acid


distributions. This allows the molecular architecture of PSAs to be designed
with linear chains, branched chains, and structures with controlled crosslink
density. In addition to the skeletal structure obtained by the polymerization
process, the polymer structure can be functionalized with groups whose
purpose is to control the adherence to the substrate. In other tests plant oils
were used as starting material, e.g., high purity oleic methyl ester. In such
compounds the single double bond allows for easy synthesis of linear poly-
mers. These base materials were transformed into acrylic derivatives [190].

6 SIMULTANEOUS MANUFACTURE OF PSAs


AND PSA LAMINATES
6.1 Radiation Curing of PSAs
Chapter 5 discussed some special cases where the monomers to be used
for manufacturing the adhesive are not polymerized, or polymerized only
partially and coated as such; in these cases, the PSA is manufactured during
the coating. These are special materials with the ability to polymerize through
radiation (electron beam or UV). As an intermediate step, there are special
‘‘classical’’ PSA formulations which can be dried or cured using radiation as
well. Although an intensive effort was carried out in this field, electron beam-
or UV-cured PSAs or PSA-related materials are still considered exotic.
Pressure is growing on converters using solvent-based materials to
look for methods which are environmentally acceptable [191]. Here the field
of chemistry can offer new groups of products, namely systems with a low
solvent and a high solids content (i.e., water-based systems, coatings with
solid resin in powder form, and 100% prepolymerized solid systems). The
following feasible solutions as alternatives for solvent-containing formula-
tions are known for PSAs: formulations on an aqueous basis, formulations
on a hot-melt basis, and radiation-curable and crosslinkable formulations.
Numerous papers have been published in the past several years in this field,
some of which could serve as a basis for industrial applications. Techniques
which do not include the disadvantages of thermal drying are the various
methods of radiation drying, in particular those including ionizing radiation
such as ultraviolet- and electron beam-curing. In this case the energy is
transferred without contact. Energy transfer can take place in a vacuum or
through a vacuum section. Electron beams belong to the directly ionizing
beams of energy levels above 3 eV. Some of the advantages of electron beam
processing for PSAs include higher productivity, compliance with regulatory
restrictions, enhanced profitability, and increased product development
capacity [192].
Manufacture of Pressure-Sensitive Labels 613

Styrene-isoprene-styrene block copolymers can be crosslinked to


become PSAs by means of reactive monomers under radiation [193–195].
Electron radiation crosslinking of PSAs has enjoyed industrial applications
for several years now making PSA tapes. Furthermore Huber and Müller
[196,197] were able to present new radiation-curable copolyesters. These are
saturated copolyesters containing reactive acrylate end groups [198]. Nitzl
et al. [194,199] and Forster [200] studied hot-melt PSA formulas based on
block copolymers; the base polymer used was Kraton D-1320X, a branched
styrene-butadiene copolymer with a functionalized end group in combina-
tion with hydrocarbon resins. The various formulas were crosslinked by
electron beam exposure at doses of 2 and 4 Mrad. The available results
show that there is no clear dependence of the adhesive properties on the
formulation [199].
As postulated by Seng [201], the most important application field of
radiation curing is the curing of silicones and PSAs. Release-coating curing
is done mainly for labels, while curing of the adhesive layer is mostly for
labels and tapes; a postcuring via an electron beam is made for tapes also
[198]. In all these domains, both techniques present advantages and
disadvantages [201]. Electron beam-curing is faster, up to 100–500 m/min
versus 100–200 m/min for UV, but needs a nitrogen atmosphere and may
damage paper. UV displays the disadvantages of rest monomers, a limited
thickness of the layer, and too low penetration of the radiation.
The state of development of radiation curing (electron beam and UV)
has been discussed [202–204]. The first application of electron beam-curing
for PSA tapes was described in 1954 [205]. Silicone acrylic PSAs were used
by Hurst in 1983 as release liner [206]. An in-line UV-curing hot-melt PSA/
silicone release manufacturing line was proposed in 1982 [206]. Other papers
discuss the problems of radiation curing for lacquers, printing inks, and
adhesives [207–211]. The coating and printing of radiation curing systems
has been discussed in a detailed manner [212–216]. The technology of
UV curing for varnishes and pigments in the printing industry (‘‘Face stock
printing for labels’’) has been discussed [217–221]. Radiation curing
possibilities for plastics and paper coating were covered by Röltgen
and others [222–227]. Guidelines for the formulation of radiation-cured
adhesives were reviewed by Hinterwaldner [228,229]. The electron beam-
curing and drying of varnishes and coatings was described by Strohner [230]
and Fuhr [231].

6.2 Siliconizing Through Radiation


An emerging coating application which lends itself to UV/electron beam-
cure technology involves silicone release coatings. There are several reasons
614 Chapter 9

for the interest in these new techniques. Because UV- and especially electron
beam-curing occurs at very low temperatures, these kinds of silicone release
coatings can be applied and used on heat-sensitive liners (e.g., PET,
polypropylene, or polyethylene films) and polyolefin-coated papers without
distorting such liners. For the same reason the application of radiation-
curable release coatings to Kraft paper liners eliminates the need for
remoisturization of the coated paper. The complete, immediate cure of
radiation-curable release coatings eliminates ‘‘blocking’’ of rolls of finished
product and provides greater assurance of uniform release characteristics
from the outside of a roll to its core [232]. Electron beam-curing enables
simultaneous curing of both sides of two-sided coated release liners. UV/
electron beam technology is a considerably more efficient means of curing
than thermal drying; the energy cost is less than that of conventional (i.e.,
thermally dried) release coatings and the need for large ovens disappears.
The practice of in-line mixing of release coatings with a relatively short pot
life could be eliminated with these types of coatings. These fully formulated
products typically have a shelf life of 6 months to 1 yr. Theoretically, the
‘‘hot’’ UV system, the ‘‘cold’’ UV system, and electron beam-curing can
be used [233]. In some cases solvent-based silicones were postcured with an
electron beam [234]. The coating weight for electron beam-cured silicones
amounts to 0.7–1.5 g/m2 with a curing time of 0.3 sec. In some cases
the release properties of electron beam-cured silicone papers improve with
storage [235]. The performance characteristics of electron beam-cured
silicones and the advantages/disadvantages of the procedure were summar-
ized by Brus et al. [236].
Radiation-cured silicones have to be coated as a continuous layer, at
a coating speed of 300 m/min, a coating weight of less than 1 g/m2, and with
a viscosity below 1500 mPa s. They contain at least 90% dimethylsiloxane.
Radiation-cured silicones are based on prepolymers; radiation-cured PSAs
also contain reactive monomers [200]. The most common radiation-cured
silicone systems are based on acrylic (silicone derivatives) which are cross-
linkable via electron beam or UV radiation. Another system is based on
mercapto-modified silicones; these are not air-sensitive and may be used for
UV curing. The initial tests were carried out with UV-cured silicones, mostly
in combination with hot-melt PSAs. In comparison with electron beam-
cured PSAs, electron beam-cured silicone liners need no more than a
0.5 Mrad dose. The highest efficiency is achieved by coupling hot-melt PSA
coatings with in-line siliconizing. In this case about 30% of the energy costs
may be saved [200]. Although new UV-curable silicones display improved
substrate and adhesive compatibility, it is not correct to assume that any
silicone will work with all adhesives. Such products possess a minimum
viscosity at 20 C of about 500 mPa s. Easy release products can run at a rate
Manufacture of Pressure-Sensitive Labels 615

of 70 m per 40 watts/cm of lamp power [237]. Such UV-curable silicone-


acrylate release coatings are catalyst free. They allow the use of low gauge
film liners, e.g., 30 m BOPP [238] at speeds of 900 m/min [239,240].

Electron Beam-Curing
Electron beam-cured silicone release papers exhibit different release forces
depending on the paper thickness [175]. A 67-g/m2 paper displays a higher
release peel force than a 90-g/m2 paper (84–98 mN/cm) [241]. Electron
beam-curing may damage the paper and increase its stiffness; PVC may
undergo yellowing. Furthermore, electron beam-cured silicone release liners
have limited storage resistance. In this application the electron beam-
curable silicone release layer is coated via rotogravure (1 g/m2) [242]. For
release liners manufactured via electron beam-curing a coating weight of
0.8–1.2 g/m2 and 0.4 g/m2 is suggested for paper and for films, respectively,
at a radiation dose of 1.5–2.0 Mrad [242]. An inert atmosphere is required
for this application; the oxygen concentration should be lower than 50 ppm
[243]. The advantages of radiation curing include the following [203]:
The paper is not dried.
Thermal sensitive films can be used.
No drying oven is necessary.
High coating weights are possible.
High running speeds are possible (up to 900 m/min) [209].
Two-sided in-line coating is possible.
On the other hand, some technological disadvantages need to be mentioned.
The prepolymer/diluting agent blend is highly viscous, so smootheners
are necessary [227]. Another problem is the toxicity of the monomers used
[208]. The average energy consumption for solvent-based or water-based
siliconizing is about 0.14 kWh/m2; a UV-cured release liner needs
0.0143 kWh/m2 [223].
In general, electron beam technology provides manufacturers of PSA
laminates with the following benefits [244]:
New products can be developed by electron beam-curing/cross-
linking using a variety of substrates, coatings, and adhesives, with
different property combinations.
A higher productivity is provided because modern electron beam
units operate at higher speeds and process wider webs with
minimum waste.
Electron beam users are in compliance with environmental regula-
tions because the systems use solventless adhesives, inks, and
coatings.
616 Chapter 9

Profitability is enhanced because electron beam technology can


generate new product and market opportunities whilst improving
the performance of existing products.
Reproducible product uniformity is achieved because electron beam
units ensure a uniform cure.
Electron beam-curing/crosslinking can enhance the shear resistance, the high
temperature resistance, and oil or solvent resistance of thermoplastic rubber
hot-melt PSAs. Integrating the electron beam processor in-line with the hot-
melt coating equipment enables single operation processing prior to slitting.
‘‘Dual cure’’ is a method where curing is initiated via radiation and
completed through a thermal method [245]. There are special SIS block
copolymers suitable for electron beam-curing [246,247].
There are two main types of electron beam generators: partial
beam and scanner [245]. The electrocurtain process was described by
Lauppi [248,249].

UV Curing
Limited space requirements and a low thermal loading of the face stock
material are the advantages of UV-cured silicones [250]. The coating weight
for UV crosslinked silicones may be adjusted within a tolerance of 5%.
UV-curing siliconizing machines were running in 1992 at a speed of
400 m/min [251]. In 1999 production speeds of up to 900 m/min were
attained [251]. UV-cured PSAs possess the following advantages [252]: rapid
curing, low energy requirements, limited machine preparation costs, absence
of solvents, and a long shelf life. They may be coated using screen printing
or roll coating technology.

7 MANUFACTURE OF THE RELEASE LINER


7.1 Nature of the Release Liner
Generally, the PSA label is manufactured as a laminate composed of a face
material laminated to a release liner with a PSA. In many cases the adhesive
is first coated onto the release liner and not onto the face stock. There are
two methods of depositing the PSAs onto the face sheet: direct and transfer
coating. The former deposits the adhesive onto the face material. In the
transfer method the adhesive is deposited onto the release paper and is
subsequently transferred to the face stock during the laminating process. It
is evident that especially in transfer coating, the performance of the release
liner will also influence the performance of the final PSA label. On the other
hand, the properties of the release liner depend on its chemical composition
Manufacture of Pressure-Sensitive Labels 617

and manufacturing conditions. Many converters manufacture PSAs and


release coatings and there are some common features for both technologies.
Although it is not the aim of this work to discuss the chemistry and
manufacturing of release liners their influence on the label (and label
technology) are reviewed next.
Other materials than silicone may be used as a release liner too.
Nonsilicone release coating materials are discussed in detail in [253,254].
The release surface will normally be provided by a layer of silicone resin or a
material providing similar release properties coated onto the liner. Silicone-
free vinyl acetate-based emulsion polymers were suggested as release coating
[255]. Derivatives of polyvinylcarbamate also are used for protective films
and tapes, but the release effect of silicones remains superior. Where a
coating of silicone polymer is used, the coating weight will typically not
exceed 2 g/m2 and a liner of 60–100 g/m2 should be used.
Paper and nonpaper materials are used as liners. As printing, convert-
ing, and label application speeds are increasing, web breaks and V-tears in the
liner cause unacceptable scrap loss and downtimes. Hence plastic films
are now competing with paper-based liners. Today solvent-based, solvent-
less, and emulsion (water-based) silicone coatings are routinely coated.
The liner is the packaging for the adhesive, the carrier material for
die-cutting, and sometimes the transparent support for the label [209].
Generally (as shown earlier in Chap. 7) carrier materials for PSAs are
flexible webs, supplied as rolls which are coated, dried, and rewound [5];
silicone derivatives are used as coating materials. Their function is to sepa-
rate the adhesive from the release paper, to offer an easy but controlled
and reproducible separation of the face paper from its adhesive layer
(or vice versa). A very low coating weight of silicone is deposited and a very
smooth defect-free surface should be obtained.
The weight of the release liner depends upon the method of label
converting, the type of die-cutting and tools, and the mechanical strength
of the sheet. The release coating is usually a silicone formulation and is
designed to provide a variable release level, again depending on the method
of converting and whether the adhesive is cooled over the whole web width
or is pattern or zone coated. For years the only release liners used
in pressure-sensitive label applications were calendered Kraft papers in the
40–50 lb weight range, with one side silicone coated. As processing speeds
increased, web breaks and V-tears in paper liners became a growing
concern. The use of 150-g siliconized PET as a release base resolved the
problems associated with high speed dispensing applications [256]. Paper
and polyester are just two of the commercially available release liner
systems. Oriented polypropylene and polyolefinic blends also are used
for high speed application systems. Another technology employs a liner
618 Chapter 9

structure of 30-lb Kraft paper extrusion-coated with 14 lb of polypropylene;


these materials offer a greater margin of security during die-cutting [257].
Polyethylene has been used as a release liner since 1970 [258].
Face stock materials and release liner materials were reviewed in
Chap. 7. A detailed description of the manufacture and performance
characteristics of such solid state carrier materials is given in [259]. Their
nature is important for quality control purposes because of their influence
on the release (peel) force. Papers used as release liners must display superior
mechanical properties, little thickness variation, and end-use tailored
properties [8]. The release force for pigmented papers depends on the
nature of the clay as well as on the level and nature of the resin [260]. The
curing method also influences the choice of the release paper [8]. The main
factors influencing the release force include the silicone formulation control
release agent (CRA) level, the coating defects present, the mechanical
destruction of the silicone layer, and the possible incompatibility between
release and adhesive formulation. Mechanical defects can be avoided by
using softer rubber-coated cylinders (60 shore hardness, SH) and the offset
process. The adhesive and release properties may be influenced by the
interaction of the adhesive and the silicone. This problem may be avoided by
using another silicone catalyst, but the activation temperature may increase.
High peel forces are necessary to separate water-based acrylic PSA labels
from Sn-catalyzed silicones. Only addition polymerization with a Pt catalyst
allows lower release forces [261].
Generally, the most difficult problem is the control of the silicone
coating weight. The silicone coating should not show any pores and should
display a good anchorage to the liner [8]. Appropriate amounts of solvent-
based silicones are 0.5–0.8 g/m2, and for solventless silicones, 0.7–1.5 g/m2
[260]. The reactive groups in the silicone or PSAs, their mobility, and the
mechanical influences determine the release properties [262].
Usually the release liner displays the minimum resistance to debonding.
However, this low release force is inadequate in certain applications as
greater resistance may be required. In these cases special silicone resins
(control release agents) are used. These agents must be compatible with the
base silicone resins and be capable of being blended with them in any ratio;
various based resins are used as control release agents:
control release resins for polycondensation in solvent systems;
control release emulsions for polycondensation systems;
control release resins for polyaddition systems with and without
solvent.
Polycondensation siliconizing is not used in Europe [8]. In 1988 in
Europe, 50% of the silicones were solvent-based, 39% solventless, and the
Manufacture of Pressure-Sensitive Labels 619

rest water-based. Since the 1980s, solventless silicones have been used at an
increasing pace [262] attaining a level of about 60%.

7.2 Coating Machines for Silicones


The same coating machines as the ones used for PSA coating are suitable for
siliconizing. Coating machines for PSAs have been extensively discussed in
the literature and were described in a detailed manner in Section 2.1 [39,263].
Silicone coating machines are similar to the adhesive coating machines.
Different coating devices should be selected for different applications [39]:
in-line manufacturing: release-Corona-primer-PSA, Corona-
primer-PSA;
manufacture of very thin layers, like solventless silicones;
manufacture of self-adhesive medical products;
manufacture of self-adhesive special papers (crêpe);
manufacture of self-adhesive woven and nonwoven materials;
manufacture of self-adhesive tamper-proof and protective films;
manufacture of pattern coatings.
The rheology of PSAs differs only slightly from that of the silicones
(coating technology), hence similar coating systems may be used for both
systems. For solventless silicones a two-cylinder coating device was used
at 120 m/min running speed [264]. A four-cylinder geometry deposits
0.5–2 g/m2 solventless silicone [265]. A five-cylinder nip fed coating device
deposits 0.3–1.5 g/m2 [39]. The machine speed for siliconizing ranges from
50–600 m/min [266]. Siliconizing at 400 m/min ensuring a 5% weight
tolerance appears possible [8,267].

7.3 Technology for Solvent-Based Systems


Different coating geometries were designed for different silicone/release
types [268]. Machine widths of 80–250 cm and speeds of 50–200 m/min are
indicated. The most important coating devices include gravure roll, roll blade
systems, and kiss coating with knife over roll. Generally, silicone coating
machines possess a coating, drying, and rehumidification station [269].

7.4 Technology for Solventless Siliconizing


A machine for solventless silicone coating is equipped with fully automatic
unwinders and rewinders, as well as with a rehumidification unit [270]. In
contrast to solvent-based or aqueous systems, nearly ten times or less liquid
mass must be applied onto the substrate (i.e., a quantity of 0.7 g/m2 of paper
620 Chapter 9

normally suffices, generally less than 1 g/m2) [264]. Solvent-free silicone


release coatings vary from 1.1 to 1.4 g/m2 [270]. Today coating quality
tolerances are guaranteed within 5% in longitudinal and cross directions.
The silicone is applied with a five-roll system running at different speeds and
provided with distinct surface structures. Today speeds of up to 400 m/min
with a working width of 2.3 m are reached and speeds of 500 m/min are
conceivable. Usually, thermal systems are used for silicone curing, with
single- or double-sided airfoil dryers, depending on the type of the substrate.
A combination of roll and blade coaters was suggested for 100%
solventless silicone coating systems [271]. Totally solventless systems are
cured with platinum or rhodium catalysts as addition reactions of Si-H groups
and vinyl groups occur. Common solventless silicone release formulations
contain: coating, release modifier, and curing agent. The platinum catalyst is
included in the polymer and the release modifier. For an extended range of
cure and release performance, four-component formulations were developed
[272]. Such recipes contain: polymer, resin, crosslinker, and catalyst.
Siloxane microemulsions also may be used as release agents [273].
A release agent with excellent mechanical stability contains an organo-
polysiloxane microemulsion obtained by emulsion polymerization of an
organopolysiloxane. Water-based silicones have been known since 1986
[273]. In 1988 20% of the silicone release liner production was based on
water-based technology [274], actually the level of water-based siliconizing
attains about 30%. With solventless and emulsion technologies release liners
can be produced at machine speeds of 1000 m/min [272,275]. The
advantages/disadvantages of the solvent-based/aqueous release coatings
are summarized in Table 9.5.

Table 9.5 Advantages/Disadvantages of Solvent- and Water-Based Release


Coatings

Advantages Disadvantages

Solvent-based Anchorage to variety of liners Cost


Absence of migrating species Flammability
Less sensitive to Environmental regulations
humidity variations
Low solids content
Faster drying Clean-up
Energy costs
Water-based Cost Foaming
Nonflammable Poor anchorage to certain liners
Clean-up Migrating species
Nonpolluting Humidity sensitivity
Manufacture of Pressure-Sensitive Labels 621

8 REHUMIDIFICATION / CONDITIONING

Due to the high thermal web strain caused by the generally high crosslinking
temperatures of the silicone, humidity is extracted from the web. Therefore,
this lost humidity must be added to the oven-dried web after the coating or
curing process in order to eliminate the tendency for curling or for flatness
correction purposes. For this purpose the paper is led through one or more
jet steamers where a highly saturated steam/air mixture is generated. Several
papers review rehumidification [269,276–279]. In some cases, in order to
increase the hygroscopy of the web, additives such as carbamide, glycerine,
or calcium nitrate are used [104].

REFERENCES

1. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,


Marcel Dekker, New York, 1999, Chapter 6.
2. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 2.
3. A. Dobman and J. Planje, Papier u. Kunststofjverarb., (1) 37 (1986).
4. Die Herstellung van Haftklebstoffen, T1.-2, 2-14d, Dec. 1979, BASF,
Ludwigshafen, Germany.
5. Adhäsion, (1/2) 27 (1987).
6. Coating, (4) 123 (1988).
7. Paper, Film and Foil Conv., (10) 39 (1971).
8. J. Paris, Papier u. Kunststoffverarb., (9) 53 (1988).
9. Coating, (3) 56 (1984).
10. E. Bohmer, Norsk Skogindustrie, (8) 258 (1968).
11. Coating, (6) 156 (1974).
12. W. Schaezle, Coating, (9) 314 (1987).
13. H. Hadert, Coating, (6) 175 (1969).
14. H. Klein, Adhäsion, (1) 15 (1976).
15. H. Klein, Coating, (11) 406 (1987).
16. J. R. Martin, Paper, Film and Foil Converter, (9) 81 (1989).
17. Papier u. Kunststoffverarb., (11) 25 (1985).
18. Coating, (4) 23 (1969).
19. H. Klein, Coating, (10) 378 (1988).
20. H. Klein, Coating, (1)18 (1986).
21. H. Klein, Coating, (12) 336 (1983).
22. H. Klein, Coating, (3) 60 (1988).
23. H. Klein, Coating, (4) 122 (1987).
24. H. Klein, Coating, (2) 49 (1987).
25. C. Massa, Coating, (7) 239 (1989).
26. G. Renz, Allg. Papier Rundschau, (24–25) 960 (1986).
27. Coating, (7) 240 (1987).
622 Chapter 9

28. G. W. Drechsler, Coating, (7) 251 (1987).


29. Vollständige GID Modul –Einheit mit allem Zubehör, Bematec SA, 1991,
Morges-Lausanne, Switzerland.
30. Der BemaTec-Compact Coater, Bematec SA, 2001, Morges-Lausanne,
Switzerland, 1994.
31. Beschichtungstechnologie, Bematec SA, Morges, Switzerland, 11. 96.
32. H. Hardegger, Coating, (6) 194 (1992).
33. H. Klein, Coating, (2) 32 (1988).
34. D. Percivalle, European Tape and Label Conference, Exxon, Brussels, April
1993, p. 133.
35. Adhes. Age, (11) 28 (1983).
36. H. D. Patermann, Kunststoffberater, (1/2) 39 (1988).
37. Coating, (2) 49 (1988).
38. National Starch Chem. Co., Durotak, Pressure-Sensitive Adhesives, Technical
Bulletin, 4/1986.
39. H. Klein, Coating, (2) 42 (1988).
40. H. Klein, Coating, (10) 372 (1988).
41. H. Baumeister, Coating, (10) 8 (1982).
42. W. Drechsler, Coating, (7) 253 (1987).
43. R. B. Köhler, P. H. Lathrop and D. M. McLeod, Tappi, Polymers, Laminating
and Coating Conference Proc., September 1988, p. 205.
44. Coating, (3) 87 (1969).
45. H. Türk, Papier u. Kunststoffverarb., (10) 22 (1985).
46. P. Herzog, Coating, (3) 73 (1988).
47. F. T. Sanderson, Adhes. Age, (12) 19 (1983).
48. Adhes. Age, (8) 24 (1988).
49. Coating, (2) 48 (1988).
50. F. Renk, Coating, (4) 146 (1988).
51. H. D. Patermann, Coating, (3) 96 (1988).
52. H. D. Patermann, Coating, (6) 224 (1988).
53. T. G. I., Twentse Graveerindustrie, Glanerbrug, Rasterwalzen, Technical
bulletin, 1995.
54. Technical bulletin, Pamarco Inc., Roselle, NJ, 1994.
55. H. Klein, Coating, (7) 225 (1981).
56. F. Weyres, Coating, (4) 110 (1985).
57. M. Benzi, Coating, (4) 120 (1991).
58. T. J. Bonk and S. J. Thomas (Minnesota Mining and Manuf. Co., St. Paul,
MN, USA), EP 120.708.81/03.10.1984.
59. Labels and Labelling, (5) 81 (1999).
60. P. M. Fahrendorf, Paper, Film and Foil Conv., (9) 106 (1989).
61. J. R. Marline, Paper, Film and Foil Conv., (9) 81 (1989).
62. Labels and Labelling, (5) 96 (1999).
63. H. Klein, Coating, (12) 142 (1984).
64. Coating, (12) 365 (1984).
65. Coating, (2) 50 (1978).
Manufacture of Pressure-Sensitive Labels 623

66. Croda Adhesives, Water Based Laminating Adhesives, Newark (U.K.), 1995.
67. Backofen & Meier AG, Bülach, Switzerland, Reverse Gravure Coating of
Emulsion PSA, 1994.
68. Coating, (8) 399 (1987).
69. Adhes. Age, (3) 87 (1969).
70. Coating, (3) 87 (1969).
71. J. Paris, Papier u. Kunststoffverarb., (9) 44 (1990).
72. H. Hardegger, Coating, (5) 153 (1992).
73. Air Products, Airflex, Polymer Chemicals Technical Report, p. 5.
74. Papier u. Kunststoffverarb., (2) 16 (1979).
75. ARC Europe, Dynamic Flow Chamber, Chambered Doctor Blade System,
Technische Information, Vergleich verschiedener Kammerrakelsysteme,
June 1994.
76. B. Voges, Klebrohstoffe, Beschichtungstechnikum, BASF, Ludwigshafen,
1993.
77. Adhes. Age, (7) 36 (1986).
78. H. Bohlman, Papier u. Kunststoffverarb., (3) 12 (1979).
79. Adhäsion, (1/2) 23 (1993).
80. Allg. Papier Rundschau, (1) 18 (1978).
81. G. Kupfer, llth Munich Adhesive and Finishing Seminar, Munich, Germany,
1986, p. 105.
82. H. M. Maschke, Coating, (6) 186 (1970).
83. Coating, (12) 344 (1984).
84. H. Klein, Coating, (11) 390 (1986).
85. Coating, (11) 393 (1986).
86. Coating, (8) 399 (1987).
87. G. Kupfer, Allg. Papier Rundschau, (16) 428 (1986).
88. Coating, (7) 238 (1989).
89. Coating, (7) 222 (1989).
90. R. Wimberger, Coating, (7) 258 (1991).
91. Coating, (7) 258 (1991).
92. R. Dehnen, Der Siebdruck, (32) 36 (1986)
93. A. Tawn, Farbe u. Lack, (5) 416 (1977).
94. J. J. Koch, Der Siebdruck, (3) 32 (1987).
95. H. Dickerhoff, Coating, (6) 187 (1993).
96. H. Klein, Coating, (7) 255 (1991).
97. E. Klas, Coating, (7) 254 (1991).
98. H. Klein, Coating, (3) 24 (1993).
99. I. Benedek, Pressure-Sensitive Formulation, VSP, Scientific Publishers,
Utrecht, 2000, Chapter 4.2.
100. Adhes. Age, (3) 41 (1985).
101. G. Kupfer, Coating, (1) 6 (1984).
102. K. Bond, Adhes Age., (2) 32 (1990).
103. I. Benedek, Pressure-Sensitive Formulation, VSP, Scientific Publishers,
Utrecht, 2000, Chapter 2.
624 Chapter 9

104. Solvay and Cie, Beschichtung van Kunststoffolien mit IXAN, WA, Prospect Br.
1002d-B-0, 3-0979.
105. Coating, (8) 311 (1982).
106. H. Hadert, Coating, (3) 54 (1973).
107. Labels and Labelling, (5) 48 (1999).
108. H. Wittke, Coating, (9) 290 (1986).
109. F. Krizek, Coating, (9) 316 (1987).
110. A Steedman, European Adh. Seal, (9) 6 (1997).
111. G. Menges, E. H. Stroh and D. Weinand, Papier u. Kunststoffverab., (9) 68
(1986).
112. Papier u. Kunststoffverarb., (9) 38 (1986).
113. R. A. Grainwille, Paper, (12) 763 (1973).
114. Der Siebdruck, (3) 240 (1987).
115. W. Barnscheidt, Wochenbl. f. Papier., (2) 63 (1974).
116. Coating, (12) 334 (1985).
117. Kaut., Gummi, Kunstst., (10) 899 (1983).
118. Papier u. Kunststoffverarb., (4) 21 (1986).
119. J. Lawton, Paper, (12) 769 (1973).
120. Coating, (6) 208 (1988).
121. Coating, (6) 211 (1989).
122. Coating, (7) 250 (1988).
123. W. Drechsler, Coating, (5) 164 (1997).
124. T. Frecska, Screenprinting, (6) 68 (1987).
125. Coating, (6) 156 (1974).
126. Converting Today, (8) 13 (1991).
127. B. H. Gregory, ‘‘Extrusion Coating Advances—Resins, Processing,
Applications, Markets,’’ Polyethylene ‘93, The Global Challenge for
Polyethylene in Film, Lamination, Extrusion, Coating Markets, October 4,
1993, Maack Business Service, Zurich, Switzerland.
128. Converting Today, (1) 13 (1982).
129. Neue Verpackung, (7) 78 (1971).
130. H. Münstedt and H. J. Wolter, Kunststoffe, 83, (10), 725 (1993).
131. Neue Verpackung, (8) 64 (1993).
132. D. Djordjevic, ‘‘Tailoring Films by the Coextrusion Casting and Coating
Process,’’ Speciality Plastics Conference ’87, Polyethylene and Copolymer
Resin and Packaging Markets, December 1, 1987, Zürich, Switzerland.
133. E. Haberstroh, Papier und Kunststoff Verarbeiter, (8) 14 (1986)
134. J. Kortwig, 12th Munich Adhesive and Finishing Seminar, 1987, Munich,
Germany, p. 6.
135. M. Le Jeune, Pressure-Sensitive Adhesives and Adhesive Coating, Cowise,
24.04.1996, Amsterdam, Netherlands.
136. T. Vogel, ‘‘Impact Air Pollution Regulations on the Pressure-Sensitive
Industry,’’ PTSC XVII Technical Seminar, 4 May, 1989, Woodfield
Shaumburg, IL., USA.
137. European Adh. Seal, (6) 32 (1995).
Manufacture of Pressure-Sensitive Labels 625

138. Neue Verpackung, (12) 49 (1991).


139. D. Ruchay, Allg. Papier Rundschau, 117 (50/51) 1321 (1993).
140. H. Lübberick, Papier und Kunststoff Verarbeiter, (5) 114 (1986).
141. H. Nitschke, Papier und Kunststoff Verarbeiter, (10) 8 (1987).
142. W. Wittke, Coating, (7) 249 (1987).
143. W. Wittke, Coating, (3) 85 (1987).
144. W. Wittke, Coating, (6) 210 (1987).
145. W. Wittke, Coating, (5) 168 (1987).
146. Coating, (12) 345 (19874).
147. Coating, (7) 180 (1981).
148. Druck-Print, (2) 60 (1989).
149. W. Wittke, Coating, (9) 334 (1987).
150. Coating, (4) 120 (1992).
151. Druckwelt, (12) 44 (1987).
152. D. Eitner, Chemie-Umwelt-Technik, (4) 50 (1992).
153. Duro-Tak, 280-1151; Speciality Adhesives, Product Data National Starch and
Chemicals B. V., Zuthphen, Netherlands, p. 3, 1986.
154. G. Wilmes, Allg. Papier Rundschau, 20, 617 (1987).
155. K. Feldmann, Coating, (6) 204 (1987).
156. Coating, (8) 207 (1984).
157. Der Polygraph, (10) 1158 (1986).
158. Klebstoffvorprodukte, Abwasserbehandlung, EDM/G;September, 1985, BASF,
Ludwigshafen, Germany.
159. H. Maack, ‘‘Coextruded and Coinjected Packaging Developments and
Recycling Aspects,’’ Maack Business Service, Speciality Plastics Conference,
1989, Zurich, Switzerland.
160. Z. Czech, European Adhes. Seal., (6) 4 (1995).
161. Finat Labelling News, (4) 35 (1995).
162. P. Greenway, International Forms, (3) 40 (1997).
163. Neue Verpackung, (8) 63 (1993).
164. G. A. H. Kupfer, Coating, (5) 169 (1993).
165. European Adhes. Seal., (9) 36 (1997).
166. Kleben & Dichten, Adhäsion, (10) 20 (1994).
167. J. B. Morrison, ‘‘The Technology Challenge of Paper Recycling,’’ PSTC, XVII
Technical Seminar, May 4–6, 1994, Woodfield, Shaumburg, IL, USA.
168. J. Y. Peng, ‘‘US Postal Service Environmentally Benign PSA Program,’’ Proc.
24th Annual Meeting of Adh. Soc., Febr. 25–28, 2001, Williamsburg, VA, p. 157.
169. Finat Labelling News, (1) 17 (1995).
170. Coating, (12) 472 (1995).
171. C. W. Paul, R. Billmers, P. Altieri and M. Blumenthal, ‘‘Starch-based Hot-
Melt Adhesives, ’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25–28, 2001,
Williamsburg, VA, p. 325.
172. Labels and Labelling, (5) 52 (1999).
173. European Plastics News, (11) 65 (1993).
174. European Plastics News, (11) 127 (1993).
626 Chapter 9

175. Ecostar Kunststoff-Zusätze Vertriebsgesellschaft, Ochtrup, Germany,


Kunststoff-Information, (1054) 4 (1991).
176. Polyethylene Films, Conversion Industry Reference Report, 1994, Ch. 1.1.2,
Maack Business Service, Zürich, Switzerland.
177. Kunststoffe, 84 (11) 1569 (1994).
178. Pack Report, (3) 15 (1997).
179. Coating, (3) 65 (1974).
180. Papier und Kunststoff Verarbeiter, (10) 40 (1990).
181. Pack Report, (4) 62 (1997).
182. Biocell 163, Tubize Plastics, Rhone Poulenc, Paris, France, 1995.
183. T. Jopski, Kunststoffe, 83 (10) 748 (1993).
184. L. P. Schreiber, Technica, (25/26) 20 (1995).
185. Wacker Chemie GmbH, Munich, Germany, EP 0603746; in European Coat J.,
(12) 964 (1995).
186. D. Cotto and J. M. Haudin, Rev. Ge´n. Therm. France, (298) 879 (1985).
187. R. Hinterwaldner, Coating, (5) 181 (1997).
188. R. Hinterwaldner, Coating, (8) 274 (1997).
189. R. P. Wool and S. Bunker, ‘‘Bio-Based Pressure Sensitive Adhesives,’’ Proc.
24th Annual Meeting of Adh. Soc., Febr. 25–28, 2001, Williamsburg, VA, p. 156.
190. S. P. Bunker and R. P. Wool, ‘‘Pressure Sensitive Adhesives from Renewable
Resources’’, Proc. 24th Annual Meeting of Adh. Soc., Febr. 25–28, 2001,
Williamsburg, VA, p. 174.
191. P. Roll and E. Foil, European Tape and Label Conference, Exxon, Brussels,
April 1989, p. 151.
192. Adhes. Age, (7) 38 (1986).
193. D. J. Clair, Adhes. Age, (23) 30 (1980).
194. K. Nitzl, T. Horna and U. Hoffmann, 16th Munich Adhesive and Finishing
Seminar, Munich, Germany, 1991, p. 100.
195. M. Dupont and M. De Keyzer, 19th Munich Adhesive and Finishing
Seminar, 1994, p. 120.
196. H. Huber and H. Muller, European Adhesives and Sealants, (3) 120 (1987).
197. H. Huber and H. Müller, 13th Munich Adhesive and Finishing Seminar,
Munich, Germany, 1988.
198. Adhäsion, (4) 4 (1985).
199. K. Nitzl, G. Peter, T. Horna and W. Hasselbeck, ‘‘Radiation curable PSA,’’
15th Munich Adhesive and Finishing Seminar, Munich, Germany, 1990.
200. A. Forster, Thesis, FH München, FB 05/1988.
201. H. P. Seng, Coating, (7) 245 (1992).
202. Coating, (2) 50 (1986).
203. R. Seidel, Allg. Papier Rundschau, (16) 432 (1986).
204. M. Schmalz and W. Neumann, European Tape and Label Conference,
Brussels, 1993, Exxon, p. 143.
205. S. O. Hendrichs (Minnesota Mining and Manuf. Co., St. Paul, MN, USA),
US Pat., 2.956.904; cited in ref. [135].
206. Coating, (6) 154 (1983).
Manufacture of Pressure-Sensitive Labels 627

207. A. R. Hurst, Production of EBC Silicones Release Paper and Film, Radcure 82,
Lausanne, Switzerland.
208. Der Polygraph, (5) 366 (1986).
209. J. Paris, Papier u. Kunststoffverarb., (6) 12 (1986).
210. H. Röltgen, Coating, (11) 399 (1986).
211. H. G. Muller, Kunststoff. J., (9) 8 (1986).
212. Coating, (11) 292 (1982).
213. Coating, (12) 332 (1982).
214. F. T. Birk, Coating, (9) 238 (1985).
215. R. Hinterwaldner, Coating, (3) 73 (1985).
216. Coating, (6) 154 (1983).
217. K. F. Roesh, Coating, (12) 353 (1983).
218. K. D. Schroter, Coating, (11) 331 (1984).
219. R. P. Cawthorne, Papier u. Kunststofjverarb., (10) 56 (1985).
220. Coating, (2) 45 (1986).
221. P. P. Zylka, Der Siebdruck, (3) 60 (1986).
222. H. Röltgen, Coating, (12) 331 (1985). .
223. H. Röltgen, Coating, (4) 102 (1985).
224. H. Röltgen, Coating, (5) 124 (1985).
225. T. Birk, Coating, (12) 278 (1985).
226. R. Hinterwaldner, Adhäsion, (3) 14 (1989).
227. G. A. Kupfer, Coating, (12) 258 (1985).
228. R. Hinterwaldner, Adhäsion, (10) 24 (1985).
229. R. Hinterwaldner, Adhäsion, (12) 13 (1985).
230. G. Strohner, Adhäsion, (1/2) 24 (1985).
231. K. Fuhr, Defazet, (6/7) 257 (1977).
232. Tappi J., (9) 27 (1987).
233. Coating, (12) 344 (1984).
234. W. Graebe, Papier und Kunststoffverarb., (2) 49 (1985).
235. Coating, (11) 396 (1987).
236. H. Brus, C. Weitemeyer and J. Jachmann, Finat News, (3) 27 (1987).
237. Labels and Labelling, (5) 44 (1999).
238. UV and EB Curable Silicones, Th. Goldschmidt A. G., Oligomers and Silicone
Division, Essen, Germany, 09.97/2000.
239. M. Fairley, Labels and Labelling, (5) 44 (1999).
240. Labels and Labelling, (1) 20 (1999).
241. R. Holl, Verpackungsrundschau, (3) 220 (1986).
242. K. Nitzl, Adhäsion, (6) 23 (1987).
243. D. A. Meskan, Coating, (7) 238 (1992).
244. R. Kardaghian, Adhes. Age, (4) 24 (1987).
245. Adhäsion, (12) 8 (1988).
246. Coating, (5) 177 (1988).
247. Coating, (4) 121 (1987).
248. V. Lauppi, Coating, (7) 188 (1984).
249. V. Lauppi, Coating, (1) 8 (1984).
628 Chapter 9

250. Chemische Industrie, (8) 64 (1989).


251. Labels and Labelling, (5) 43 (1999).
252. J. Tin, W. Wen and B. Sun, Adhes. Age, (12) 22 (1985).
253. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 4.3.
254. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.1.
255. R. A. Bafford and G. E. Faircloth, Adhes. Age, (12) 234 (1987).
256. A. W. Norman, Adhes. Age, (4) 35 (1974).
257. J. M. Casey, Tappi J., (6) 151 (1988).
258. L. C. Fehrmann, Adhes. Age, (6) 48 (1970).
259. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 6.2.
260. Coating, (11) 396 (1987).
261. R. Hinterwaldner, Coating, (4) 87 (1984).
262. R. Hinterwaldner, Coating, (5) 172 (1987).
263. G. Kupfer, Coating, (3) 25 (1986).
264. H. Klein, Adhäsion, (7) 185 (1976).
265. H. Zander, Adhäsion, (11) 37 (1988).
266. R. Thomas, Allg. Papier Rundschau, (16) 437 (1986).
267. Adhäsion, (6) 6 (1987).
268. R. Thomas, 4th PTS, Adhesive Seminar, Munich, Germany, 1984, p. 33.
269. R. Pagendarm, Coating, (5) 182 (1988).
270. H. U. Hermann, 2nd International Cham Tenero Meeting of RSM Industry,
Locarno, Switzerland, March 29th, 1990, p. 21.
271. Adhes. Age, (7) 36 (1986).
272. Labels and Labelling, (5) 12 (1999).
273. CAS, Emulsion Polymerization, (10) 3 (1988), 108.152233f.
274. Coating, (1) 3 (1988).
275. A. Soldat, A. Fall and G. Lechtenbohmer, 18th Munich Adhesive and
Finishing Seminar, Munich, Germany, 1993, p.145.
276. Coating, (4) 242 (1988).
277. H. Hadert, Coating, (6) 175 (1969).
278. Coating, (19) 277 (1982).
279. E. Mandershausen, Coating, (19) 271 (1974).
10
Test Methods

As in the case of many other finished products a wide range of test methods
were developed for PSA-based labels, tapes, and coatings. A number of
organizations such as FINAT, AFERA, PSTC, and TLMI have established
(standard) test methods which are widely used in the industry, although
there remain significant differences between the various methods used [1].
These methods provide a good basis for the evaluation of adhesives, but
some modifications or additional tests are required when testing materials
for specific applications. Therefore, the principal PSA manufacturers and
converters have developed their own methodology. It is not the aim of this
book to discuss the standardized PSA testing methods in detail. Here it
appears of more interest to describe the special methods worked out for
special end-uses or for specific PSA applications. It is evident that for a label
manufacturer the properties of the liquid adhesive, its coating behavior,
and the PSA label performance characteristics are very important. From the
point of view of methodology, these areas differ quite a bit.

1 EVALUATION OF THE LIQUID ADHESIVE

Pressure-sensitive adhesives are coated in the liquid state (i.e., as a


dispersion, as a solution, or as a molten adhesive). Therefore, the evaluation
of the adhesive starts with the examination of the characteristics of the
liquid PSAs, generally by testing the flow properties. Evidently, the flow
properties of the solvent-based PSAs (stable systems), water-based PSAs
(shear-sensitive systems), and hot-melt PSAs (highly viscous, temperature-
sensitive systems) differ. The properties of the liquid adhesive to be tested
include the solids content, the viscosity, the flow properties, and the free
monomer content [2]. Later, the following properties of the dried PSA coating
should be tested, namely the 90 and 180 peel adhesion, the cohesion (shear),
the tack, the Williams plasticity, the shrinkage (plasticizer influence of

629
630 Chapter 10

the PVC) or dimensional stability behavior, the storage properties, the


temperature resistance, the peel from the release liner (release force), as well
as the aging on the substrate.

1.1 Hot-Melt PSAs


The most important processing parameter of hot-melt PSAs is their viscosity
and thus this parameter should be tested according to common methods.
Not only the viscosity but also its time/temperature dependence (aging)
appear very important. Melt viscosities are often determined in the
temperature interval range of 140–200 C using a Brookfield viscometer
(spindle LVF 4) [3]. In [4] a test with a Brookfield viscosimeter RVT
at 180 C, spindle 7, and 20 rpm is suggested. Recent developments in
the measuring techniques/devices, using cone and plate geometries, allow a
more rapid determination of the viscosity curves at different shear rates and
temperatures requiring a minimal amount of adhesive.

1.2 Solvent-Based PSAs


Generally, the processability of solvent-based PSAs depends on their
viscosity and solids content. Coating devices are designed for a given
viscosity, whereas the coating weight and the drying speed depend on
the solids content. In general, the solids content and viscosity should be
measured for formulated compositions. Because of the nonideal flow
behavior of the adhesive solutions (dispersions), viscosity measurements at
different flow rates should be made. However, the most common industrial
method is the cup-flow method where a one point viscosity measurement
(flow time versus cup orifice) is carried out; Ford cup or DIN cup 3 or 4 are
common [5]. Table 10.l summarizes the most commonly used test methods
for viscosity measurements.

1.3 Water-Based PSAs


Generally, the properties of aqueous dispersions to be tested [6] include the
viscosity and its stability in time, the thinning (diluting) characteristics, the
solids content, and the temperature stability. With regard to the processing
properties of the dispersion, there are several requirements: the absence of
foaming, no drying on the rolls, the wet-out, the mechanical stability when
being pumped, and the lack of odor. The liquid flow properties of water-
based PSAs are crucial for their processability. Aqueous dispersions are
shear-sensitive systems, where mechanical influences during storage,
Test Methods 631

Table 10.1 Test Methods for Viscosity Measurement of Water-Based PSAs

Test Conditions

PSA Supplier Apparatus Type Spindle RPM Temperature ( C)

AKZO Brookfield RVT 3 50 20


Polysar Brookfield RVT 2 30 —
Rhone-Poulenc Brookfield RVT — 50 23
BASF Contraves STV C11 — — 23
Hoechst Brookfield RVT 5 20 23
Hoechst Contraves MS-C11 — — 23
Rohm & Haas Brookfield LVT 3 30 —
Sealock Brookfield RV 3 20 20

handling, and coating operations may cause the formation of grit or foam;
therefore the examination of the mechanical stability remains very
important. On the other hand, water-based adhesives are difficult to coat
because of their low viscosity and high surface tension, and because of the
low surface tension of the release liner or face stock web. The density and
particle size of the dispersion may influence its stability, the coating proper-
ties, and drying ability as well. Foam properties determine the processing
speed and coating quality appearance.
The particle size and particle size distribution influence the viscosity,
surface tension, and mechanical stability of the dispersion. Mechanical
stability influences the viscosity; viscosity and surface tension determine the
coatability. For the PSA conversion, coatability remains the main property
of the liquid adhesive because the wettability also leads to the choice of the
adequate coating geometry, of the face stock, and of the release liner.

Coating Properties
Wettability, the desired coating weight under required conditions (running
speed, temperature), and optical appearance are factors of the coatability
behavior. The viscosity and surface tension influence the wettability: they
should be tested a priori as should the viscosity (thinning, diluting response)
and mechanical stability.
The methods for measuring the viscosity were discussed earlier
(see Section 1.1). For water-based PSAs the viscosity (Brookfield RVT,
spindle 4, 20 rpm) at room temperature should be tested immediately after
compounding [7]; in some cases the thixotropy index also is measured.
The thixotropy index is the ratio of the Brookfield viscosity at 0.5 rpm to
632 Chapter 10

the Brookfield viscosity at 50 rpm. This index shows an increase of the low
speed viscosity of 5% for conventional polyacrylic PSAs used as thickener,
as compared to 33% for some special products.

Wetting Characteristics
The wetting characteristics of a dispersion are examined either directly or
indirectly. The direct evaluation of the wetting characteristics implies the
subjective examination of the wet-out of the original or slightly diluted
dispersion on a release liner or on a face stock material. The indirect
characterization of the wetting properties refers to the measurement of
the surface tension of the dispersion or of the wetting angle.

Direct Examination of the Wetting Out. The evaluation of the wet-


out should be carried out on the surface of the carrier material which is to be
coated. Depending on the nature of the coating (direct/transfer), either the
face stock or the release liner should be used as the test surface. The test of
the wet-out on a face stock material is mostly used for film carriers.
The liquid typical PSA first is applied onto a release liner before
transfer application onto the face stock; both solvent-based and solventless
silicone liners should be included in the evaluation. After drying a com-
parison of the coatings should be made, based on the visual inspection of the
presence (or absence) of defects (cratering, pinholes, fish eyes, etc.) [8]. The
wetting out can be characterized by crater number and wetting defects [9].

Vinyl Wetting Test. A film of the experimental latex is cast on the


PVC sheet using the draw-down bar. The film coating is inspected for fish
eyes and cratering. The fewer fish eyes and craters the film contains the
better the vinyl wetting of the latex.

Indirect Examination of the Wet-Out. Wetting out can be tested via


sur-face energy and surface tension. The evaluation of the wetting out and
the existing norms have been described [10,11]. The evaluation of the surface
tension according to Union Carbide or ASTM yields no real information
about the adhesion [12]. This test is made more in order to characterize the
coating performance of an adhesive or of a surface to be coated. The coating
performance and wetting out depend on the surface tension, but the surface
tension measurement alone is not sufficient to characterize the actual
behavior of an adhesive on a coating machine. In the coating of a latex
at high speeds the dynamic aspect of wetting also appears critical. The
measurement of the contact angle of a latex on a face stock material or
the surface tension of the latex with different surfactant systems determines
the choice of the surfactant [13]. The wettability of adhesives also may be
Test Methods 633

examined using a wettability balance [14]. Pitzler [15] stated that under
certain conditions (low surfactant level) the results of the ring and plate
method for surface tension measurement can differ. Pitzler is proposing
dynamical contact angle measurements with advancing and receding angle
measurements, and stated that for wetting out the advancing contact angle
is more important. It should be emphasized that this depends on the coating
speed. Independent from the analysis of the static wet-out, dynamic tests
under shear also are necessary. Surface tension tests were standardized
according to the FINAT methods. Similar test methods were published by
DIN and ASTM. Surface tension measurements and (wet) film-weighing
machine measurements allow the evaluation of adequate surfactants, their
diffusion velocity, the reduction of the surface tension, and the surfactant
concentration required [16].
The evaluation of the surface tension, with the aid of a wettability/
wetting angle measuring device has been described [17]. A simple method
to test the dynamic surface tension is the use of a stalagmometer (capillary
tube). This method—the pendant drop method [18]—involves the measure-
ment of the surface tension of the aqueous dispersion (as compared to
water) using a stalagmometer to determine the number of liquid drops.
The dynamic surface tension is given by the ratio of the number of water
droplets (multiplied by the surface tension of water) to the number of the
emulsion droplets.

Wetting Angle. The method of the wetting angle may be used to test
the face stock, the release liner, or the PSAs. The same method applies in
order to determine the nature (solvent-based, water-based, or hot-melt) of an
unknown PSA (or label). In this case it is assumed that the water-based
adhesive always leaves traces of surface agents on the release liner.
Therefore, a delaminated release liner which had been originally coated
with a water-based adhesive allows a better wet-out for pure water. In order
to determine the water-based or solvent-based nature of an unknown PSA
sample, the wetting angle of water on the delaminated release liner should
be measured (deionized water should be used). For example, typical wetting
angle data for a solvent-based rubber-resin and water-based tackified acrylic
PSA coated release liner are 101–103 and 91–93 , respectively.

Wettability of a Release Liner. In the evaluation of the wettability of


an adhesive on a release liner care should be taken because of the change of
the wettability of the siliconized paper with time (Table 10.2). Following a
small increase after a few days, it decreases with increasing storage time (i.e.,
the wettability increases with increasing age of the siliconized paper).
634 Chapter 10

Table 10.2 Wettability of Siliconized Paper as a Function of Age

Characteristics of Silicone Coating (wetting angle, )

Nature of the Silicone Coating

Age of the coating (days) Solventless Solvent-based

0 103 109
103 107
100 113
96 109
1 108 87
107 89
103 106
103 105
2 108 107
106 108
110 110
107 111
3 106 111
104 108
105 111
101 110
7 106 104
108 105
102 105
103 106
30 98 106
98 106
100 107
98 106

Wetting Angle for Adhesive Extract. Extracting with water the water-
soluble surface-active agents from a PSA-coated laminate (adhesive nature
unknown) and comparing the wetting angle of the extract with those of a
known hot-melt or solvent-based PSA, it is possible to identify the nature of
the adhesive (Table 10.3).

Foam Content
The most important tests used for the characterization of the foaming ability
of a dispersion consist of the following:
1. Stirring and measuring the foam density.
Test Methods 635

Table 10.3 Test of the Water-Based Nature of PSAs. Wetting Angle for Aqueous
Laminate Extracts

Surface-Active Agent Level Wetting


Adhesive Nature in the Recipe (%) Angle ( )

Tackified water-based acrylic 1.5 62.0


Solvent-based rubber/resin 0 94.9
Solvent-based rubber/resin 0 84.8
Tackified water-based CSBR acrylic 0.8 77.0
Water-based acrylic 5 47.0

2. High speed stirring (42,000 rpm) and measuring the foam height.
3. Foam generation by air bubbling.
4. Foam dynamics as a function of time.
Generally, tackified dispersions display a higher foam content than untacki-
fied ones. The evaluation of the foaming behavior is carried out using
different stirring devices and measuring the volume, height, or weight of the
foam column [19]. The foam number after 5 min is used for foam charac-
terization [20]. Byk [21] showed that the best results are obtained by stirring
the diluted dispersion for 1 min with a high speed mixer. A volume of 50 ml
of the stirred, foamed, aqueous dispersion is weighed. The weight of the
liquid (dispersion/air mixture) is an index for the foaming characteristics.
Surfactant solutions (or water-based formulations) should be subjected to
high speed agitation at 25–60 C for 3 min in a blender, and the foam volume
should be measured [22].
Procedure for Testing the Foam Height. According to this method
foaming is characterized by the volume of the foamed liquid [23]. The test is
carried out as follows:
The emulsions are diluted to a solids concentration of 7%; 100 g
of the diluted latex is poured into a blender.
The content is stirred for 30 sec at 8000 rpm and subsequently for
150 sec at 15,000 rpm.
The generated foam and remaining liquid are poured into a
1000-ml graduated beaker. The volume of foam and liquid is
reported as foam height (ml).
Another method is:
First, a well-defined sample volume is poured into a 1000-ml
graduated cylinder.
636 Chapter 10

The dispersion then is stirred at 4500 rpm for 30 sec. The volume is
recorded and a 5-min time period is started.
Volumes should be recorded at the 15 sec mark and then each
minute thereafter.
With the aid of the weight/volume correlation, density values can be
calculated. With this method the sample contains both entrained
air and froth foam, and the lowest density value corresponds to the
case of the highest amount of foam. Errors related to the transfer of
the foam into another vessel are eliminated.

Mechanical Stability (Sieve Residue)


The term mechanical stability refers to the ability of the emulsion polymer
to resist coagulation under the influence of shearing stresses as encountered
during agitation or pumping. Shear under machine conditions depends on
the running speed and viscosity, and the shear characteristics of the coating
device. In the most frequently employed test the polymer is stirred at high
speed in a blender for a specific time (typically 10 min). The amount of
coagulum (instability) resulting from this treatment is determined by
filtering the latex through an 80-mesh screen [24]; a 100-mesh screen also is
sometimes used [25]. There are standard test methods for determining grit,
lumps, or undissolved matter in water-borne adhesives [26]. For instance
ISO 705 (third edition, 1985-08-15) describes the determination of coagulum
content (sieve residue) for rubber latex [27].
A rapid measure of mechanical stability is obtained by filtering a
sample through a 100-mesh screen, then stirring it at high speed in a blender
for 5–10 min. The emulsion is again passed through the screen and washed
through with water until only solid, if any, remains on the wire. If there is no
solid residue the emulsion is considered to possess very satisfactory mecha-
nical stability. If so desired the amount of solid can be weighed after drying
the screen in a circulating draft of a vacuum oven at 80–100 C.
The Hamilton Beach blender test is another way of testing the
mechanical stability. The percent residue created by subjecting the emulsion
to high speed agitation for a set period of time is a measure of the stability
of the emulsion (or the lack thereof). According to [34] the Warring Blender
test is not useful in assessing high shear, due to the lack of confined
geometry to produce high shear field. Earlier rheometers normally lacked
the capability to remove viscous heating during the high shearing which
could produce shear-induced flocculation, instead of coagulation. Better
results are obtained using a high shear rheometer (e.g., Hake Rheostress
300) testing the viscosity stability for 600 sec at 70 K/sec. Such laboratory
Test Methods 637

methodology for high shear stability of acrylic emulsions was confirmed by


pumping with a positive displacement progressive cavity pump.

Sieve Residue (Tackifying Resin). This is a measure of how much


deposit the user can expect to find in the filters. Normally, the specification
allows a maximum of 0.05%; a more realistic figure, however, is a maximum
of 0.015%.

Grit Content. The grit represents the coagulum which did not pass
through a 200-mesh screen [30]. It can be tested visually, either by coating
a thin (2–5 g/m2) adhesive layer on a transparent face stock material or by
using a screen (gravimetrically).

Drying Ability and Conditions


While screening different water-based PSAs with similar adhesive but
different processing conditions it appears very important to estimate
their drying ability. The drying ability is the drying speed (loss of water
with time) under real conditions. Unfortunately, there is no standard
method nor apparatus for this test. Preliminary screening tests may
be carried out by the weight loss determination of the coating in time
(see Fig. 6.5).
Tests with PSA laminates require a dry adhesive coating. Different
companies use quite different drying conditions as can be seen from
Table 10.4. ASTM has developed methods of determining ‘‘drying, curing,
or film forming of an organic coating at room temperatures’’ (TE5-
1740-65T).

Table 10.4 Drying Conditions

Coating Characteristics Drying Conditions

Coating Temperature Time


Face Stock weight (g/m2) Nature ( C) (min) Ref.

PET — Solvent-based 90 45 [28]


— — — 100 5–10 [29]
— — Water-based 120 10 [30]
PET — Water-based 130 3 —
PP 25 Water-based 220 0.5 [25]
— — — 93/204 — [31]
Paper — — 70 5 [32]
PET — Solvent-based RT/90 15/5 [33]
638 Chapter 10

There are some properties which allow a screening of water-based


PSAs from the point of view of the coating properties. Parameters such as
solids content, particle size, and density yield indirect information about the
machine properties of the adhesive.

Particle Size
Different methods used for the determination of the particle size and
distribution are known; either simple or more sophisticated pieces of
equipment are used. Particle size is related to mechanical stability,
water resistance, viscosity, and surface tension. The measurement of the
particle size can be carried out with a stalagmometer, by determining the
number of liquid drops for the pure dispersion and the modified one, upon
adding an increased surfactant level to the dispersion. The method is:
The dispersion to be tested should be diluted (1 : 1.5) with water.

Increasing quantities of surfactant (in fact a 10% wet weight


surfactant solution) in ortions of 1–2 cc should be added and
the surface tension should be measured. The surface tension/
concentration of surfactant plot for the unknown dispersion is to
be compared with the plot of a known dispersion (using the same
surfactant). The slope of the surface tension versus surfactant
concentration plot is equal to the moles of soap absorbed per gram
of polymer. If the same surfactant is used as in the synthesis of the
emulsion, and the original amount of soap per gram of polymer
solids is known, the surface area per gram of polymer can be
calculated, and hence the average surface radius of the polymer
particles.

For emulsions, the particle size is a measure of the quality of the


dispersion. Particles with a diameter of 0.25–0.5 m can be observed with
a microscope. Other methods for measuring particle size and particle size
distribution exist (i.e., laser light scattering).

Solids Content
When measuring the solids content of resin emulsions it is very important
that the sample remains in the oven for only 40 min, as the base used for
emulsification is volatile. Karl Fischer water content measurements tend to
indicate a solids content of about 1% higher than solids content determined
in an oven.
Test Methods 639

Density
This is a measure of how much air is contained in the emulsion. A low
density (i.e., high air entrainment) causes problems through dehydration of
the emulsion at the surface (especially if the temperature exceeds 30 C).

2 EVALUATION OF THE SOLID ADHESIVE

Many different test methods were established in order to characterize the


adhesive, converting, and special properties of the solid (i.e., dried) PSAs.
As these depend on the coating weight, the measurement of this parameter
is discussed next. The adhesive properties of the solid PSAs are discussed
in Chap. 6.

2.1 Test of Coating Weight


A simple gravimetric method can help to keep coating weight specifications
within 0.5 g/m2, resulting in considerable savings. The coating weight is
tested by dissolving the adhesive layer in a solvent and determining the
weight difference between coated and uncoated face material.
Apparatus:
Balance with a precision of 0.001 g.
Die-cutting device for 100-cm samples.
IR lamp (300 W).
Method:
Prepare a 100-cm2 sample by die-cutting the laminate and weigh it
with a 0.001-g precision.
Wash off the adhesive layer with a common solvent for the
adhesive (e.g., toluene).
Dry the cleaned face material under the IR lamp for 20 sec at 20 cm
distance from the lamp.
Check the weight of the cleaned, dried face material again.
The weight difference indicates the coating weight.
Generally, for paper/PET coatings, this method yields more reliable results
than comparing the weight of the coated and uncoated face material.
In-line gravimetric measurements of the coating weight can be carried
out by inserting siliconized paper or aluminum foils in the laminating
640 Chapter 10

station. Methods based on beam penetration and absorption (IR, -rays,


etc.) also are often used in-line.

2.2 Other Properties


Other properties of solid (dry) PSAs or raw materials to be examined consist
of the softening point of the resin, the Williams plasticity/modulus, the
nature of the adhesive, the FDA/BGA compliance, the tensile strength, and
the cold flow.

Softening Point of the Resin


Tackified dispersions differ through their softening (melting) point. The
tack/shear balance of an adhesive is related to the softening point of the
tackifier resin (see Chap. 6). Softening points as measured by the Mettler
device are 5–10 C higher than ring-and-ball softening points. Other,
subjective methods for the examination of the softening behavior of a
coating also have been developed; an example is given next.
The softening point of the adhesive coating is determined subjectively
(adhesive transfer, cohesive break) by peeling off (90 ) the laminate affixed
on a heated substrate (test-bank) in different temperature ranges:
Apparatus: Koffler heating stand
Method:
A 1  30-cm adhesive strip should be affixed to the heated stand.
The strip should be peeled off instantaneously at a 90 angle.
The temperature of the heated zone giving adhesive transfer should
be noted.
The softening point provides information about the nature of the adhesive
(hot-melt/solvent-based, competitive samples) and the formulation (tackifier
amount and nature). The softening range can be determined according to
ASTM E-2867 too (4).

Williams Plasticity
The Williams plasticity number indicates the deformability of the adhesive
mass under a static load. Williams plasticity has recently regained favor
for PSA testing purposes. It was used extensively in the 1950s for following
molecular weight changes in reactions during polymer production, but
was replaced by the Mooney viscosity which is normally faster and more
convenient to measure [35].
Test Methods 641

Table 10.5 Screening Values for Williams Plasticity Number and Interdependence
of the Williams Plasticity/Modulus

Rheological Characteristics

Williams Storage Loss


Plasticity modulus modulus
Number ( l02) ( 105) ( 105) Tan  Adhesive Nature Ref.

1.9 1.4 0.7 0.51 — [39]


5.1 — — — Solvent-based acrylic [37]
3.2 — — — Solvent-based acrylic [37]
5.3 — — — Solvent-based acrylic [37]
2.2 1.6 0.82 0.46 Water-based acrylic [39]
2.2 1.7 0.82 0.48 Water-based acrylic [39]
1.9 1.4 0.7 0.50 Water-based acrylic [39]
1.9 1.4 0.6 0.43 Water-based acrylic [39]
2.4 2.1 0.9 0.44 Water-based acrylic [39]
1.6 1.7 1.0 0.59 Water-based acrylic [39]
1.0 — — — Hot-melt acrylic [40]

The Williams plasticity is used for screening hot-melt PSAs [36]; it is


run at 100 F for 15 min [37]. The sample coated on a release liner is dried for
24 h, weighed (to 2.0 g), conditioned for 15 min at test temperature, the load
applied, and the plasticity measured 10 min after the application of the load
[38]. In some cases the Williams plasticity is evaluated at different tempera-
tures (38–200 F). Some screening values for the Williams plasticity number
are summarized in Table 10.5.
In this test method the thickness of the sample (in mm) is measured
while subject to a constant load (5 kg) at 100 F (38–37.8 C) [41].
Apparatus:
Plastometer with gauge (Williams Plastometer, Scott Testers Inc.)
in accordance with ASTM-D-926.
An oven.
Release paper.
Method:
A wet film of the adhesive is coated onto siliconized release
paper, so as to produce a dry film. It is dried at room temperature
and further dried for 5 min at 135 C in a circulating air oven. The
adhesive is removed from the silicone paper, and a pellet (about
2 g) is formed in the shape of a ball.
642 Chapter 10

Condition the ball-shaped sample (placed between two silicone


papers) for 20 min at 38 C.
With the plastometer and sample at 38 C, place sample (between
two pieces of release paper) into plastometer, and record the
thickness of sample after 14 min.
The Williams plasticity number (PN) is the thickness of the pellet
in mm after 14 min compression at 37.8 C in the plastometer under
a 5-kg load.
This method yields valuable information about dried (coated) materials.
Some companies use this method; others consider it inadequate for
dispersions.

Cold Flow
In this test the dimensional change of an adhesive sample after storage under
load during a long time (1 week) is measured [42].
Apparatus:
Glass slides.
A 1000-g weight.
Method:
Prepare a 0.3-g spherical sample of the adhesive.
Place the sample to be tested between two glass slides with a 1000-g
weight on top of the glass sandwich.
After 1 week at room temperature the diameter of the squeezed
adhesive disk is measured (cm).
Calculate the ratio of the diameters of the experimental and
standard adhesive. The cold flow resistance is equal to the diameter
of the squeezed disk of the experimental adhesive divided by the
diameter of the squeezed disk of the standard adhesive. Cold flow
also may be evaluated using a load of 3 psi at a temperature of
120 F [43].

Tensile Strength
The method ASTM D-638 is used for the evaluation of adhesive films. For
this purpose, 3.2  0.4-mm thick and 19  0.5-mm wide dumbbell-shaped
specimens are prepared and strained in a tensile tester until break occurs.
The tensile force at break is registered and divided by the original cross-
section area of the narrow section of the specimen.
Test Methods 643

Apparatus: Tensile tester


Method:
A 500-m adhesive layer is coated onto a siliconized release liner.
After drying (at 70 C for 30 min) the adhesive layer is converted
into a test specimen. The adhesive sample should be conditioned
for 24 hr at 21 C and 55% relative humidity (RH). The sample
should be strained at a rate of 300 mm/min in a tensile tester until
break occurs.
Although the data concerning the dependence of the tensile strength
value on the coating weight are contradictory [44–46] a good correlation was
found between tensile strength and peel values for styrene block copolymers
and HMPSAs based on such copolymers [47]. The test of the tensile strength
at break yields unrealistic values (too high) for the cohesion; the tensile
strength at 300% elongation provides data in good agreement with
measured shear values. Tensile tests were used to study strain hardening
observed during fibrillation. The strain hardening observed in the fibrillation
part of the probe test curves can be correlated to tensile tests [48]. By the
analysis of PSA fracture mechanics in the course of peeling Feldstein
correlated peel force to tensile strain and stress  [49]. Generally axis-
symmetric adhesion tests are based on tensile measurement.

FDA and BGA Compliance


Generally, PSAs should comply with the following FDA and BGA
regulations:

FDA 21 CFR* 175 105 (adhesives)


176 170 (aqueous food)
176 180 (dry food)
178 3400 (emulsifier)
BGA Proposal XIV for polymer dispersions, 01.08.85 170.
Mitteilung Bundesgesundheitsblatt 28, 305 (1985).
Prüfung auf Physiologische Unbedenklichkeit.

*Code of Federal Regulations.

3 LAMINATE PROPERTIES

Most properties of PSAs may be examined when coated (i.e., in the


PSA laminate). All PSA samples should be examined for their main
644 Chapter 10

characteristics; those manufactured for special purposes should also be


examined for their special application related properties.

3.1 General Laminate Properties


Independent from their end-use, all PSA laminates possess an inherent
degree of tack, peel, and shear (i.e., an adhesion/cohesion balance as well
as some converting properties). The coating weight influences all these
properties (see Chap. 6). On the other hand, the adhesive and converting
properties depend on the aging response of the adhesive.

Adhesive Properties
The first step in the evaluation process of the adhesive (or other) properties
involves the preparation of the specimen. Generally, the PSA label specimen
to be tested is manufactured in the laboratory when tests concern the PSA,
the face stock material, or the release liner (i.e., when screening the laminate
components).
Quality control tests are carried out on specimens taken from the
industrial production. Generally, laboratory samples are prepared via
laboratory (direct/indirect) coating, drying, and conditioning of the coated
sample. However, the laminate components, the equipment, and the drying
and conditioning parameters used by different PSA suppliers or converters
can vary quite a bit.
Test methods have been established in the United States (ASTM) and
in Germany (DIN). European Norms (EN) also have been defined. Several
adhesive manufacturing associations have developed test procedures that
are internationally accepted and are being used in the trade as reference
methods. The test procedures for PSAs generally examine three key aspects:
the tack, the peel adhesion at 90 and 180 , and the shear resistance
(cohesion). Procedures for measuring adhesion of self-adhesive materials
to a test surface normally rely on a peel test [50]. Such test procedures were
published by a number of organizations such as FINAT, AFERA, and
PSTC. They also have been incorporated in buying specifications of the
German military procurement office (BWB) [51].
Standard test methods for adhesives were described by MacDonald
[52] and Symietz [53]. Test methods for PSAs are derived from those used
originally for gummed papers [54]. Shear, tensile, and peeling tests
were conducted on gummed papers in order to obtain the performance
characteristics with respect to the three classical stress modes. The basic
types of stress to which a joint may be subjected fall into three categories:
tensile, shear, and peel [55]. For the evaluation of the adhesive properties,
Test Methods 645

reference materials like tapes were used in some cases [56]. Thus as a
reference ordinary office-grade Scotch tape was selected; it has a rolling
ball tack of about 180 mm whereas rolling ball tack data of masking tape
typically are 75–130 mm. On the other hand, it is difficult to compare the
shear strength of an ordinary adhesive tape because the backing of this
commercial product will not support the test weight [57]. However, as a
reference value the masking tape shear resistance is 1–17 hr under PSTC-7
conditions.
Preparation of the Specimen. The preparation of the specimen
includes the coating of the adhesive onto a solid state material, the drying
of the coated material, and in some cases laminating it on the release liner/
face stock material. The preparation of the specimen also includes in most
cases the conditioning of the samples.
A layer of liquid adhesive (or molten in the case of hot-melt PSAs)
thick enough so as to produce 20 g/m2 of dry adhesive film is spread on
a Mylar film by means of an adjustable adhesive film applicator and
subsequently dried. A sheet of silicone paper is applied on the dry (cold)
adhesive film. Test specimens are cut before removing the silicone paper. In
some cases the indirect or transfer coating method is used. For hot-melt
PSAs coatings are sometimes made from solutions. Generally, different
kinds of laboratory equipment are used in order to coat the adhesive. Hand
operated devices such as the Bird, Erichsen, Drage, Biddle applicators, etc.,
may be used [19,57].
Face Stock and Substrate Used. A series of performance tests can be
carried out on experimental adhesives using essentially the end-use
substrates. For reasons of comparison most tests are made on standard
plates. The largest single factor in the measurement of bond strength is the
test plate. However, normalized test procedures differ in their definition of the
surfaces required. FINAT uses float process plate glass, AFERA suggests
brushed steel, while PSTC and BWB propose polished steel. FINAT specifies
glass as the most suitable material; the surface smoothness is well defined
and the chemical nature of the material is stable and constant [50]. Other
parameters also influence the results (e.g., drying of the test plate after
cleaning). Different procedures and surface agents will leave different deposits
on the test plate, which may influence adhesion of the self-adhesive material
to the plate.
Drying Conditions. Drying conditions influence the content of liquid
components in the adhesive layer, the composite structure (coalescence), and
the roughness of the adhesive layer. They can also affect the mechanical
properties of the solid state laminate components. In most cases different
646 Chapter 10

coating weights, different test surfaces, and different drying conditions


are used.
Performance data are determined from laboratory prepared test
laminates composed of 1 mil dry adhesive coated on 2-mil polyester protected
with siliconized release paper [37,40]. Different kind of adhesives (thermo-
plastic, thermosetting, self-crosslinking, and emulsion PSAs) are dried at
different temperatures and oven times. Hence, an exposure to ambient air for
15 min and 2 min oven time at 200–250 C are suggested for solvent-based
adhesives; a 1 min ambient air exposure followed by 3 min oven (212 F)
exposure is suggested for emulsion polymers. A tackified CSBR coated on
paper is tested by coating it on release paper, drying for 5 min at 70 C, then
2 min at 100 C, and then transfer-coating onto paper (1  0.1 mil dry).
Samples on PET are prepared by directly coating adhesives onto chemically
treated films and drying for 5 min at 158 F and 2 min at 212 F. The adhesive
then is laminated onto a release liner and conditioned for a minimum of 2 hr
at standard temperature and humidity before testing [38]. Drying at 90 C for
3 min followed by conditioning at 23  2 C and 50% RH for 16 hr also is used
in practice. The adhesive is coated onto the face stock material with a Meyer
bar so as to yield a 20–25 g/m2 dry adhesive after 2 min in a ventilated oven
at 105 C [58,59]. Then a protective liner is placed on the adhesive film and
conditioned at 20 C and 50% RH before testing. There are quite different
recommendations concerning the coating/drying conditions of laboratory
samples (Table 10.6).
Sample preparation is carried out as follows [24]: a wire rod or
wire wound Meyer rod is used to apply 25 g/m2 (dry weight) of the adhesive

Table 10.6 Coating-Drying Conditions of Laboratory Samples of Water-Based


Coatings

Drying Steps

1st Step 2nd Step

Temperature ( C) Time (min) Temperature ( C) Time (min) Ref.

1 70 5.0 100 2.0 [32]


2 115 3.0 23 1440.0 —
3 RT 30.0 105 5–10.0 —
4 220 0.5 — — [25]
5 105 2.0 — — —
6 70 10.0 — — —
7 RT 15.0 90 5.0 [33]
Test Methods 647

Table 10.7 Test Conditions

Air Temperature ( C) Relative Humidity (%) Standard

1 23 þ 1 50  3 DIN/ISO
2 23 þ 2 50  6 BWB-TL
3 23 þ 2 50  5 FINAT
4 23  1 50  5 AFERA
5 23 þ 2 50 þ 2 PSTC
6 23  1 50  2 TLMI

onto a 25-m polyester film. The coating then is dried for 3 min at
115 C and conditioned for 24 hr at 23 C and 65% RH before being
tested. Coatings on polyethylene should be dried at 60 C on soft PVC at 45–
50 C.

Test Conditions/Conditioning. A specification for materials testing


always has to start with a definition of environmental conditions. Methods
will normally specify 23 C at 50% RH, but there are considerable
differences in accepted tolerances (Table 10.7) [60–62]. If the laminate to
be tested contains paper, broad tolerances in the testing environment will
raise considerable problems. One should therefore aim at achieving
environmental conditions according to DIN ISO 137 (1982) [63].
Some test equipment allows the measurement not only at a constant
peel rate but also with a constant force. From the point of view of appli-
cation technology, this mode of measurement is often of greater interest.
The PSA-coated films should be conditioned under different conditions [64].
Fresh films should be stored for 24 hr at 23 C and 50% RH with no contact
with the release paper. Aged films should be conditioned at 40 C for 14 days
and then for 24 hr at 23 C. For water-based PSA formulations the
coalescence (film forming) occurs slowly as compared to hot-melt PSAs
and solvent-based PSAs (interpenetration of resin and dispersion). Practical
end-use conditions may strongly differ from test conditions, in these cases
tests under actual application conditions are required. In [65] the application
climate for the main pressure-sensitive products (labels, tapes, and
protective films) is described.

Measuring the Tack


Tack as an adhesive characteristic was discussed in detail in Chap. 6,
Section 1.1. Generally, the test procedure to measure tack consists of two
steps: bond formation and bond separation. Some of the test methods (quick
648 Chapter 10

Table 10.8 Test Methods for Tack Measurement

Principle of the Method

Name/
Stress nature Procedure variants Norm Unit

1. High speed, Slowdown of Rolling ball PSTC 6 in., cm


repeated a rolling Rolling in., cm
dynamical item cylinder
bonding/
debonding
2. Low speed, Debonding: Loop tack PSTC 5 g/in.2
dynamical via peel, Quick stick TLMI L-IB1 lb/in., oz/in2
bonding/ tensile forces TLMI L-IB2 oz/in., N/25 mm
debonding FTM 9 N/in., N/25 mm
Debonding: gr N, gr/cm2
via tensile Polyken tack
forces Probe tack

stick or loop tack and Polyken tack) simulate the real bond formation
conditions; others, like rolling ball tack, measure the tack under quite
different experimental conditions. Table 10.8 summarizes the different test
methods for the characterization of the tack as suggested by suppliers
of PSAs or raw materials for PSAs. The first and least sophisticated test
for the tack is to put the thumb on the adhesive that was coated onto a
sheet of paper. The next level of complexity involves sticking a loop of
adhesive-coated paper to different substrates. Each time the loop is pulled
from the surface to which it sticks, the tester forms an opinion about how
sticky the adhesive is [66]. The normalized test methods are used to compare
different adhesives [67]. The data are not absolute values, but may be used to
compare the various products.

Quick Stick. Quick stick (PSTC 5) measures the instantaneous


adhesion of a loop of adhesive-coated material using no external pressure to
secure contact. Because of the special loop form of the specimen quick stick
also is called loop tack. According to another definition the quick stick tack
value is the force required to separate at a specific rate a loop of material
which was brought into contact with a standard surface (FINAT test
method 9 or FINAT FTM 9). The FINAT FTM 9 quick stick method
differs from AFERA (4015) and the PSTC 5 quick stick method as the latter
measures peel at 90 without making a loop. The quick stick method is
Test Methods 649

relatively simple and may be carried out using common tensile strength test
machines [68]. Unfortunately, the contact time is long and depends on the
area, the contact area differs, and when using the FINAT method the
peeling angle is not constant. Quick stick has the advantage of allowing tack
to be measured on a wide range of substrates such as stainless steel, glass,
polyethylene, and paper. It can, however, result in high tack readings on
smooth substrates such as glass for adhesives with a low finger tack on
rough surfaces. The EPSMA method for double-sided tapes varies in that it
specifies that a loop of PET film is brought into contact with the tape on a
panel; this procedure discounts the effect of the material rigidity.

Loop Tack. Loop tack is a measure of the force required to remove a


standard adhesive-coated film loop from a standard stainless steel plate after
a short contact of the test strip with the steel plate in the absence of pressure
[41,42]. A 0.5  4-in. strip of one mil Mylar film coated with the sample
adhesive is formed into a loop with the adhesive on the outside; the loop is
applied to the test plate until the PSA loop contacts 0.5 in.2 of the surface
area on the plate. The loop is then removed from the plate at a rate of 12 in./
min and the loop tack is the maximum force observed. According to another
method, a loop formed from a 2.54-cm wide strip of coated paper is lowered
by a tensile testing machine onto a glass plate. The force required to remove
the tape measured at a rate of 300 mm/min at 20 C is the measured loop
tack. Tack (measured as loop tack) depends on the substrate; the following
values were obtained for different substrates (in N/in.) [20]:

Stainless steel 2.8


Glass 3.5
Polyethylene 2.7
PVC 2.5

In a modified PSTC 5 loop tack (quick stick) test polished stainless


steel is used as substrate [59]. When a contact surface of about 1 in.2 is
achieved, the loop is separated from the panel at an 8 in./min separation
speed. Quick stick or loop tack is the average of five test values. In a further
modification (modified PSTC 5) a 200 mm/min rate may be applied. In
some cases a 2.54  12.7-cm adhesive-coated strip also is used. Tack
should be tested in N/20 mm on chromed steel plate [41]. Quick stick to
stainless steel and to Kraft paper also are measured [43]; peel build-up and
re-stick also are tested. Loop tack also can be tested through the test method
ATM 136-84 [20]. Table 10.9 summarizes the test methods concerning
loop tack.
650 Chapter 10

Table 10.9 Loop Tack Test Methods

Test Conditions

Face material Substrate Contact surface (cm2) Speed (mm/min) Ref.

PET (50 m) Stainless steel 2.54  1.27 — [41]


PET Stainless steel — 200 [59]
PET Stainless steel 1.27  1.27 300 —
Paper Stainless steel — 300 —
Paper Stainless steel 2.54  2.54 200 —

A test plate of 40  135 mm should be used for loop tack [58]. The
nature of the joint failure should be characterized as follows:
A ¼ adhesive break (no adhesive on the plate)
B ¼ cohesive break (adhesive traces on the plate and carrier)
A0 ¼ adhesive transfer (adhesive on the plate)

The intermediate steps are characterized in percent.


Loop tack is probably the most common test method. A finite element
analysis of the loop tack test is being developed by [69]. In this case the loop
backing is modeled as a flexible, inextensible, elastic strip whose curvature
is proportional to the bending moment. The Polyken probe method is the
most well known among the probe tack tests [70].

Polyken Tack. Polyken tack is a measure of the tackiness of PSAs.


The Polyken probe tack testing machine provides a means of bringing the
tip of a flat probe into contact with PSA materials at controlled rates,
contact pressures, dwell times, and subsequently measuring the force
required to break the adhesive bond in g/cm. A Polyken Probe Tack Tester
(Kendall) according to ASTM has been described [71]. Here a 120-g force
and a 0.5-cm sample diameter are used. Polyken tack is tested with a 100-g/
cm contact pressure, 1-sec dwell time, and a 1-cm/sec test rate [72]. Since this
method implies a very small contact area, small imperfections in the
adhesive film can lead to disparate values.
The use of a probe to test tack is an extension of thumb tack testing, a
probe test method having been used in the 1940s [73]. Kamagata et al. [74]
tried to control conventional thumb tack testing by using a small balance
[31]. Another earlier method was the Ball Tack Test using a steel ball as
contact surface [74]. In the 1950s Wetzel [75] proposed a probe testing device
that could be fitted into a standard extensiometer. Later this method was
Test Methods 651

redesigned into the Polyken Probe Tack Tester. The probe is attached to a
force gauge; the sample to be tested is attached to another weight to control
the applied pressure which then is lowered onto the probe and pulled off.
The equipment applies a 100-g/cm2 pressure with a 5-mm diameter stainless
steel probe, a contact time of 1 sec, and a rate of removal of 1 cm/sec. There
also is the Kendall probe tack tester for research investigation purposes. The
Polyken Probe Tack Tester and the test method also have been suggested for
the evaluation of readhering adhesives [71,74].
A more precise version of the probe tack method is the one developed
by Druschke [68] enabling the use of contact times as short as 0.01 sec. The
probe tack method proposed by Wetzel [75] and developed by Hammond
[71] has the disadvantage that a special apparatus is necessary and the
contact area is too low (< 0.2 cm2) [68]. Table 10.10 presents the experi-
mental conditions used for probe tack tests.
Historically the Probe tack test equipment was developed from the
common industrial apparatus used in the field of other, non-pressure-
sensitive adhe-sives, for the measurement of the bond strength as a function
of the time. The industrial renaissance of this test method is due to its ability
to be quantified exactly (normed contact force, contact time, contact and
de-contact speed).
The scientific rebirth of this procedure is forced by the possibility of
a theoretical interpretation of the test results, which allows the correlation
of the contact mechanics, rheology, and macromolecular basis of the adhesive
(see Chap. 3, Section 3 also). Owing to a complicated stress distribution,
peel test values cannot describe the molecular origins of the adhesion. For
this reason axis-symmetric adhesion tests are favored in studying the
interfacial and bulk contributions to adhesion. With this method accurate
measurements of the relationship between the energy release (G) and
the crack tip velocity (v) can be made. The Probe tack test is a variant of

Table 10.10 Probe Tack Test Methods

Test Conditions

PSA (Thickness, m) Contact pressure Contact time (s) Debonding rate Ref.

PEHA (100) l MPa 1 30 (m/s) [76]


SIS 1 kPa 1 — [77]
SIS (70) — <1 — [78]
PIB (70–80) 5N <1 1.0 mm/s [79]
SIS triblock (10) 1 kPa 1 [79]
652 Chapter 10

Figure 10.1 Schematic diagram of the axis-symmetric adhesion test.

the axis-symmetric adhesion test (indenter test). A schematic presentation


of the axis-symmetric test is given in Fig. 10.1.
The main disadvantages of the Probe tack are given by the extreme
sensitivity of the method concerning the quality of the contact area. For
Probe tack, the effective work of adhesion is given by the correlation
between the area under the tensile load versus displacement curve
normalized by the area of contact at maximum compressive load [80].
This technique consists of the formation and separation of a bond between a
thin polymer layer and a hard flat surface. A thin compliant layer of the
adhesive is bonded to a rigid substrate. A rigid flat punch is brought into
contact with the adhesive at a fixed rate. The layer is then separated from the
punch, and the energy required for the pull out process is measured by
finding the area under the load vs. displacement curve. For strong adhesives,
failure during the pull out phase occurs by cavitation and fibrillation of
the adhesive. Therefore the tack test can be used to determine a phenomeno-
logical model for fibril deformation and to predict the peel rate and the peel
force too [81]. According to Kendall [82] the sphere adhesion (i.e., Probe
tack with spherical punch) result is rather similar to the peel test. (As is
known the geometry of a spherical punch pressing against a flat substrate
has been used in fundamental studies of the tackiness of crosslinked
elastomers). PSA samples may be in the form of cylinders too. The case
of contact between two crossed cylinders with equal radii of curvature is
identical to the contact between a sphere of the same radius with a flat
surface [83]. Although they present alignment difficulty and fibrillation upon
pull-off, flat punch tests have often been the choice of industrial research
Test Methods 653

Figure 10.2 Discrepancies between tack test data (rolling ball, loop tack, and
Polyken) for pressure-sensitive adhesive formulations based on natural rubber (1),
styrene-butadiene rubber (2), and styrene-isoprene-styrene block copolymer (3).

laboratories and quality assurance and are the basis of commonly used
ASTM standards [84]. The softness of the adhesive and the roughness
strongly affect the measurement and may produce discrepancies in the test
results. Such discrepancies are accentuated in the case of styrene block
copolymers (Fig. 10.2). It is well known from the testing practice of PSPs
that classical tack test methods (Polyken and Rolling ball) give results which
differ from the Loop tack data. They deviate from the parallelism generally
observed for changes in loop tack values and peel resistance values. As
illustrated by the data of Fig. 10.2 such a deviation is a function of the
adhesive system. The greatest deviations are observed for pressure-sensitive
adhesive formulations based on styrene-block copolymers. Until recently
there was no explanation for such behavior. The investigations made by
Hooker et al. [77] suggest that such deviations are due to the increased
rigidity of styrene block copolymers. It has been shown that for SIS-based
PSA, at normal temperature, surface roughness is detrimental for good
contact. As shown in [76] Probe tack is the optimal technique to study
adhesion properties of soft polymers for short contact time conditions.
Polyken tack is less strongly affected by the resin softening point (it is
applied under load!). Probe tack tests of PSAs on silicone release coatings
were carried out too.
Custom designed Probe tack apparatus allows the simultaneous acqui-
sition of a nominal stress and strain curve and the observation of the adhe-
sive film from underneath the transparent substrate. Probe tack tests
demonstrate the interdependence between tack and cohesion. Experimental
results show, that the beginning of the Probe test (where the cavitation
654 Chapter 10

process takes place) quantitatively corresponds to the small strain shear


properties of the adhesives while the end of the test (fibrillation, strain
hardening) can be qualitatively—so far—related to the large strain
elongation processes [48]. Sophisticated equipment (see the Mechano
Optical Tack Tester) allows the determination of the tack strength and
tack energy also. (The tack/cohesion interdependence was discussed in
Chap. 6, Section: ‘‘Tack Measured as Cohesion’’). The use of the indenter
test method to correlate the macroscopic aspects of adhesive debonding with
the rheology and with the macromolecular characteristics, was discussed
in detail in Chap. 3, Section 3.

Rolling Ball Test. The rolling ball tack (RBT) test measures tack as
a function of the distance traveled by a steel ball on an adhesive-coated
substrate (see Chap. 6, Section: ‘‘Measurement of the Tack’’; Subsection:
‘‘Tack as Coefficient of Friction’’ also). The Probe tack test can be viewed as a
constant deformation rate test on the behavior of the fibrillated materials, the
rolling ball cannot. The rolling ball tack, according to PSTC Method 6
is measured as follows: a stainless steel ball is allowed to run down a slope
from a point down the inclined plane onto the adhesive tape; the distance
traveled along the sample is measured in centimeters. In this test an 11-mm
diameter steel ball rolls down a plane having a length of 18 cm and inclined at
an angle of 21 300 , the adhesive thickness being at least 25 m. In a modified
rolling ball test, according to PSTC 6, an inclined angle of 21 300 , a necessary
length of 12.5 cm, and a ball weighing 7.59 g with a diameter of 1.229 cm
(0.484 in.) is used. In a different manner from other tack testing methods, the
rolling ball method does not measure the debonding force [36]. The apparatus
used is relatively simple, but unfortunately the test is influenced by adhesive
residues on the ball or by the viscosity and thickness of the adhesive layer.
Table 10.11 summarizes the characteristics of the tack test methods by rolling
ball. As illustrated by data of Table 10.12 low tack pressure-sensitive
products cannot be characterized adequately using the RBT test.

Table 10.11 Rolling Ball Tack; Characteristics of the Test Method

Travel Path Rolling Ball Characteristics

Method Length (mm) Angle Diameter (mm) Weight (g) Ref.

— 173 25 14 4 [87]


PSTC 6 180 21 300 11.1 — [88]
PSTC 6 125 21 300 12.29 7.59 [59]
— 180 21 300 — — [59]
Test Methods 655

Table 10.12 Rolling Ball Tack Values for Various PSPs

Adhesive Properties

Product Tack (RB, cm) Peel (180 , N/25 mm)

Label 3,0 1,0


Tape 2,5 3,0
Protecting film 25,0 1,0
Separation film >100,0 0,2

Table 10.13 Dependence of the Tack on the Coating


Weight; Rolling Ball Versus Rolling Cylinder

Tack Values

Coating Rolling Rolling


Weight (g/m2) ball tack cylinder tack

16 10.4 4.5
10.7 4.8
10.0 4.2
19 4.5 4.5
5.0 5.0
5.0 4.0
27 3.4 3.0
3.1 2.9
3.1 3.0

Rolling Cylinder Tack. The rolling cylinder tack (RCT) method is a


modified version of RBT, known as the ‘‘Douglas tack test.’’ Here one uses
a stainless steel cylinder with a diameter of 24.5 mm, a length of 19.05 mm,
weight of 75 g, and a travel path 203 mm long with an angle of 5 [3].
The reproducibility of the RCT values is better than the reproduci-
bility of the RBT values; the running path is shorter and less material is
necessary (as compared to competitive samples). For paper laminates values
of 1.5–2.5 are very good and values of 7.5–8.0 fair. No more than 15 cm
can be accepted. For film laminates values of 2.5–3.5 are very good, values
of 4.5–5 are fair, and values above 7–8 cm are not acceptable. Rolling
ball tack is more sensitive towards the coating weight than rolling cylinder
tack (Table 10.13).
656 Chapter 10

Other Tack Test Methods. Another method proposed by Bull [85] (in
a manner different from the rolling ball method where the adhesive surface
is static, and the adhesiveless contact surface rotates) has a polished rotating
adhesiveless cylinder in contact with a rotating adhesive-coated cylinder.
This method measures the force (resistance) necessary to move the cylinders.
Unfortunately, a special apparatus is needed. Like in the rolling ball tests,
debonding is influenced by the compression of the adhesive by the cylinder.

Tack Measured with the Toothed Wheel Method. Druschke [68]


summarized the different factors influencing the adhesion (and the test
method); they are:
the surface tension of the adhesive, face stock, and adherend;
the mechanical (viscoelastic) properties of the adhesive, face stock,
and adherent;
the surface structure of the adhesive, release liner, and adherent;
the thickness of the adhesive, face stock, release liner, and
adherent;
the test conditions.
Taking these parameters into account, Druschke improved the
rotating drum method. A further development of the rotating drum
method is the pitch-wheel method [86]. In this method the polished wheel
used in the rotating drum method is replaced by a pitched wheel, and the
contact force is measured using a common tensile strength test machine.
Druschke [68] proposed the toothed wheel method which permits very short
contact times, as does the rotating drum principle. In the new method, a
toothed wheel rotating freely around its axis is fixed in the upper clamp of a
tensile tester with a liner. A rotating drum is pushed against the wheel until
an almost pressureless contact is achieved. The tensile tester measures the
force acting against the rotating drum with the adhesive layer.

Mechano Optical Tack Tester. This method is a variant of the probe


tack test. Tack measuring with a Mechano Optical Tack Tester, MOTT
(a device developed by Elf-Atochem) [78,81] allows the determination of the
tack strength, tack energy, and the actual contact area, for tests at controlled
contact force, contact time, and release rate. It measures simultaneously the
contact area between the probe and the PSA together with the tack force
and energy during the contact and separation phases. This apparatus is
made up of a quartz prism probe linked to a force transducer. The whole
system can be moved downwards and upwards at constant speed in the
bonding and debonding phases. During the contact phase the actual
thickness of the adhesive is measured with a micrometer fixed to the prism
Test Methods 657

support. An optical device which moves with the probe provides a measure
of the contact area at any time. The ratio of the transmitted intensity and the
initial intensity of the light gives the percentage of contact area. A constant
contact area is applied for very short contact time (less than 1s) and the
probe is pulled up. The force and the contact area are monitored vs. time.
The product of the area subtended by the debonding force vs. time by the
pulling rate gives the tack energy.
Tack was measured as plasticity too (see Chap. 6, Section 1.1). The
influence of various test parameters on the tack (e.g., adhesive nature, face
stock material, and coating weight) was discussed in Chap. 2, Section 1.2
and in Chap. 6, Section 1.1.

Evaluation of the Peel Adhesion


Peel adhesion and the principles of its measurement were discussed in detail
in Chap. 6, Section 1.2. A combined method (an assembly of different test
methods) was proposed for the evaluation of different adhesive character-
istics of hot-melt PSAs [89]. Peel adhesion should be tested according to
PSTC 1, shear according to PSTC 7, and tack according to PSTC 6 (roll-
ing ball), Polyken, or quick stick tests. Melt viscosity is measured
with Brookfield RVT spindle 7, 20 rpm. The softening point should be
determined with the ring-and-ball method (ASTM E 2867) at 180 C [89]. A
computerized system for measuring the peel adhesion interfaces the slip peel
tester to a personal computer. This system measures peak, valley, and
average peel forces in various geometries [90]. Using a statistical software
package peel strengths were measured for two PSA tapes under controlled
stress and controlled rate [91]. These examples illustrate that although there
are standard methods for measuring the peel, really quite different ‘‘home
made’’ methods are used to measure different adhesive performance
characteristics. Kaelble [36] showed that peel depends on the modulus and
thickness of the adhesive, the solid state component of the label, and on
the peeling angle. The best known test of the adhesion is the method used
for lacquers according to DIN 53151 and ISO 2409. Adhesive strength (peel)
also may be determined according to ASTM D 1000. Here the peel or tensile
strength measurement should be used for the evaluation of the bond
strength. This was a question in the first years of the development of PSAs.
For a long time tensile strength measurements and peel measurements were
carried out in parallel [54]. Kemmenater and Bader [92] proposed a tension
measuring device for the peel. FIPAGO accepted a pendulum device [93,94];
in this test, the adhesive strength was tested via peel work in mm/kp
[95]. Today the most common method for testing the adhesive strength of
PSAs is the peel test. Different versions of this method using different peel
658 Chapter 10

Figure 10.3 Test methods used for peel evaluation. 1) 180 peel; 2) 90 peel; 3)
T-peel.

angles, peel rates, and substrates are used. The history of peel testers was
described by Liese [93]. FIPAGO recommended the adhesive strength tester
of PKL used for gummed papers. The Werle Tack tester measures adhesive
strength in a similar manner based on the peel adhesion [94]. A systematic
study of the test methods for wet adhesive tapes was reviewed during the
FIPAGO meeting in 1964.
For different end-uses, labels for different applications, varying peel test
angles are used. For a flexible and a rigid adherent a 180 peel or 90 peel, and
for peel between two flexible adherends a 180 T-peel is carried out (Fig. 10.3).
Common angles are 90 and 180 ; for label stock the 180 geometry according
to FINAT FTM 1 has found wider application than the 90 geometry [50].
As a result of the viscoelastic properties of PSAs and the mechanical
properties of the face material, the adhesive bond strength depends on the
peel rate (see Chap. 2). In most test equipment single point measurements at
5 mm/sec are specified. Test results show that the suitability of a range of
products may vary with the peel rate [50]. This should be considered when
practical application conditions differ strongly from test conditions. A
similar statement holds true for release forces from siliconized release paper.
Dwell times (the contact time between adhesive and substrate) also influence
the peel value. This is very important especially for peel tests on difficult
surfaces. In some cases (for single-coated tapes) the 180 angle PSTC 1
Test Methods 659

method is modified in order to allow 20 min or 24 hr contact of the adhesive


with the test panel [41].
Peel adhesion is determined in accordance with ASTM D-3330-78,
PSTC 1, and FINAT FTM 1 and 2, and is a measure of the force required to
remove a coated flexible sheet material from a test panel at a specified angle
and rate of removal. According to US norms, a 1-in. wide coated sheet is
applied to a horizontal surface of the clean, stainless steel test plate with at
least 5 in. of the coated sheet material in firm contact with the steel plate. A
rubber roller is used to firmly apply the strip and remove all discontinuous
and entrapped air. The peel adhesion of the sample then is measured in a
tensile tester. In [91] peel was measured using ASTM method D3330/D3330
M-99 Standard Test (‘‘Methods for Peel Adhesion of Pressure-Sensitive
Tape’’). Adhesion of the latexes was tested using ASTM D897 [96] and
ASTM D330-90 [97].
According to the 180 peel adhesion AFERA 3001 method, one
10-mm wide specimen of PSA film (prepared at least 24 hr before running
the test) is applied on a stainless steel plate. Ten minutes later the specimen
is peeled off by means of a tensile tester at a speed of 304 mm/min. The
measurement is carried out at 23  2 C. In fact, tests according to PSTC 1
and AFERA 4001 PB are carried out with a 25-mm wide sample, peeled off
at 300 mm/min speed, and the peel force is given in g/25 mm [95]. FIPAGO
uses the measurement of the peeling work for the evaluation of the peel
adhesion. A modified PSTC 1 method [98] prescribes a polished stainless
steel panel as a substrate. After 1 min dwell time the tape is peeled away
from the panel at an angle of 180 and at an 8 in./min separation speed. The
180 peel adhesion is the average of five test values and is expressed in g/in.
width of tape. Another modified PSTC 1 method on a 2.5  25  125-mm
glass plate uses a tape applicator. The peeling force is registered at 12 mm
intervals and given in N/25 mm. Alternatively, a strip of 2.54-cm wide
coated paper is bonded to a glass plate by applying a constant pressure [70].
A tensile testing machine then is used to measure the force required to peel
the paper strip at an angle of 180 at a rate of 300 mm/min at 20 C. The
testing speed for film laminates should be 300 mm/min and 75 mm/min for
paper. Peel adhesion should be tested according to FINAT FTM 1, but at
a peeling rate of 200 mm/min; 180 peel adhesion should be checked.
Apparatus:
Tensile testing machine according to DIN 51220, Class I, and DIN
51221 (tensile strength).
Glass slides (2.5  45  125 mm).
Tape applicator D-4271 (Sandes Place Research Institute).
Cleaning solution.
660 Chapter 10

Method [99]:
Clean the glass slides for 24 hr in the cleaning solution; flush with
deionized water and then dry the glass slides.
A strip of coated material (25  250 mm) is bonded to the glass
plate by applying a constant pressure (by tape applicator or roll).
For instance, samples for peel were rolled down with two slow
passes of a 2 kg roller with a 10 cm diameter and a 0.5 cm thick
rubber covering [100]. Rolling of the tape can be carried out with
a 4.5 pound roller too [101].
The noncontacted portion of the sample then is doubled back
(nearly touching itself ) so that the angle of removal of the strip
from the plate will be 180 . The free end of the test strip is attached
to the adhesion tester. The test plate then is clamped in the jaws of
the tensile testing machine.
The peel force should be rated at each 12 mm interval.

90 Peel Adhesion. The 90 peel adhesion is measured as the force
required to remove an adhesive-coated material from a test plate at a specific
speed and at an angle of 90 [95]. According to Johnston [102] 180 peel is
a combination of tensile strength and shear, while a 90 peel is one of tensile
only. Peel decreases with increasing angle, 90 peel is about twice that of 180
peel [103]. Wilken [50] obtains the maximum peel value between 90 and 180 ,
about the same peel resistance value for 90 and 180 peeling angle, and shows
that below 90 the peel increases as a result of the increase of the shear stress:
P180 ffi P90 > P!0 ð10:1Þ

Our data show that 180 peel is higher than 90 peel (see Tables 6.11 and
6.12). According to [104] for carrier materials with higher thickness
and stiffness the 180 peel gives higher values than the 90 peel. The peel
resistance increases proportionally with the carrier thickness (PETP) and
then more slightly:

P180 > P90 ð10:2Þ

The difference between the peel values at different peeling angles (180 or
90 ) depends on the chemical nature of the PSA. Soft rubber-resin PSA
gives much lower 90 peel. For the 90 peel test the apparent strain energy
release rate can be calculated from the equation: G ¼ P/b where P is the
PSA tape width. A sample of PSA-coated material 1-in. (2.54 mm) wide and
about 4-in. (about 10 cm) long is bonded to a target substrate surface (glass
Test Methods 661

unless otherwise stated) [41]. One end of the adhesive-coated sheet is peeled
off the target substrate and held in a plane perpendicular to the target
substrate which is restrained in its initial plane. The coated sheet is pulled
away at a constant linear rate of 300 mm/min and an angle of 90 . The
peel adhesion at 90 is twice the value of 180 peel [24]. For 180 peel
measurements the results depend on the face stock material [78]. For thicker
and stiffer materials a higher peel value will be obtained than at 90 [105].
Especially for soft PVC, 90 peel gives more exact values [106]; 90 peel of
PSA tapes was measured on glass plate coated with release [107]; a special
case of the 90 peel adhesion test is the butt tensile test. This test (90
adhesive test) restricts the evaluation to a single narrow area, in contrast to
a standard peel test in which a continually new adhesive part is being
examined [108]. Special peel adhesion measurements at different angles and
temperatures (between 45 and 90 , at 50 C and 110 C), as well as hot shear
gradient (SAFT) measurements were proposed. The parameters influencing
the peel test were discussed in detail in Chap. 2, Section 1.2 and Chap. 6,
Section 1.2.

T-Peel Adhesion. The T-peel adhesion test evaluates the ‘‘unwind’’


characteristics of the wound tape [25]. The tape is bonded to the other piece
of the same film. The ends of the two films are inserted into the opposing
jaws of a tensile machine which then pulls the two films apart at an 180
angle. T-peel adhesion may be measured according to ASTM D-1976, DIN
53281, or EN 1994D. This method is intended primarily for determining
the relative peel resistance of adhesive bonds between flexible adherends or
when standard peel values are too low. The adherents should have such
dimensions and physical properties as to permit bending them through any
angle up to 90 without breaking and/or cracking. For instance, Hamed and
Preechatiwong [109] used T-peel measurement for the adhesion test
of SBR pressed between PET and cotton fabric at 140 C. The T-peel test,
according to ASTM D-1976-61T is carried out at a speed of 5 in./min; EN
1994D uses a speed of 300 mm/min.
A major attraction of peel tests is their apparent practical simplicity.
However, in both testing and data interpretation, attention to detail is
important if reliable and useful results are to be obtained. Williams et al.
[110] proposed a peel test protocol. The peel test protocol [111] used for
flexible laminates has the purpose to provide guidance on the measurement
of the peel strength of the laminate and then to show how the adhesive
fracture energy, Gc, can be determined from the peel strength and other
measurements. The protocol is divided into two parts: one for a ‘‘fixed arm
peel’’ test and the other for a ‘‘T-peel’’ test. These geometries are different
but there is a common theme in converting the peel strength measurement
662 Chapter 10

into an adhesive fracture energy, Gc. The current challenge is to model


accurately any extensive plastic deformation which may occur in the flexible
peeling arm, since if this is not accurately modeled then the value of Gc
deduced may suffer a high degree of error. As described in [112] for self-
adhesive polyolefin films (SAF) the so-called peel cling and lap cling are used
to characterize the product. In this case the peel cling is a T-peel of the
SAF on itself; lap cling is adhesion of the SAF on itself, tested by shear
measurement. Cling of cling-stretch SAF is measured according to ASTM
D-4649 after 1, 6, 14, and 40 days after manufacturing.

Influence of the Peeling Rate. The peel strength is proportional to the


ratio of G00 and G0 at the respective debonding and bonding frequencies. At
low peel rates the peel adhesion depends mainly on the work of adhesion,
but at high peel rates it is influenced by the work of deformation. Therefore
factors to be varied by the peel test include the peel rate, time after
application, and substrate. The influence of the adhesive thickness, viscosity,
backing rigidity, and peeling velocity in the cohesive regime have to be
tested [113]. Peel and tack increase with increasing test speed, but only up to
a limit [114]. At very high measuring speeds the tack increases again. An
increase of the temperature gives increased molecular mobility resulting in
increased tack and reduction in shear resistance. A high stress rate is equal
to a drop in temperature. Indeed, universal testing of a roll of tape occurs
at 12 in./min, while modern slitting and unwinding machines, or tape
application processes, may run at several hundred feet per minute. At harder
unwind, adhesive pick-off or transfer may occur [108]. For industrial
practice it is very important to know how stable the release value is with
regard to the aging and peel rate conditions [115].

High Speed Peel Tester. This equipment allows a 10–300 m/min


testing speed for release force measurements [116]. With this method the
separation force of the PSA-coated face stock from the release liner is
evaluated at speeds similar to those typically used to convert and dispense
the laminated material. It therefore provides a better characterization of the
material than FINAT FTM 3. According to [117] the stripping speed was
originally set to be 12 in./min, but later the test speed moved upward
through 400 in./min and higher. Label applications can run with 500 feet/
min, thus high speed release testing equipment has been designed in
this range. Very low peel values (from the release liner) may create label-fly
during conversion or application; high values may produce web break when
skeleton stripping or dispensing failure during automatic application occurs.
Release force increases with the time. It is to be noted that high speed
stripping of release (at 15 m/min) gives 25–30 g release force in comparison
Test Methods 663

with low speed stripping (at 3 m/min) of 9–16 g. High speed stripping values
do not depend on storage time. As shown [118–121] the bulk viscoelastic
properties of the adhesive are correlated to the release profile observed from
peel force vs. peel velocity measurements.
The test is carried out like a 180 peel test at jaw separation speeds
between 10 and 300 m/min. In order to ensure good contact between the
release paper and the adhesive, the sample is placed between two flat metal
or glass plates and kept for 20 hr at 23  2 C under pressure. There is an
influence of the way of stripping on the high speed release test results (i.e.,
face from liner or liner from face). Generally, higher results are obtained
when removing the face material from the release liner. According to TLMI
Adhesion Test No. VIII-LD the sample to be tested is secured to a rigid
plate using double-sided adhesive tape. Excess moisture in the carrier
paper can be troublesome, and it is preferred that the papers have a
moisture content per weight of about 4% or less. Predrying should be
employed if necessary.
On increasing the thickness of the release backing the apparent high
speed release values were approximately doubled, over the whole speed
range, when the backing was taken away from the label. Such tests are
very important taking into account the velocity dependent influence of
the release agents on the release liner. As shown in [118] the largest impact of
the HRA on release force is typically at low to intermediate peel velocities
(0.005 to 0.17 m/s).
The peel force measurement is usually made at a single pull rate and
ambient temperature [122]. Generally, the PSA is polymer-based; conse-
quently, it behaves like a viscoelastic medium. The bond strength is not
constant for varying strain rates and temperatures. The rate of separation
and temperature are related to the adhesive strength in a complicated
manner, and these factors can influence the nature of the failure mode
(adhesive or cohesive). This can lead to a situation in which an adhesive is
selected on the basis of its performance at one set of strain rate/temperature
conditions, but ultimately fails because in use it experiences a different set
of loading conditions. Rather than judging adhesive strength at a single
separation rate and temperature, surface contour diagrams should be used
to illustrate the adhesion over a wide range of conditions [120]. This
technique allows a comparison of different PSAs while examining the
maxima and minima of their rate/temperature graphs as well as how a PSA
laminate meets or exceeds a given peel value. A temperature range of 20 to
158 F may be selected, a range of separation rates of 0.50–15 in./min was
suggested.
The time/temperature influence on the peel test was discussed
in Chap. 6. Druschke [68] summarized the different dwell times used by
664 Chapter 10

Table 10.14 Longitudinal Carrier Film Deformation


During 180 Peel Debonding of a Pressure-Sensitive Laminate

Elongation (%)

Code Full length Without postelongation

1 9.0 5.5
2 6.0 5.0
3 10.0 5.5
4 7.0 3.0
Note: PE film with a thickness of 45 m was tested.

AFERA (10 min), PSTC (1 min), and FINAT (20 min and 24 hr) for peel
testing purposes.
Influence of the Carrier Material. As discussed in Chap. 3 and Chap. 6,
Section 1.2, the carrier material strongly influences the peel tests. The
decisive influence of the plastic deformation of the carrier was demonstrated
in [123]. As illustrated by the data of Table 10.14 such deformation plays
an important role for special products having a thin plastic carrier (e.g.,
tamper-evident labels, forms, protective films etc.). Peel tests with flexible
carrier materials at a large angle, with a rigid carrier (disk sample), and with
a semiflexible carrier were compared [82]. Generally the common test of
the adhesive is carried out using standard, nondeformable (without plastic
yielding), isotropic materials with controlled surface properties. Generally
polyethylene terephthalate is applied as the carrier film [103,124]. In
practice, the elongation of the carrier film during debonding is the sum of
the elongations (for the full carrier length) during peel off and after peel off.
Principally, the carrier elongation, which contributes to a supposed coating
weight decrease does not include ‘‘post-elongation.’’
Influence of the Substrate. The influence of both components of a
joint on the debonding is well known (see Chap. 2). The strength of
the adhesive bond for the substrate was found to be ordered like the
corresponding magnitudes of the thermodynamic work of adhesion [125].
Peel adhesion should also be examined as a function of the substrate
(adherend). For example, peel was tested on stainless steel and HDPE [38].
Samples of 20  280 mm on a stainless steel plate with a bond length of
100 mm were evaluated. For peel from HDPE, 0.25-cm thick, high density
polyethylene panels were used. For PUR modified acrylics on film face stock
material adhesion performance was measured on stainless steel and HDPE
as substrates, in accordance with ASTM D330-90 [97].
Test Methods 665

The most important factor in measuring the bond strength is the test
plate [50] which can be float process plate glass (FINAT), brushed steel
(AFERA), or polished steel (PSTC and BWB). Glass remains the most
suitable material. The FINAT Test Method 1 proposes the use of float
process plate glass as substrate or test material. All peel test methods except
FINAT suggest stainless steel plates as the adherent [68,100,103,126].
Standard tests of tack and peel are made on steel (PSTC-5 and 1),
but measurements on polyethylene, polypropylene, cellulose, PET, PA, and
glass are often suggested. Adhesion on different substrates (aluminum, glass,
nylon, acrylic sheet, polycarbonate, rigid polyethylene, rigid polypropylene,
polystyrene, stainless steel) was tested [127].
Generally, peel adhesion should be evaluated with and without the
use of a primer [128]. The surface energy of the adherent, the compa-
tibility of the surface with the adhesive, and its pH also influence the
adhesion [108].
Paper as substrate displays a special behavior. Paper possesses an aniso-
tropic structure, it is very resistant to in-plane stresses and very sensitive
to out-of-plane or z directional stresses. Thus, paper is very susceptible to
delamination. Paper delamination is characterized by a high initiation force
and a relatively low propagation force. Experimental conditions strongly
influence paper delamination [129]. Bikermann and Ehitney [130] first
reported that paper is much more likely to delaminate in peel if the tape
overlaps the edge of the paper. In the case of interfacial failure the peel curve
is noisy but approximately constant, resembling the peeling from stainless
steel. The peel curve at higher peel rates is more complicated. The peel force
initially rises until the peel force applied to the paper surface is high enough
to remove fibers after which the force drops to a low steady-state value,
corresponding to catastrophic delamination (paper failure). Yamauchi
and coworkers [131] reported peel forces and failure modes as a function
of peel rates and paper types. Recent work at McMaster University has
extended the examination to the role of the surface energy and application
pressure too [132].

Peel on Cardboard. It is essential that PSAs are evaluated against the


final intended surface, namely the National Bureau of Standards reference
material 1810 [108]. Because cardboard has a grain that is dependent on the
direction of the manufacturing process, this should be determined prior
to any testing so that the same direction is used in any measurement. An
adhesive tape can cause 100% delamination of the cardboard when stripped
in one direction, but may only cause 5–10% fiber tear when stripped in
the perpendicular direction. In the cardboard fiber tear test, a strip of tape
is applied to the cardboard substrate, affixed with a 2-lb roller and removed
666 Chapter 10

immediately by a pulling force extended on the tape at a 90 angle to the


substrate surface. The adhesion properties of a product vary directly with
the properties of its surface covered by fiber [25].

Peel Adhesion of Plasticized PVC Film. A PVC film/1 mil PSA layer/
siliconized paper laminate is proposed for the evaluation of the adhesion of
plasticized PVC film [72]. After specified aging times, the siliconized paper
is removed and the adhesive-coated PVC film then is adhered to a stainless
steel panel. After a 24 hr dwell time the 180 peel strength is determined
according to PSTC test method 1.

Polyethylene Peel on Polyethylene Plate. This method tests the 180


peel adhesion on polyethylene using a rigid polyethylene surface (plate). The
test method is the same as that used for polyethylene foil but with HDPE
plates instead; the cleaning of the plates should be carried out with n-hexane.
Adhesion to HDPE is lower than to LDPE [133]. There are different bond
failure modes:
1. Face failure: a 100% transfer of adhesive to the substrate.
2. Clean release failure: 100% adhesive release from the substrate.
3. Cohesive failure: the adhesive splits between the face and the
substrate.
The nature of the failure is very important for removable PSAs. Papers with
a different surface treatment yield different peel values for the same
adhesive. Clay-coated papers give paper tear if the superficial strength of the
paper is not sufficient. Permanent adhesives should cause destruction of the
paper (paper tear); partial destruction is noted as paper strain [134] which
generally occurs at a force of 1500 g/in.

Drum Peel Test. The adhesion can be tested using a drum peel
configuration on a single screw driven mechanical test frame. This test
allows a 90 peel test. Such peel testing was accomplished for medical
PSAs using a peel test wheel (e.g., TA.XT2i, texture analyzer, by Texture
Technologies Corp., Scarsdale, NY) [135].

Special Peel Test Methods. It has been demonstrated, that single pull
peel testing does not always predict the joint failure that occurs as a result of
cyclic loading at subcritical loading levels [136–138]. Therefore in [136] a
combined peel test method was proposed. Samples were exposed to a single
pull, 180 peel test, at the displacement rate of 250 m per second. These
data were analyzed to determine the load at first yield. Samples were then
exposed to subcritical cyclic loading between 20 and 300 mN per cycle, at
frequencies of 1, 5, and 10 Hz to determine the rate of peel under the cyclic
Test Methods 667

Figure 10.4 Dependence of the debonding force on the elongation for 1) standard
and 2) zip-peel product.

protocol. The results clearly indicate that these adhesive joints when
critically loaded at levels below the load at first yield in the single pull
experiment, can open up at appreciable rates.
From the adhesive performance point of view it is a measure of how
much ‘‘loose’’ silicone will be transferred on to the laminated adhesive and
what percentage of peel and tack performance are retained (subsequent
adhesion) after delamination from the release liner. Therefore special peel
tests are carried out [115].
For high surface energy, film-based PSPs instantaneous debonding
above a critical debonding force plays an important role [139] (Fig. 10.4).
For such products zip-peel (sawtooth-like force/time plot) is desired. The
test of zip-peel products requires the use of a modified tensile strength
equipment with a spring-balance (Prohaska device), having the resistance of
the maximum peel force of the product.

Other Adhesion Test Methods


In practice the debonding of labels or protective film occurs mainly by
peeling off. Tapes and packaging material applied using tapes may suffer
other debonding operations too, where delamination starts at an arbitrary
place of the assembly which is not situated at the end of the laminate. Such
debonding can be compared to blistering. The most common and
convenient method to measure the fracture energy (Gc) is peel testing. A
disadvantage of peel testing is the significant plastic deformation that the
peel arm undergoes. The extent of plastic deformation is a function of peel
angle and peel rate and therefore the measured Gc values are said to
668 Chapter 10

Figure 10.5 Schematic presentation of the shaft loaded blister test.

be geometry dependent. An attractive alternative to peel testing is the


pressurized blister. The small angle between the film and substrate
minimizes plastic deformation and the known stress distribution results in
a measured fracture energy closer to the true Gc. However, the experimental
set up is very difficult and if the flow rate of the fluid is not controlled,
catastrophic debonding occurs. The shaft loaded blister test (SLBT) may
be a suitable alternative to either test in that it has both the ease and
convenience of peel testing and the plastic deformation characteristic of
the pressurized blister test.

The Shaft Loaded Blister Test. The shaft loaded blister test (Fig. 10.5)
has a number of advantages over the more conventional peel test, because of
the easy experimental set up and because the low angle of deflection between
the peel arm and the substrate reduces plastic deformation [140–143]. The
theoretical framework for measuring the fracture energy using the shaft
loaded blister test was developed by Wan and Mai [144]. The SLBT test is
experimentally similar to the pressurized blister test [145]. The two critical
assumptions utilized for the SLBT are that the film undergoes pure elastic
stretching, or in other words, no bending occurs, and the load is
approximated as a point load. Utilizing the shaft loaded blister test the
apparent strain energy release rate was measured for a PSA tape bonded to
either an aluminum or Teflon substrate [143]. Comparative measurements
were carried out by pull test and 90 peel test. A good agreement was found
for the low energy substrate only.
Test Methods 669

The Probe tack test method contact using two crossed cylinders with
equal radii of curvature was proposed for the test of adhesion too [83].
Comparative studies of the changes in surface adhesion of PSAs were
carried out with AFM (atomic force microscopy) and the spherical indenter
test [146]. The forces probed with AFM are on the order of nN, while those
probed with the spherical indenter are on the order of mN. Penetration
depths are only several to hundreds of nanometers for AFM, but several
microns for the spherical indenter.

Measuring the Shear


The cohesion and the shear resistance and their theoretical basis were
described in detail in Chap. 6, Section 1.3. The influence of the viscoelastic
properties on the shear was described in Chap. 2, Section 1.2. As discussed,
the cohesion of an adhesive may be estimated as its resistance to shear
forces. Similar to tack and peel measurements, there are standard and
special methods for the evaluation of the shear properties. There are a lot of
different methods providing information about the cohesive strength. These
include shear or holding power, shear at a 20 angle, heat distortion tem-
perature, heat resistance, 90 peel, shear resistance versus humidity, etc. [147].
Different methods for measuring shear based on the examination of shear,
peel, and tensile strength were described [148]. A tensile strength measure-
ment also may be used for the evaluation of the cohesion [106]. On the other
hand, shear measurements may characterize other adhesive properties.
Shear tests are used for the characterization of the mechanical properties of
adhesives [149]. The temperature resistance via 45 shear at 70 C has been
measured [106].
There are a lot of different methods based on quite different test
parameters for shear testing purposes. The most important methods
(FINAT FTM 8 and PSTC 7) measure holding power or shear adhesion,
and record the time to failure at room temperature. Other methods are
carried out at 65 C (Shell), 50 C (Ashland), or 70 C (Rohm). Some tests
involve temperature measurement (shear adhesion failure temperature,
heat distortion temperatures). Here the temperature is gradually increased
and the temperature at bond failure measured; for the shear adhesion failure
temperature (SAFT) a 40 F/hr gradient is generally used. The slip distance
is measured by Wacker Chemie GmbH (70 C, 5 min). Others measure
the temperature at which a change from adhesive failure to cohesion failure
occurs.
The test conditions differ also. Shear should be tested on 5 cm2
samples, at room temperature or 50 C with a 1-kg load, and conditioned for
12 hr [58]. Holding power is tested statically according to different norms
670 Chapter 10

such as AFERA 4012, TL7510-011, or US Federal Test 20554 (PPT-T-


60-D) [150]. These examples illustrate that there are many different shear
measurement methods although standardized ones were defined.

Standard Tests for Shear Strength. Shear is determined in accordance


with ASTM D3654-78, ASTM D3654-88, PSTC 7, and FINAT Test
Method 8, and is a measure of the cohesiveness (internal strength) of an
adhesive. It is based on the time required for a statically loaded PSA sample
to separate from a standard flat surface in a direction essentially parallel to
the surface to which it was affixed with a standard pressure.

Statical Shear—Room Temperature Shear. Each test is conducted on


an adhesive-coated strip applied to a standard stainless steel panel in a
manner such that a 0.5  0.5-in. portion of the strip is in fixed contact with
the panel with one end of the strip being free. The steel panel with the coated
strip attached is held in a rack such that the panel forms an angle of 178–
180 with the free end of the tape extended; the latter then is loaded with a
force of 500 g applied as a hanging weight from the free end of the test strip.
In the test of the shear resistance to a fixed load at 178 to the horizontal, the
2 tilt is to prevent peel off. The elapsed time required for each test strip to
separate from the test panel is recorded as shear strength. According to
FINAT FTM 8 (a 178 shear adhesion test) a 25-mm wide specimen of PSA
film (prepared at least 24 hr before running the test) is applied to one end of
a stainless steel plate in order to obtain a contact area of 25  25 mm. Five
to ten minutes after applying the specimen on the plate, the assembly is
suspended from a plate holder, which maintains the plate at an angle of 2
from the vertical. A 1-kg weight is immediately attached to the clamp.
The time recorder is automatically switched off by the weight as it falls.
According to AFERA 4012-P1 (modified) a 25  12.5 mm sample is used
in combination with a 1000-g weight [151]. Besides the standard, normalized
shear test methods there are several modified versions. In another test
method, a strip of coated paper is fixed to the edge of a glass plate so that an
area of 6.45 cm2 (1 in.2) is in contact with the glass. The force required to
remove the strip at an angle of 2 is measured at a separation of 1.0 mm/min
and 20 C [70].
As in the case of peel measurements the influence of the contact time
(dwell time) on the shear test results should be considered. Shear adhesion
testing according to PSTC 7 (1000 g/in.2 load) implies a dwell time of 24 hr.
Shear should be measured after a 20 min dwell time, using a 25  25 mm
or 25  12.5 mm area, with a load of 5 or 10 N [2]. As test substrates,
stainless steel or glass should be used. Tests should be carried out at 25 C
and 50% RH. Failure time or slip distance should be defined as shear
Test Methods 671

resistance. The influence of the sample dimensions should also be taken into
account. Room temperature shear should be measured using an increased
contact surface in order to avoid errors by coating defects. For a modified
version of FINAT FTM 8, a load of 2 kg and a contact area of 12.5 cm2
on aluminum test plates are used. A strip of PSA-coated material
(20  150 mm) should be affixed with a light pressure on an aluminum test
plate (previously cleaned with toluene) so that the contact length is 62.5 mm.
At least three measurements should be performed.
The basic problem with the standard lap shear test is that the average
measured shear strength is not a material property that characterizes the
adhesive uniquely. Instead, it is a rather vague quantity that also is strongly
dependent on the geometry of the joint being tested [152]. For instance, if
the overlap were doubled and all other variables left unaltered, the strength
of the joint would not be doubled. Similarly, if the joint geometry and
adhesive are kept constant and the adherend material changed, the apparent
strength of the adhesive will change dramatically. From the point of view of
the measurement efficiency, shear tests are more difficult and less efficient
than tack or peel measurements. Static, normal temperature shear possesses
low reproducibility and requires long measurement times. In order to get
precise results in a short time, hot shear measurements were introduced (i.e.,
shear tests are carried out at elevated temperatures). Another possibility to
accelerate shear tests is the use of the dynamical shear test. For adhesive
characterization purposes not only the time, but also the nature of the
failure are very important.
The mode of bond failure is described by shear tests with different
codes such as P, C, PS, and SPS, where P means panel failure, C means
cohesive failure, PS means failure with panel staining, and SPS means slight
panel staining [153]. Modes of shear failure include:
1. Face stock failure: 100% transfer of the adhesive to the substrate
(adherent).
2. Clear release failure: 100% adhesive release from the substrate
(adherent).
3. Cohesive failure: (zip effect) adhesive splits between the face stock
and the substrate. This results as a direct consequence of an
imbalance between cohesive and adhesive properties and is an
adhesive weakness.
Room temperature shear tests (0.5  0.5 cm, 1000 g) and elevated
temperature shear tests (with a 4.4-psi load) were used [38]. To force the
cohesive failure rather than substrate or panel failure, the polymer was
tested on aluminum foil as substrate, after a 72-hr dwell time and a 20-psi
load. Values of 1500 min to more than 3000 min were recorded.
672 Chapter 10

Hot Shear (Holding Power). This test is a modification of PSTC 7. A


1.27  2.54-cm specimen of the adhesive is mounted on a 7.5  20-cm
stainless steel panel; an aluminum foil is added as reinforcement for the
face material. The panel then is positioned with its longitudinal dimension
so that the back of the panel forms an angle of 178 with the extended piece
of tape, the 2.54-cm dimension of the adhesive extending in the vertical
direction. The assembly then is placed in a 70 C oven and a 1-kg weight
is attached to the free end of the tape. The time required for the adhesive to
fail cohesively is reported in minutes. There are many different variations
of hot shear methods. Shear resistance at elevated temperatures is measured
at 200 and 250 F (2.2 psi) [154]. High temperature shear at 400 F, based on
a 5 kg/in. bond, 0-min dwell time was evaluated [148]. Shear at 70 and
50 C creep are determined also. The method is similar to PSTC 7; only the
weight is reduced from 500 g to 250 g and the temperature is increased from
25 C to 50 C.

Shear Adhesion Failure Temperature: Dead Load Hot Strength


Test. The shear adhesion failure temperature (SAFT) determines the
temperature at which a pressure-sensitive specimen delaminates under a
static load in a shear mode [155]; it is a method of determining the resistance
to shear of a tape under constant load under a rising temperature. A
variation of the hot shear test method can be achieved by mounting the
laminate construction in an oven and suspending a weight from the sample
in such a way that a vertical shear stress is applied to the bond. The oven
temperature may be raised gradually to define the temperature at which the
bond fails [156]. When measuring the SAFT the sample is subjected to a
30-min dwell time at room temperature. The load is attached, the sample
conditioned for 20 min at 100 F, then the temperature is increased at
1 F/min until shear failure occurs upto a maximum of 350 F. The 178 C
(modified PSTC 7) test is a measure of the ability of a PSA tape to withstand
an elevated temperature rising at 40 F/h under a constant force [156]. For a
rubber-resin-based adhesive tackified with a Wingtack resin a temperature
range to failure of 78–87 C was found [157].
Hot-melt PSAs based on Kraton D display SAFT values (on Mylar)
of 60–90 C [158]. The test assembly is supported by one end in a vertical
position, with a weight attached to the other end and heated in an oven for
periods of 15 min; the temperature starts at 30 C and is raised progressively
by 5 C increments after each 15 min period.
The SAFT value is (T þ 5) C, T being temperature at which the bond
failure occurs. The adhesive tape is applied to a glass plate (25  25-mm
overlap) and a 2-kg roller passed over the tape, once in each direction [3].
The glass plate is placed in a temperature-programmed oven and clamped at
Test Methods 673

an angle of 2 to the vertical. A load of 500 g is hung from the tape and the
oven temperature increased by 4 C/min. The temperature at which the tape
drops off the glass plate is recorded.
The validity of the time/temperature superposition principle is
demonstrated by the parallel use of dynamic shear tests (controlled forces)
and SAFT (controlled temperature). Practically, it is easier and faster to use
dynamic shear [70].

Dynamic Shear. This method tests the shear properties of the


adhesive in a tensile tester under an increasing load (force). Current
static-shear test methods use a constant load at longer test times; they show
(related to the nature of the test) poor reproducibility and need very long
test times. As is shown SIS block copolymers have such high room
temperature shear values, that only highly plasticized formulations can be
evaluated in a convenient period of time [159]. Moser and Dillard [160]
developed a dynamic test correlated with the static test. For a current
dynamic shear test following lamination and appropriate conditioning, the
samples are placed in a typical tensile strength testing machine, and shear
adhesion can be defined by applying shear stress at a given rate (dynamic
shear).
Apparatus:
Tensile tester
Glass plates (see peel test)
Method:
A strip of coated material is fixed to the edge of the glass plate so
that an area of 6.45 cm2 (1 in.2) is in contact with the glass.
The force required to remove the strip at an angle of 2 is measured
at a rate of 1 in./min and 20 C.
Other variants of the method exist [70]. Molecular disentanglement can take
place at a testing rate of 0.01 in./0.25 mm per minute, thus a dynamic shear
test can be set up on an adhesion tester at this slow speed [102]. While the
width of the sample can be of any standard width, the height of the sample
needs to be limited, so that it does not become a tensile test of the backing.
That is the maximum shear resistance achieved should be in the region of 2 lb/
1 kg or so. This would normally give a height of around 0.125 in./2 mm.

Dynamic Lap Shear. Here a 3  1-in. aluminum/aluminum assembly


is used with an adhesive coating weight of 3 mils. The aluminum plates are
pulled apart in a 180 configuration at a speed of 1 in./min [161]. The
standard shear method is essentially a static test and therefore no upper value
674 Chapter 10

is defined. A dynamic test method is preferred, using a machine that has the
capability of running extremely slowly (e.g., 0.02 in./min) [108]. A new
dynamic method at a testing speed of 2.5–100 mm/min (as a preliminary test)
and 50 mm/min as a common testing speed, a laminating pressure of 154 kPa,
and temperatures of 23, 40, and 70 C (allowing a rapid assessment of holding
power) was developed [150]. A strip of the PSA-coated material is adhered by
its adhesive layer to a primed polyethylene substrate with an adhesive contact
bond area of 1  0.5 in. When testing is carried out at an elevated
temperature, this substrate is first reinforced with aluminum foil to impart
rigidity.

Twenty Degree Hold. The 20 hold test is similar to a standard shear
test except that the test plate is inclined 20 from the vertical direction [37].
This test measures a combined peel and shear strength of the adhesive
mounted on a 1 mil Mylar film when applied under standard forces to a
corrugated cardboard substrate. Shear from cardboard according to PSTC 7
is tested with a 200-lb bursting strength cardboard. The tape is applied in the
corrugation direction [25]. With regard to carton sealing properties,
cardboard shear, cardboard adhesion (% fiber tear), and high humidity
aging (80% RH/125 F, 3 days) are evaluated [43].
Apparatus:
Shear tester block
500-g weight
Method:
Samples of adhesive-coated PET are applied on a cardboard
substrate such that a 1.5-in. edge of the sample is aligned parallel to
the corrugated flutes (ridges) of the substrate.
After application, the sample is rolled with a 4.5-lb rubber roller one
time parallel to the 1.5-in. edge of the sample at a roller speed of
12 in./min.
The sample is mounted in a shear tester at an angle of 20 to the
vertical.
A 500-g weight then is affixed and a timer is started.
Hold values are reported in minutes.
The shear test shows the following disadvantages: the information
obtained remains limited, the reproducibility is low, and the time required is
too long. More reliable results are given by creep tests (measuring a creep
time slope in order to determine the viscous and elastic components) or by
measuring the creep (cold flow) directly (see later).
Test Methods 675

Automotive PSAs Shear. This test measures the slip after a given
time, using a 500 g/in.2 test sample surface at 158 F, according to the Fisher
Body Procedure TM-45-134 [33].
Thick Adherend Shear Test. A suitable test for the measurement of
shear properties of an adhesive is the thick (6 mm, the overlap length is
5 mm) adherend shear test (TAST) since it is conducted using a tensile
testing machine. However, finite element analysis shows that the stress
distribution in the thick adherend shear test is not pure shear. The values of
maximum shear strain gained from the thick adherend shear test are not
comparable to those obtained from tests in pure shear [163].
Table 10.15 summarizes the experimental conditions used in shear
measurements; Table 10.16 lists the various test methods and cohesive
strength values.
It should be noted that the use of the measured shear values has to be
made in connection with the test conditions, the adhesive formulation, and
the other adhesive characteristics. Some years ago tack was discussed as a
cohesion related property. Later shear resistance has been considered as a
cohesion and cuttability related characteristic. Recent developments in TPEs
demonstrate that such generalized assumptions are not valid. The use of the
shear resistance as an index of the cohesion, more as an index of the
nonelastic component of the cohesion is not correct. On the other hand the

Table 10.15 Experimental Conditions Used for Shear Measurements

Sample
Dimensions Weight Temperature
(mm  mm) Substrate (g) ( C) Method Ref.

1 25  12.5 Glass 1000 RT AFERA4012-Pl [151]


2 25.4  12.7 Stainless steel 1000 70 PSTC 1 modified [151]
3 25.0  25.0 Stainless steel 1000 40 FTM 8 —
4 12.5  12.5 Stainless steel 500 — ASTM D-3654-78 [70]
5 12.5  12.5 Stainless steel 500 RT PSTC 7 —
1000
6 12.5  12.5 — 1000 RT — [35]
7 24.5  24.5 — 500 20 PSTC modified —
1000 65
8 12.5  12.5 — 500 RT PSTC 7 modified [162]
24 hr dwell time
9 10  50 — 1000 5 — —
Dynamic
10 645 mm2 Glass — RT 1 mm/min [70]
676

Table 10.16 Test Methods and Values for Shear

Test Condition

Dimensions Weight Face stock


Temperature (C ) Angle ( ) of the sample (mm) (g) material Value Units Norm/Method Ref.

1 RT 178 6.2  6.2 500 PET 120 min PSTC 7 [166]


2 RT 180 12.5  12.5 500 PET > 100 hr FTM 8 [167]
3 RT 180 25  25 10 N PET 0.7–168 hr FTM 8 [2]
4 RT 180 — — PET 250 min FTM 7 [24]
5 RT 180 — — Paper 36–162 min FTM 7 [168]
6 20 180 25  25 1000 PET — — — [59]
7 65 180 25  25 500 PET — — — [59]
8 70 180 25  25 1000 PET > 30 hr PSTC 7 [169]
9 RT 180 25  25 1000 PET > 85 hr — [2]
10 RT 180 25  25 1000 PET 5–10 hr — [147]
11 70 180 — — PET 0.5–2 hr — [170]
12 RT 180 25  25 1000 PET > 200 hr — [171]
13 70 180 25  25 1000 PET > 100 hr — [172]
14 RT 180 25  25 1000 PET > 200 hr — [173]
15 RT 180 25  25 1000 PET > 5000 min — [172]
16 70–90 180 25  12.5 — — — — — [86]
17 RT 180 6.45 cm2 — — 80–150 N Dynamic [70]

18 — 180 25  25 1000 PET 75–92 C SAFT [172]
Chapter 10
Test Methods 677

use of the cohesion as an index of tack is generally not possible. This


statement is illustrated by recent advances in the use of the Probe tack test
method.
Creep Test. For special, soft products creep tests are recommended
also. The apparatus used for the shear creep measurements suggested in
[164] was similar to that described by van Holde and Williams [165]. The
applied shear stress was 3.5  104 dyne/cm. The creep displacement was
measured with a capacitance probe transducer for a period of 100 min
following the application of the load.
Global Evaluation of the Adhesive Properties. A complete evaluation
of the adhesive properties is possible after measuring all performance
characteristics (i.e., tack, peel resistance, and shear resistance). There
are quite different graphical representations allowing the comparison of the
above properties as a function of the formulation and in relation to one
another. Figures 10.6, 10.7, and 10.8 present the main types of diagrams
used for the evaluation of the adhesive properties. Figure 10.6 shows
that the adhesive properties can be evaluated separately (one property)
for different formulations (A) or different properties for a given formulation
(B). Figure 10.7 presents an overview of the performance characteristics

Figure 10.6 The main types of diagrams used for the evaluation of the adhesive
properties. A) Adhesive characteristics for different formulations (peel, tack, and
shear). B) Adhesive characteristics (peel, tack, and shear) for a given formulation.
678 Chapter 10

Figure 10.7 The main types of diagrams used for the evaluation of the adhesive
properties. C) All performances as a function of the chemical composition. D) All
performances for a given chemical composition.

Figure 10.8 The main types of diagrams used for the evaluation of the adhesive
properties. E) Space diagram of the adhesive characteristics (peel or tack or shear)
as a function of the chemical composition and component characteristics. F) Plane
diagram of the adhesive characteristics.
Test Methods 679

for different chemical compositions (C) or for a given formulation (D).


Figure 10.8 shows that the properties can be plotted separately as a function
of the chemical composition and a base characteristic of the main
components, in a three-dimensional (E) or two-dimensional graph (F).
The general test methods of the adhesion include the tack, peel, and
shear resistance measurement. Such methods are used for adhesive coated
and adhesiveless PSPs also. Their principles and practice for PSAs have
been described earlier. As discussed, such methods use standard carrier
materials. It is evident that the adhesive properties of the finished product
may differ from the characteristics of standard samples, due to the influence
of the carrier material and to manufacturing conditions. On the other hand
there are PSPs where the adhesive is not coated but imbedded in the finished
product, or the carrier itself possesses an adhesive nature. For other
products the adhesive is a part of a composite structure having partially the
characteristics of a carrier (from the point of view of mechanical resistance).
For such products the general test methods will substantially differ from the
common methods used for PSAs [65,174].

Product Related Tests. For splicing tapes the peel and shear resistance
are measured. Peel resistance for splicing tapes is tested as standard peel
on stainless steel and as peel on siliconized liner [175]. The shear resistance
is tested as dynamical, statical, and practice-related shear. Both statical
and dynamical shear measurements are carried out at room and elevated
temperatures. Statical and dynamical shear measurements are carried out
according to modified standard methods with paper as substrate. The
practice-related shear test uses a heated cylinder. For packaging seal tapes,
falling and flap test (this is a combination of shear, peel adhesion, and tack)
have to be carried out also [176]. For insulation tapes mechanical, electrical,
and thermal properties are required. Elongation, shear, and peel resistance,
are required too. Shear resistance should be measure after solvent exposure
also. For UV light-cured PSAs a bending test is used [122]. A skin adhesion
test is carried out for medical tapes. The skin underlying the tape sample is
inspected visually, in order to evaluate the amount of adhesive residue left
on the surface of the skin. The samples are assigned a numerical rating from
0 to 5. By a removability test the following evaluation criteria are used:
A—% adhesive transfer (less than 5%), L—legging, SL—slight legging.
Such examples illustrate the specific, product related character of the
practical tests. Their detailed discussion is given in [65,174].

Rheological Evaluation of the Adhesive Properties. The interdepen-


dence of the adhesive properties with the rheological characteristics was
discussed in detail in Chap. 2, Sections 1.1 and 1.2, and Chap. 4, Section 2.
680 Chapter 10

Developments in the equipment available for the test of the rheological


properties have allowed the use of the rheological characteristics for
screening of the PSA formulations. The main parameters investigated are
the viscosity, the Tg, and the modulus. For their study viscosimetry, DSC,
and DMA are used.
For a decade dynamic mechanical analysis has tried to measure,
explain, and forecast pressure sensitivity on the basis of the rheological
characteristics of the polymers, taking into account the time-temperature
dependence of the material characteristics (modulus). Such a method
presents the following main disadvantages: it is limited to the adhesive,
it describes the behavior of ‘‘pure’’ polymers, without micromolecular
additives, and it is limited to the phenomena which occur in the bulk
polymer. In the classical domain of adhesive-based PSPs the actual limits
of a pure rheological approach are illustrated by the test and application of
crosslinked products. Natural rubber is a crosslinked product, however its
elastical properties are due mainly to chain entangling, and its crosslinked
network is very elastic. Advances in macromolecular chemistry allowed
the synthesis of other crosslinked polymers having rubbery elasticity. The
compounds with physical crosslinking (thermoplastic elastomers) are hot
processable and 100% solids materials. The compounds with chemical
crosslinking (e.g., carboxylated styrene-diene copolymers) are supplied as
water-based dispersions also. Both are relatively hard materials; the styrene
block copolymers are elastomers without self-adhesivity, the carboxylated
styrene-diene copolymers are products with a low level of self-adhesivity
(they need a high concentration of tackifier) due to the bulky styrene
monomer and to the tight crosslinked network. Such polymers behave like
filled systems. They cannot be characterized by common test methods of the
adhesive properties or of the rheological parameters. As discussed earlier
by the usability of the Probe tack test method for SBCs, for such
polymers discrepancies appear, due to the stiffness (rigidity) of the adhesive
which is the result of the special build-up of the network. Principally the
same problem, i.e., the insufficient control of the length, and the distribution
of the network bridges can hinder the characterization and use of
carboxylated styrene-diene copolymers, where the Tg and sol/gel content
alone do not describe the practical behavior (peel and shear resistance) of
the adhesive.
However, for formulations of ‘‘pure’’ macromolecular adhesives DMA
helps by the screening of the possible recipes. For instance, the mobility of
the midblock in TPEs can be evaluated by DMA. It is related to the softness,
viscous flow, i.e., to the loss modulus, to the situation of the rubbery plateau
modulus, and to the ratio of the two moduli (storage and loss moduli), i.e.,
tan . The situation of tan , related to temperature, the tan  minimum
Test Methods 681

temperature (Tmin), and the tan peak temperature (Tmax) characterizes the
rheological behavior. The melting, disappearing of the endblock domains is
indicated by the Tmin and the crossover temperature of the loss and storage
moduli (Tcross). Between Tmin and Tcross the end domains soften and
disappear. The difference between Tcross gives information about the melt
viscosity of the formulation. In tests of the release properties the general
shape of the release profile is consistent with the shape of the adhesive’s
viscoelastic function—specifically the loss tangent (tan ) as a function of the
dynamic frequency—in the peel window [118]. Advances in DMA examining
the whole product (adhesive and carrier) were obtained also. Willenbacher
[177], uses torsional resonance oscillation, to achieve the moduli G0 and G00
of the sample (PSA tape); two maxima are observed, corresponding to
the adhesive and to the carrier. As discussed earlier (Table 10.5), Williams
plasticity is related to the modulus values also. A complementary characteri-
zation for PSAs is allowed by the melt rheology. The study of the viscosity
and of the storage modulus allows a better characterization of the HMPSAs.
A detailed discussion of the DMA data for the formulation of the adhesive is
given in [178]. As discussed by Muny [117], the cost of this equipment is still
quite high, which limits its use to only larger organizations.

Other Properties
There are some properties of the coated adhesive which influence the
adhesive performance; they include the aging properties of the adhesive
and the coating weight. On the other hand, the adhesive and converting
properties together determine the application field of the label; therefore the
converting properties should also be tested.

Coating Weight. The coating weight influences the adhesive,


converting, and end-use properties. The dependence of the adhesive
properties on the coating weight was extensively discussed in Chap. 6.
Industrial in-line measurement and control of the coating weight, as well as
spot checks of the coating weight are carried out. The coating weight may
be checked using rays or IR radiation [157]. For laboratory purposes the
gravimetric method is generally used. In order to avoid the discrepancies
due to the sensitivity of the face stock material towards the liquid adhesive,
aluminum foil may be used. The coating weight is measured by dissolving
the adhesive layer and determining the weight difference between the coated
and uncoated face material. The method was described in Section 2.1. The
use of fluorescence techniques to determine coating weight of release is well
known [179].
682 Chapter 10

Aging Properties. Aging tests are carried out on liquid and dried
PSAs. For liquid PSAs these are mostly limited to testing the viscosity of
molten hot-melt PSAs. For the dried PSAs or laminates aging tests are
carried out in order to check the solid components of the laminate or the
PSAs (shrinkage, migration). These aging tests use one of the standard
methods for the characterization of the adhesive properties (or their
combination) after the storage of the adhesive (laminate) under well-defined
conditions (temperature, light, humidity) for a defined time. The most
important adhesive characteristics evaluated after aging are the peel adhesion
and the tack; generally, peel increases with time and tack decreases. A more
complex test concerns the aging of removable PSAs, where ‘‘clean’’ remov-
ability should be examined also. In some cases aging tests are carried out in
order to make a choice of the adequate adhesive. Aging tests for hot-melt
PSAs are carried out (melt viscosity stability) at 177 C for 4 days [180]. For
HMPSAs principally two different test series are carried out: color and
adhesivity changes at elevated temperature 150 C, 24–100 hrs) and the
standard adhesive properties after aging at elevated temperature (70 C, 1–5
weeks) are tested [181].
The tests of the aging characteristics of plastics have been summarized
[180,182,183]. In a similar manner adhesive-coated laminates are exposed
to weathering tests and their adhesive properties are tested subjectively or
with a mechanical device. Generally, the weaknesses of special classes of
adhesives are well known; hence aging test methods are developed to
examine these shortcomings after storage. Consequently, the conditions
proposed for accelerated aging (storage) can vary quite a bit. The following
environments were proposed for aging tests of PSA laminates [180]: 65 F
and 50% RH, þ73 F and 50% RH, þ150 F and 50% RH, and þ73 F and
95% RH. Temperature and humidity cycling (MIL STD 331) and diurnal
cycling from 65 to þ165 F at 95% RH are carried out. Test specimens are
examined after 0, 8, 16, 27, and 40 weeks exposure. Aging at 70 C for 1 day
corresponds to 0.5 yr of natural aging [169]. Aging tests are carried out at
room temperature, 54 C, or 71 C for 1 to 4 weeks [184].
For aging tapes, a storage at 80% humidity at 66 C for 6 days and at
50% humidity at 29 C for 4 hr according to PSTC 9 was proposed [185].
Exposed film tack may be evaluated by leaving tapes in a laboratory
environment and periodically checking the finger tack. Aging for 1 day or
1 week at room temperature, 8 days at 88 C, and 8 days under a sunlamp is
done for chloroprene latices [186]. The aging of SIS-based hot-melt PSAs
was tested after storage at 150 C for 24, 48, 72, and 100 hr [183]. Tests were
carried out at 70 C after 1, 2, 3, and 4 weeks [178]. Pure SIS films were aged
at 95 C for 150 min [186]. Adhesive build-up is tested for hot-melt PSAs
[187]. No significant increase in 180 peel adhesion to stainless steel was
Test Methods 683

noted from initial testing through 24 hr. With hot-melt PSAs, peel decreases
after aging. A hot-melt PSA with a 180 peel adhesion of 8.50 lbs/in.
(control) displays only 3.70 lbs/in. after 7 days at 120 F, and 4.70 lbs/in.
after 7 days at 150 F [43]. Aging of SIS-based hot-melt PSAs is performed at
95 C, on polyester, and in a dark environment and dry air [178]. As stated
by Kauffmann [188] the effect of antioxidants can be investigated by
accelerated aging only partially. The test procedures for the choice of
an antioxidant are based on the artificial aging of the product and IR
estimation of the carbonyl and hydroxyl absorbances.
The aging behavior may be estimated by measuring peel or tack
changes. Peel strength should be tested initially and after 14 days at 40 C, as
well as shear strength and rolling ball tack. Exposed film tack after days
at 70 C may be measured also. The UV stability of hot-melt PSAs is
tested after 0, 1, 3, and 5 hr via 180 peel, shear, tack, color, and density
measurements [185]. Table 10.17 summarizes the test conditions for aging.
The aging test is carried out by storing a laminate sample, 100 
100 mm, in an oven, between two wooden plates, under a load of 1 kg/
100 cm2, and at 70 C for 4 days. The degree of the adhesive degradation
during aging, and the migration (bleeding) are examined:
Evaluation of the adhesive degradation (finger tack test). The
degree of degradation is rated as follows:
D0 ¼ OK, no changes in adhesive properties.
D1 ¼ Slight legging, may be used however.
D2 ¼ About 1 mm legs, adhesion lowered.
D3 ¼ Fiber yield, adhesive transfer on the fingers.
D4 ¼ Smearing, fluidity of the adhesion, loss of cohesion.
D5 ¼ Dry, nontacky adhesive.
Evaluation of the migration (bleedthrough): after peeling off
the liner, coated and uncoated strips of the face material should be
compared. The quality is rated as follows [32]:
M0 ¼ No difference, very good.
M1 ¼ Coated surface generally lighter, fair.
M2 ¼ Migration, point-like sites, poor.
M3 ¼ Migration, numerous point-like penetration sites, pinhole-
like or more.
M4 ¼ More than 50% of the surface covered by bleeding.

Aging Test (not FTM). This method tests subjectively the adhesion
of the coated PSA layer after aging under light and temperature exposure.
For current, industrial control tests, a ‘‘30 min aging time’’ should be used.
684 Chapter 10

Apparatus:
Sun tester.
Oven, without air blow cooling (weatherometer), substrate
temperature 60–70 C.
Glass plates (slides) 2.5 mm thin, 25 mm wide.
Method:
A strip of PSA-coated material (25  100 mm or longer) is applied
to the glass slide, with the coated adhesive side against the light,
with a 20-mm portion at the end of the slide left exposed, the rest
being covered with release paper/aluminum foil.
After 20 min aging of the exposed part, the next 20 mm portion
should be exposed and so on.
After aging the adhesive quality (finger tack) of the aged material
should be subjectively evaluated.
D0 ¼ OK, no changes.
D1 ¼ No major changes, good-fair, slight legging.
D2 ¼ 1 mm legs, the adhesive is softer.
D3 ¼ Legging, adhesive transfer.
D4 ¼ Smearing, loss of the cohesion.
D5 ¼ Loss of the tack.
Accelerated Weathering Testing. One mil films of adhesive coated on
one mil PET film are exposed directly (adhesive side) to the light source in a
UV device. A straight UV cycle is imposed (no moisture) [191].
Aging Test of the Release Force. Accelerated aging before carrying
out a high speed release test can be carried out by placing a set of self-
adhesive strips between two flat metal or glass plates and keeping them
for 20 hr in an air-circulating oven at 70  5 C. The strips then should be
removed and conditioned for at least 4 hr. The environment’s effect on the
release charac- teristics of the release paper is measured.
Aged siliconized paper is applied to freshly prepared adhesives and
the force required to remove the paper measured in the 180 mode [180].
Adhesive film aging may be tested at 150 F and 15% RH at 4 and 12 day
intervals. Exposure under these conditions for four days according to
ASTM D-3611-77 represents two years of natural aging.
Aging of PVC Face Stock. Aging tests on PVC (coated with PSAs)
may be carried out after 1 week at 158 F [192]. This test appears equivalent
or more severe than 7.5 days exposure at room temperature [72]. Aging of
PSAs and protective agents are described in [193].
Test Methods 685

Table 10.18 Shrinkage Values for Different Face Stock Materials

Face Stock Characteristics Test Conditions Shrinkage (%)

Nature Thickness (m) Temperature ( C) Time (min) MD TD Ref.

PET 30–350 190 5 3.0 2.0 [194]


PET — — — 3.3 3.5 —
LDPE 30–200 70 60 — 1–3.0 —
HDPE — 80 15 1.0 2.0 —
CPP 100 90 5 1.0 — [195]
BOPP 50–75 130 5 <8 <4 [194]
BOPP — 135 7 5.0 5.0 [196]
PVC 40–200 70 60 — 0–2.0 —
Paper — — — 0.8–1 — [197]

Dimensional Stability. Dimensional stability testing was standardized


according to FINAT FTM 14. Shrinkage values for different face stock
materials are summarized in Table 10.18 (see also Table 6.42). Test condi-
tions and shrinkage of PSPs are discussed in detail in [198,199].

Plasticizer Resistance. Different practical tests are described in the


literature that deal with the plasticizer resistance of common adhesives
[200]. Most of them use a PVC film (soft PVC with a well-defined plasticizer
content) and store the PVC between the adhesive surfaces to be tested.
The plasticizer migration is estimated as a weight difference [201,202]. Most
plastic films are compounded materials, and contain additives. There are
possible interactions between micromolecular components of the film (e.g.,
PVC) and the adhesive. Plasticizers and emulsifiers from the film can migrate
into the adhesive. On the other hand, monomers, oligomers, and surface
agents from the adhesive can migrate, causing stiffening of the face stock, or
its shrinkage, and the loss of the adhesive properties. The insufficient aging
resistance of the rubber-based adhesives excludes their use on plastic films.
On the other hand, several EVAc-based adhesives show high shrinkage. Thus,
the most suitable adhesive materials for film coating are the acrylic PSAs.
The shrinkage of a plasticized vinyl film with a 1.0 mil transfer coated
adhesive film, conditioned 24 hr at standard humidity and temperature, is
measured as the change in length and width; then it is heat aged on the liner,
reconditioned, and measured again. Shrinkage on vinyl is tested after 7 days
at 158 F. The plasticizer migration resistance of the polymer is acceptable
for vinyl applications if the shrinkage, when heat aged on a release liner, is in
the range of 0.3–0.9% [38]. For transfer tapes used for structural bonding
686 Chapter 10

plasticizer resistance is tested at elevated temperatures (70 C, 30%


plasticizer) [203].
In the mounted shrinkage test a PSA-coated PVC face stock is bonded
to a test surface, often stainless steel, aluminum, and/or release liner. The
laminate then is exposed to elevated temperature accelerated aging, typically
70 C for one week, and the percentage shrink back of the PVC film from its
original dimensions is measured. Both machine direction (MD) (subject
to greatest stress-induced elongation) and cross-machine direction (CMD)
measurements are carried out [43]. In another test, a 4  4-in. coated sample
is adhered to a formica panel and aged for 7 days at 110oF [38]. Evaluation
criteria include shrinkage in the machine direction of less than 0.5%, and
shrinkage in the cross direction also below 0.5%.
Plasticizer migration is tested by 180 peel after 24 hr dwell time, 24 hr
aged peel, and 1 week aging on the liner at 158 F. The peel retention after
7 days at room temperature and 158oF, the shrinkage on the liner (6–8%),
and mounted shrinkage (11–17%) were measured (24 hr at 158oF) [154].

Migration. When pressure-sensitive adhesives are coated onto porous


substrates such as paper, an assessment of the tendency of the adhesive to
migrate or bleed through the paper should be made (see also Chap. 7). The
migration of the adhesive can discolor the face surface and reduce the
effective adhesive film thickness, thus changing various adhesive properties.
On the other hand, the migration may dilute the adhesive and decrease
the adhesive properties. Migration is tested by storing the label samples at
increased temperatures and then examining them against a black cardboard
[204]. Suggested storage temperatures are 60 C and 71 C for hot-melt and
solvent-based PSAs, respectively.
There are different methods to test the bleeding through of an
adhesive.The test can be run by placing samples of coated label stock, based
on paper of a certain quality (e.g., litho paper), into an oven kept at elevated
temperatures (minimum of 70 C). The test samples are inspected for
discoloration at weekly intervals [195]. Migration for splicing tapes is tested
according to the well known method, i.e., by statical load of multilayer
samples at different temperatures and times [205].
Face stock penetration is tested using a 60-lb white paper. The
adhesive is coated on siliconized paper and laminated with the Kromecoat
litho paper [72]. The adhesive laminate is stored in a 158 F oven for specified
times (5 days). Bleedthrough should be tested under a load of 1 lb/in.2. No
migration at 140oF on standard label stock should occur [43].

Edge Ooze. This property refers to the flow characteristics of the


coated adhesive.
Test Methods 687

Method:

A sample (a stack of sheets, having a mean height of 3 cm) should


be stored at room temperature for 24 hr.
After storage the sample should be guillotined through the middle.
The appearance of the cut surface should be evaluated subjectively.
The evaluation of the cuttability (edge ooze) should be repeated
after nine succesive cutting operations of the sheet stack.
Rating:

1 Very good, no blocking, sheets can be moved (from the sheet


stack).
2 Slight bonding on the cut surface, sheets can be moved only after
separating the cut pieces.
3 Pronounced adhesive residue on the cut surfaces, slight gumming
of the cut and cutting surface, sheets cannot be moved (from the
sheet stack).
4 Very pronounced flow-out of the adhesive, very pronounced
gumming on the cut and cutting surface.

3.2 Special Laminate Properties


In this section the test methods designed to evaluate the special features of
PSA labels used in quite different application fields (e.g., removable film,
deep freeze) are covered.

Removable PSA Labels


The most important special features of removable PSA labels are the
removability, the migration/bleeding behavior, and the edge lifting
properties.

Removability. In the automotive industry special removable tapes


have to display very high removability. Thus, selected tapes were removed
cleanly from enameled, lacquered, or melamine test surfaces after 1 hr at
150 C or 30 min at 121 C. For common PSA labels the experimental
conditions are not so severe, but generally instantaneous removability and
removability after aging (room temperature and high temperature aging)
are carried out. Removable adhesives can be formulated so as not to
display adhesion build-up (i.e., the adhesion to the substrate does not
increase to the point where the label cannot be removed cleanly even
after exposure to heat) (see also Chap. 6). The most important tests
688 Chapter 10

covering removability are the following:


removability from different surfaces, after aging at elevated
temperatures over a period of time, evaluated subjectively (see
Table 6.18);
removability from glass, after aging at elevated temperatures,
evaluated as peel from glass;
mirror test, residue-free removability from mirror glass, evaluated
subjectively.

Removability of Hot-Melt PSAs. The removability is tested by aging


the stainless steel panels with 1  6-in. strips of label stock at room
temperature and elevated temperature (48 C), cooling for 1 hr at room
temperature, followed by peel adhesion measurements of the label stock
[206]. Other test cycles include 24 hr at room temperature, 24 hr at 48 C,
1 week at room temperature, and 1 week at 48 C. For hot-melt PSAs
the initial peel from stainless steel should be generally below 1.0 lb/in. or
16 oz/in. The resulting peel (after build-up) should be below 2.5 lbs or 40 oz
[196]. For removable hot-melt PSAs the initial peel, 24 hr peel at room
temperature, peel adhesion at 48 C, 1 week peel at room temperature, and
1 week peel after aging at 48 C are measured. Tack values for removable
hot-melt PSAs (90 quick stick) are 0.5–1.3 lb/in., rolling ball tack values
are 1–3 in., and peel values are 0.9–2.6 lb/in.; the SAFT amounts to
128–156 F [206].
Not only the peel value but also the failure mode is important in the
removability test of PSAs. Therefore, the amount of adhesive transfer for
peelable tapes or labels is evaluated after storage at 40 C for 7 days [20].
Removability of Water-Based PSAs. The evaluation of removable
water-based acrylic PSAs is carried out after 1 week storage at 70 C on PET
and HDPE [2].
Adhesive Residue. When the removability test (described earlier)
is performed, the surface (substrate) underneath the label sample is usually
inspected to determine the amount of adhesive residue left on the surface of
the adherend. There are different, subjective rating scales. In some cases
each sample is assigned a numerical rating from 0 to 5 based on an arbitrary
scale. In other cases adhesive transfer is estimated with A, L, and SL ratings,
where A means adhesive transfer, L corresponds to legging, and SL denotes
slight legging or stringing. High temperature resistance may be tested as
the percentage adhesive transfer after 30 min at 250oF [43]. Removability
should be evaluated on glossy stain-resistant acrylic enamel after aging at
elevated temperatures and peeling at an angle of 45 .
Test Methods 689

Apparatus:
Test panel covered with glossy stain-resistant acrylic enamel paint
(automotive paint).
Samples of labels (tapes) 1.27  10.16 cm adhered at room tempe-
rature to the test surface.
Ventilated oven.
Method:
Samples are fixed to the panel.
A 2-kg rubber-coated metal roller is run twice over the samples.
Samples are left for 1 hr at 128 C or 150 C in a ventilated oven.
Samples are peeled back at an angle of 45 at an approximate rate
of 1.9 m/min.
The panel is removed from the oven and examined. A nontacky
deposit is reported as residue.

Repositionability. With repositionable adhesives the adhesive-coated


paper may be readily lifted and removed from the contact surface and
reapplied at least eight additional times to paper surfaces without reducing
the adhesive properties [207]. Measurements done with a 7-g/m2 adhesive coat
weight, on a 70-g/m2 paper as face stock material, yield the following peel
adhesion values on newspaper (test conditions: 180 peel, 300 mm/min): 110 g
initially, 85 g after 50 repeated peel tests, and 75 g after 100 repeated peel tests.
Readherability. The readherability of the PSA on paper has been
tested by measuring of the peel after repeated lamination and delamination
(50–100 steps) [208].
Lifting (Mandrel Hold). Different lifting tests are carried out in
order to check if the peel/shear resistance of the adhesive balances the lifting
forces of the face stock (deformation of the face stock material on a curved
surface). In a test called the ‘‘Curved Panel Lifting Test at 150 C,’’ an
aluminum panel with a radius of curvature of 23 cm and a length of 35.5 cm
in the curved direction is used. The tapes to be tested are applied to the
aluminum panel in its curved direction. The assembly-bearing panel then is
put into an air-circulating oven at 150 C for 10 min, allowed to cool, and
examined for failures. A rating of ‘‘pass’’ means no lifting has occurred; any
lifting at either end of the strip is noted as ‘‘end failure’’ and the total length
of tape which has lifted is rated.
This method should be used as a practical test for the peel of
removable PSAs. The peel is subjectively evaluated on the basis of lifting
or flagging of the sample affixed on round, cylindrical items.
690 Chapter 10

Apparatus:
H-PVC and polyethylene cylinders (21 mm).
Glass cylinders (16 mm).
Method:
Two samples (labels 15  50 mm for PVC, and 15  40 mm on
glass) are applied in such a manner that the machine direction
of the face is either parallel or perpendicular to the axis of the
cylinder.
The samples are stored at room temperature.
The lifting (delaminated) length of the samples is measured at
regular intervals (i.e., 1 day, 1 week).
Lifting (winging) also is evaluated after 5 hr on vertical surfaces.

Filmic PSA Labels


In film-coating of PSAs, a range of special requirements must be fulfilled.
The first screening of the adhesive (and laminate) covers a lot of special tests
carried out in the laboratory or on a pilot coater (Fig. 10.9). Laboratory
tests carried out with water-based PSAs estimate the coatability and the
adhesive properties of PSAs. An additional aging test completes the labo-
ratory screening. In the coatability test the wet-out is checked for minimum
viscosity and minimum wetting agent level. In the preliminary test of
the adhesive properties, the peel, tack, and shear balance are estimated. The
preliminary evaluation of the adhesive properties includes the estimation of
the shrinkage, color, and storage behavior. In the pilot phase of the test,
the coatability, convertability, adhesive, and aging properties of the
laminate and its processability are examined. The coatability should be
tested by wet-out, through the rheology, and optical appearance. The
convertability of the dispersion should be checked as a function of the speed,
foam formation, and behavior on the metering roll. The discrepancies
between laboratory-measured and pilot-manufactured label material should
be evaluated with regard to the adhesive and aging properties. The
processability of the laminate should be examined as far as cuttability,
release, and storage are concerned. Figure 10.10 summarizes the steps of
the preliminary screening on a laboratory and pilot scale for PSAs for film
coating, including: coagulum, transparency, wet-out, shrinkage, tack/
polyethylene peel, water whitening, wet anchorage, and wet adhesion.
For water-based PSAs for transfer coating onto soft PVC, the
following properties are to be tested for their emulsion and adhesive
properties, as well as for their water whitening and water resistance.
Test Methods 691

Figure 10.9 Screening test for PSA for film coating. Evaluation of the suitability
(in %); A, adequacy; slr, solventless release; sbr, solvent-based release.

Emulsion Properties. The relevant emulsion properties are:


Viscosity: the viscosity should be measured with a DIN 53211
Ford cup of 4 mm, at a temperature of 20 C or with a Brookfield
viscosimeter, LVT, spindle 3, 12 rpm. Target values are 17.5 sec 
1 sec or 150–200 mPa s.
Solids content: the solids content should be measured according to
DIN at 140 C, over a period of 20 min, through weight difference
measurements. The target value should exceed 50%  1%.
Wet-out: the aqueous PSAs should be coated with an Erichsen
knife coater (70 m) on solventless silicone release paper at
692 Chapter 10

Figure 10.10 Screening test used on laboratory and pilot scale for the evaluation
of the general characteristics of a PSA film laminate.

a coating weight of 20 g/m2 or less, dry. Wetting out is ade-


quate if the wet PSA layer displays no transversal shrinkage within
10–15 sec.

Adhesive Properties. The most important adhesive properties


include:
Peel from glass: the PSA layer is coated using an Erichsen 70-m
knife; the target value is 10 N/25 mm.
Peel from polyethylene: the strip should be affixed on a
polyethylene film laminated onto a test panel; the target value
is 6 N/25 mm.
Rolling cylinder tack: as discussed in Chap. 6, the rolling cylinder
tack test measures tack as a function of the distance traveled by a
steel cylinder on an adhesive-coated substrate after it has rolled
Test Methods 693

down an inclined plane. The target value averages 2.5 cm  1 cm,


for a coating weight of 20 g/m2 (dry).
Hot shear: the hot shear should be measured according to a
modified FINAT FTM 8. Face stock materials include PVC, PE,
PP, or PET. For PVC materials the film should be reinforced with
paper in order to avoid elongation phenomena. As equipment a
special heated panel should be built, tilted 2 from the vertical
to which the steel plate with the sample on it can be attached.
When the temperature reaches 50 C, the free end of the suspended
sample should be loaded with a 2-kg weight. The longer the sample
holds, the better the hot shear resistance. The target value is a
minimum of 20 min.
Water resistance may be proved using different test methods. As is
known, the water resistance of conventional adhesive films is measured
by the elution rate and transparency of the film, immersed in water.
Water whitening:
The PSA is coated onto a 100-m film by the aid of an Erichsen
knife coater (70 m). After 30 min room temperature drying
supplemental drying should be carried out at 70 C for 10 min.
The sample is then covered with release paper. For comparison
purposes the face stock should possess ungummed areas also.
A coated (partially ungummed) strip of the film is immersed
in distilled water at room temperature. The moment color changes
appear (whitening of the adhesive layer) should be noted. The
longer the transparency persists, the better the water-whitening
resistance. Target values are rated as follows: 10 sec, fair; 15 sec,
good; and 17 sec, very good. In a similar manner comparison of
the water sensitivity was carried out recently by submerging slides
covered in latex into water and measuring the % transmittance
over time. Water whitening was tested as % water absorbed after
24 hours, and the % transmittance at 450 nm after 24 hour water
soak [209].
Loss of transparency: PSA-coated film can suffer water whitening
upon immersion in water. Water whitening is characterized by
the speed of the loss of transparency (the water-whitening test just
described) and by the absolute value of the loss of transparency, as
compared to the transparency of a dry coating.

The film laminate should be covered with a transparent (clear) PVC


film and a bubble-free film/adhesive/film laminate should be prepared.
Another part of the same film coating (clear film þ adhesive) should be
694 Chapter 10

immersed in distilled water at room temperature for 7 min. After 7 min the
immersed sample should be laminated with the same transparent (clear) film
as used for the dry laminate. Transmittance (transparency) of both samples
(film-dry adhesive-film and film-wet adhesive-film) should be measured
comparatively with a colorimeter. The differences in transparency between
dry and wet laminate are measured; target values are < 8% ¼ fair
and < 3% ¼ good.
Wet anchorage: film coatings should not display rub-off of the
adhesive layer (hand made test) after immersion in water (7 min).
Wet adhesion on glass (see Chap. 6): a release-free film coating is
immersed in water (room temperature, distilled water) for a period
of time of 7 min. Then the wet film coating is affixed on glass
(bubble free and not floating). The adhesion of the wet adhesive on
glass (the recovery of the adhesive properties over time) should be
tested, by trying to remove the sample after different storage times
(i.e., 1, 2, 5, 10, 15, 30, and 60 min). The shorter the time for
adhesion recovery, the better the wet adhesion. Adequate values are
less than 10 min; at least one test value at more than 30 min is
required.
Wet adhesion on polyethylene is important for plastic bottles. High
speed label application should be used for plastic bottles based on
polyethylene copolymers. Affixed labels should resist immersion
in water or surface-active agent solutions.
Dry samples should be affixed on squeeze bottles. After 24 hr the
labeled bottles are immersed in a diluted solution of surface-active
agents (SAA) (0.5–1% common detergents, room temperature).
Labels are peeled off after a wet storage for another 24 hr; peel
forces are evaluated subjectively. No edge lifting is allowed.
Wet samples are tested in the same way as dry samples, but wet
labels (labels immersed for 7 min in water) should be used.
Summary of Adhesive Properties.
1. Peel adhesion:
The FINAT FTM 1 method could be simplified, but at least
instantaneous and 24-hr peel values on a standard surface are
needed.
At least one measurement from glass or stainless steel is needed.
Instantaneous polyethylene peel and 24-hr peel are the most
important.
If values are similar, peel on solid polyethylene plates should be
measured.
Test Methods 695

2. Tack:
For similar materials rolling ball tests should be repeated with the
rolling cylinder test.
Loop tack on polyethylene is more important than rolling ball tack.
3. Shear:
If the measured shear values are too high, smaller size samples,
samples with a higher coating weight, or hot shear measurements
are to be used.
Films can be reinforced with paper in order to avoid elongation
effects.
Tack/peel requirements for sheet/roll materials differ. Roll labels for high
application speeds need a higher aggressive tack and polyethylene peel. Film
sheets need to display a better cuttability (see Fig. 10.11). Figure 10.12
summarizes the relevant test methods used for the evaluation of the adhesive
properties as a function of time, temperature, and substrate.

Water Resistance
The water resistance is generally tested in order to characterize the water
resistance/insolubility of a PSA laminate. Water resistance/insolubility is
required when using labels under humid or wet conditions. For this
application the appearance of the label and its adhesion/end-use properties
under wet conditions are to be examined. With regard to water resistance,

Figure 10.11 Requirements for PSA for roll/sheet labels.


696 Chapter 10

Figure 10.12 The main test methods used for the evaluation of the adhesive
properties. RB, Rolling ball; RC, rolling cylinder; LT, loop tack; ST, steel; GL, glass;
PE, polyethylene film; PEP, polyethylene plate; HS, hot shear; CCSH, corrugated
cardboard shear.

paper and film labels should be included in the evaluation; film-based


products require special water resistance also. In areas such as solar and
safety window films, graphics, point of purchase advertising, and clear labels
the films may be applied to the glass by first wetting the adhesive and glass
surface to allow positioning. The water is then removed from between the
adhesive and the glass with pressure. During this wet application step the
adhesive must resist hazing and become optically transparent rapidly after
removal of the water. Water solubility is required for wash-off labels, mainly
based on paper face stock. Water resistance can be examined as the change
of the adhesive properties under the influence of water (or water vapor),
or as the change of the optical and dimensional characteristics of the label
(tape) under the influence of water. The interdependence of the parameters
characterizing the water resistance performance of a PSA or PSA laminate
was discussed in Chap. 8. For certain applications a high water resistance
(i.e., no influence of water on the PSA laminate) is required. For others a
strong influence of water on the laminate is desired. One should differentiate
Test Methods 697

the water resistance from the humidity resistance. Water-resistant products


are special laminates; humidity resistance is generally required for most
laminates. The evaluation of the water resistance, measured as the stability
(or change) of the adhesive properties under the influence of the humidity,
is mainly based on the examination of the peel adhesion, tack, or shear after
storage in a humid environment.
Water resistance can be evaluated by storing the laminates under
water and removing them 24 hr prior to testing, to allow recovery of the
adhesion. According to [210] the resistance to water or solvents is tested by
immersion in the liquid for periods up to one week. Water resistance was
evaluated immediately after 24-hr water immersion or after a 7-day water
immersion, followed by a 24-hr recovery period [154]. Another test measures
water resistance after a 24-hr immersion, bonded to stainless steel for
30 min, then placed in room temperature tap water, followed by a 180 peel
measurement 24 hr later [38]. In the Weyerhouser (shear) test a laminated
Kraft paper strip (coated with PSAs) is immersed in water with a 350-g load
attached; the time to failure is measured. Water-whitening tests measure the
change of the optical appearance (see Section 2.2 of Chap. 6).
A clear Mylar label attached to stainless steel and immersed in water
at 74 F for 48 hr shows neither whitening nor blushing [20]. There is an
obvious objection to water whitening as a measure of water resistance.
Highly water-sensitive materials like gelatin or soluble cellulosic films are
not whitened at all. However, in water insoluble and slightly swollen latex
films, the degree and speed of whitening are a good index of water
sensitivity.

Humidity Resistance. As discussed in Section 1.2 of Chap. 6 and


taking into account the strong influence of the environmental conditions on
the adhesive properties of PSA laminates, test conditions are standardized
and a standard value of 50% RH is proposed. Testing humidity resistance
simulates real application environments, with increased humidity. The
humidity resistance is tested for 7 days at 100 F, 85% relative humidity, and
under condensing humidity conditions [154]. The humidity resistance was
measured for PSAs coated on a 2 mil PET film, bonded to stainless steel for
30 min, then placed in a humidity cabinet at 100 F and 95% RH. After
7 days a 180 peel test is carried out [38]. In other tests wet aging under
100% RH, at 40 C, for 240 hrs was carried out. Measurement of the
hydrophobicity of special latexes was performed by measuring the contact
angle of a drop of water placed on paper impregnated with the latexes. The
contact angle of the latex film produced with the nonmigrating surfactant
not only had a higher original contact angle (125 vs. 80 ) but did not drop
even after several minutes [96].
698 Chapter 10

Water Removability. Labels have to adhere in the presence of atmos-


pheric moisture up to about 30 C, but should easily be removed by washing
at temperatures of 45–80 C. Water removability is measured by applying
paper (8.2  3.7 cm) carrying 20 g/m2 of the dry composition to a glass bottle
and then agitating the bottle gently under water at 60 C. The label should
easily detach from the glass, with the PSA being carried on the label
and being nontacky in the presence of water. When submerged in tap water
at 60 C the label removed itself under 1 min [211]. Water dispersibility
requires, that the paper used as carrier material has to be dispersible
and the adhesive water soluble, but resistant towards organic solvents.
Water dispersibility is tested at different pH values.

Wash-Off Adhesives. Wash-off adhesives are designed to easily


dissolve in water. The test method can be described as follows: the water
removability is measured by applying a coated paper (8.2  3.7 cm) with
20 g/m2 dry adhesive to a glass bottle, and then agitating the bottle gently
in water at 60 C. The label should easily detach from the glass with the
adhesive film still attached to the label, and being nontacky in the presence
of water [70]. The water solubility should be tested for affixed labels as old
as 3–6 months, stored at room or elevated (60–70 C) temperatures.

Washing Machine Resistance. Special labels of fabric (textile) used


on clothing have to be tested for washing machine resistance. Such labels
should resist more than 50 wash cycles and up to 10 chemical (dry) cleaning
process cycles [212].

Deep Freeze Labels


A great deal of attention has been paid to the performance of PSAs at low
temperatures as adhesive failure at those temperatures may occur [213]. The
surface temperature of the adherend is another factor. Most general purpose
adhesives are formulated for tack when applied to surfaces with
temperatures as low as 38–40 F. If the temperature of the adherent is
lower, a higher degree of adhesive cold flow is required to provide proper
wet-out. Little concern needs to be given to products that will be labeled at
room temperature and subjected to lower temperatures later. In general, low
temperature properties are tested as a function of the application field,
but all of the adhesive properties have to be checked, in order to make a
meaningful evaluation of the low temperature properties. For comparative
purposes, static shear, peel adhesion, dynamic lap shear, and low
temperature lamination were evaluated [161]. The measurement of the
peel adhesion and of the tack at 5 and þ5 C as well as at 20 C give a
Test Methods 699

reasonable indication of the end-use performance, both for low temperature


(i.e., chill) and deep freeze applications.

Static Shear Load Test at Low Temperatures. In the static shear load
test at low temperatures a modified version of the PSTC test (1 in.2 contact
surface on stainless steel, 1000-g weight) was suggested [161]. Slippage and
delamination of the PSA-coated strips after five days storage in a constant
temperature cold chamber were observed. For dynamic lap shear testing at
low temperatures, 3  1-in. samples are coated with adhesive, equilibrated
in an environmental chamber at the desired temperature, and pulled apart
in a 180 configuration at a speed of 1 in./min.

Low Temperature Adhesion (Peel). For the evaluation of cold peel


adhesion, samples are applied when enclosed in an environmental chamber
and allowed to equilibrate at the desired temperature for 10 min prior
to being peeled in a 180 configuration at 12 in./min [161]. The temperature
at which a PSA label is laminated to a substrate can be just as critical
as the usage temperature. Therefore, in low temperature lamination tests,
adhesive-coated strips and stainless steel panels are conditioned at 20 C
for 30 min, then laminated together at this temperature. After warming the
samples to 23 C (room temperature), 180 peel adhesion values are mea-
sured [161]. According to common tests a 60-lb paper label and an adherend
surface are put into a test chamber at 35 F for 30 min. Labels are applied to
various surfaces, including a high density polyethylene (HDPE) panel, a poly-
ester container, and a corrugated board; other test conditions also are
suggested [72]. After a dwell time of 1 hr at 35 F, the labels are removed
by hand from the surfaces at an approximately 90 angle and a rate of 10–
12 in./min. Labels that tear upon removal are considered satisfactory [72].

Cold Tack. Cold tack is measured by attaching a label strip at the


given temperature to a cardboard surface adjusted to the same temperature.
The strip is applied with a 100-g roller and the laminate stored for 30 min
at the required temperature in the deep freezer. Afterwards the label strip
is removed slowly by hand under a defined angle and the percentage of
cardboard fiber tear estimated. FINAT FTM 13 describes the standard
method for low temperature performance.

Special Laminate Properties


Special laminate properties such as specific adhesion, environmental
resistance, roll storage aging, and release force from liner may be tested
for certain applications. Some of these performance characteristics and the
corresponding methods used were described previously. Specific adhesion is
700 Chapter 10

the ability of the adhesive to develop a useful bond level on a particular


surface to which it is applied. In most cases polyethylene adhesion is
checked. Environmental resistance is the retention of bond strength and/or
other properties upon exposure to water, high humidity, heat, cold,
chemicals, sunlight, etc. Roll storage aging is the resistance to adhesive
property changes upon long term contact with the face material (and liner).
Migration of plasticizers from PVC face materials, of slip agents, or the
nature of surface agents, for example, can have a severe effect on some
adhesives and relatively little on others.
The release force from the liner is the adhesive force to the liner deter-
mining possible predispensing of labels; it should not be so tight as to cause
web break or liner tear.

REFERENCES

1. P. Caton, European Adhes. Seal., (12) 19 (1990).


2. National Starch Chem.Co., Durotak, Pressure Sensitive Adhesives; Technical
bulletin, Zutphen, The Netherlands, 4/1986.
3. British Petrol, Hyvis, Technical bulletin, 1985.
4. R. Jordan, Coating, (2) 37 (1986).
5. BASF, Prüfmethoden, Polymerdispersionen/Polymerlo¨sungen, Bestimmung der
Auslaufzeit, PM-CDE 005d/July 1979, Ludwigshafen, Germany.
6. H. Hanke, Coating, (9) 261 (1986).
7. Allied Colloids, Viscalex HV 30, TPD, 6004 G.
8. W.R. Dougherty, 15th Munich Adhesive and Finishing Seminar, Munich,
Germany, 1990, p. 70.
9. W. Weilen, H.F. Fink, O.Klockner and G.Koerner, Coating, (10) 376 (1987).
10. ASTM D-2578-65; Standard Test Method for Wetting Tension of PE and PP
films.
11. K. Templer and F. Wultsch, Wochenbl. f. Papierfabr., (11/12) 483 (1980).
12. E. Prinz, Coating, (10) 244 (1979).
13. A. Fau and A. Soldat, ‘‘Silicone Addition Cure Emulsions for Paper Release
Coating,’’ in TECH 12, Advances in Pressure Sensitive Tape Technology,
Technical Seminar Proc., Itasca, IL, USA, May 1989, p. 7.
14. G. Habenicht, M. Baumann, R. Penzl and K. Jalal, Adhäsion, (6) 17 (1988).
15. G.Pitzler, Coating, (6) 218 (1996).
16. R. Heusch, Fette, Seifen, Anstrichmittel, (11) 123 (1969).
17. M. Osterhold, M. Breuchen and K. Armbruster, Adhäsion, (3) 23 (1992).
18. R.J. Roe, J. Phys. Chem., (6) 2013 (1968).
19. Coating, (2) 39 (1970).
20. Surfynol Technical Bulletin, Air Products and Chemicals Inc., 1991.
21. Byk-Mallincrodt, Technische Information, Entschäumer, 1990.
22. GAF, International Speciality Products, Wayne, NJ, USA, Technical bulletin,
7583-003 Rev. 1, p. 5, 1990.
Test Methods 701

23. Unibasic, Emulsion Polymerization, IV-VI, MD, USA, p. 18, 1990.


24. Product and Properties Index, Polysar, Arnhem, The Netherlands, 02/1985.
25. Pressure Sensitive Adhesives, Technical Bulletin, Rohm & Haas, 1986.
26. Adhes. Age, (5) 32 (1985).
27. Kaut., Gummi, Kunststoffe, 39, (1) 60 (1986).
28. P.K. Dahl, R. Murphy and G.N. Babu, Org. Coatings Appl. Polymer Sci.
Proc., (48) 131 (1983).
29. Adhäsion, (10) 307 (1968).
30. R.P. Mudge (National Starch Chem. Co., Bridgewater, NJ, USA), EPA
0225541/11.12.1985.
31. J. Johnston, Adhes. Age, (12) 24 (1983).
32. H. Müller, J. Türk and W. Druschke, BASF, Ludwigshafen, Germany, EP
0.118.726/11.02.1983.
33. P. Tkaczuk, Adhes. Age, (8) 19 (1988).
34. S.D. Tobing, O. Andrews, S. Caraway, J. Guo, A. Chen and Shelly Anna,
‘‘Shear Stability of Tackified Acrylic Emulsion PSAs,’’ Proc. 24th Annual
Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.273.
35. A. Midgley, Adhes. Age, (9) 17 (1986).
36. Coating, (3) 360 (1985).
37. Ashland Oil Inc., Chemicals, Bull. No.1496-1, 1982
38. D.G.Pierson and J.J.Wilczynski, Adhes. Age, (8) 52 (1980).
39. T.H. Haddock (Johnson & Johnson, USA), EPA 0130080 Bl/02.01.1985.
40. P.A. Mancinelli, ‘‘New Development in Acrylic HMPSA Technology,’’ in
TECH 12, Advances in Pressure Sensitive Tape Technology, Technical Seminar
Proc., Itasca, IL, May, 1989, p.165.
41. C.P. Iovine, S.J. Jer and F. Paulp (National Starch Chem.Co., Bridgewater,
USA), EPA 0212358/04.03.1987, p. 3.
42. Tappi J., (9) 105 (1984).
43. Technical Information, LHM, Celanese Resins Systems, 991.
44. M. Schlimmer, Adhäsion, (4) 8 (1987).
45. G. Simon, Adhäsion, (4) 98 (1974).
46. O. Hahn and Xiao Su Yi, Adhäsion, (7/8) 25 (1986).
47. D.J.St. Clair, Adhes. Age, (11) 23 (1988).
48. A. Roos and C. Creton, ‘‘Adhesion of PSA Based on Styrenic Block
Copolymers,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001,
Williamsburg, VA, p.371.
49. M.M. Feldstein, Polym. Mat. Eng., 81, 427 (1999).
50. R. Wilken, Finat News, (1) 53 (1988).
51. R. Wilken, Coating, (8) 212 (1982).
52. N.C. MacDonald, Adhes. Age, (2) 21 (1972).
53. D. Symietz, Adhäsion, (11) 28 (1987).
54. H. Mohle, Adhäsion, (6) 260 (1966).
55. C. Watson, European Adhesives and Sealants, (9) 6 (1987).
56. M.A. Johnson, Radiation Curing, (8) 4 (1980).
57. Coating, (1) 29 (1968).
702 Chapter 10

58. Prüfung von Haftkleber, D-EDE/K, BASF, Ludwigshafen, Germany,


Juni/Juli, 1981.
59. Coating, (1) 16 (1984).
60. K. Taubert, Adhäsion, (10) 377 (1970).
61. European Standard, Entwurf, prEN, T-peel test, CEN, Brussels (1994).
62. TLMI Manual, 1994, p. 45.
63. J.R. Wilken, Allg. Papier-Rundschau, (5) 122 (1986).
64. P. Dunckley, Adhäsion, (11) 19 (1989).
65. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 8.
66. R.P. Muny, Adhes. Age, (12) 18 (1986).
67. Adhäsion, (6) 24 (1982).
68. W. Druschke, Adhäsion, (5) 30 (1987).
69. N.L. Williams, R.H. Plaut and D.A. Dillard, ‘‘Analysis of the Loop Tack Test
for PSAs,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society, Myrtle
Beach, SC, Febr.20, 2000, p.252.
70. A. Kenneth, J.R. Stockwell and J. Walker (Allied Colloids), EP 0.147.067/
29.11.1983.
71. P. Hammond Jr., ASTM Bulletin, 360(5): 123–133.
72. L. Krutzel, Adhes. Age, (9) 21 (1987).
73. H. Green, Ind. Eng. Chem. Anal. Ed., (13) 632 (1941).
74. K. Kamagata, T. Saito and M. Toyama, J. Adh. Soc. Jap., (6) 309 (1969).
75. F. Wetzel, ASTM Bulletin, (221) 64 (1957).
76. A. Chiche and C. Creton, ‘‘Role of Surface Roughness of the Adherent Surface
on the Debonding Mechanism of PSA,’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr.25, 2001, Williamsburg, VA, p.187.
77. J.C. Hooker, C. Creton, P. Tordjemann and K.R. Shull, ‘‘Surface Effects on
the Microscopic Adhesion of Styrene-Isoprene-Styrene-Resin PSAs,’’ Proc. of
the 22nd Annual Meeting of the Adhesion Society, Panama City Beach, FL,
Febr. 21, 1999, p.415.
78. P. Tordjeman, E. Papon and J.J. Villenave, J. Polymer Sci. Polym. Physics,
Part B, 38, 1201 (2000).
79. N. Willenbacher and A.E. O‘Connor, ‘‘Effect of Molecular Weight and
Temperature on the Tack of Model PSAs from Polyisobutene, ‘‘Proc. 25th
Annual Meeting of Adh. Soc., and the Second World Congress on Adhesion and
Related Phenomena, Febr.10, 2002, Orlando, FL, p.378.
80. C.M. Flanigan, A.J. Crosby and K.R. Shull, Macromolecules, 32, 7251
(1999).
81. Y.Y. Lin and C.Y. Hui, ‘‘Modeling the Failure of an Adhesive Layer in a Peel
Test,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg,
VA, p.230.
82. K. Kendall, ‘‘Molecular Adhesion and Elastic Deformations,’’ Proc. 24th
Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.11.
83. K.L. Johnson, Contact Mechanics, Cambridge University Press,
Cambridge, 1985.
Test Methods 703

84. H. Lakrout and C. Creton, ‘‘Probe Tack Tests of PSAs with Flat and
Spherical Punches,’’ Proc. of the 22nd Annual Meeting of the Adhesion Society,
Panama City Beach, FL, Febr. 21, 1999, p.313.
85. Allg. Papier-Rundschau, (14) 340 (1987).
86. W. Druschke, Adhäsion und Tack von Haftklebstoffen, AFERA, 1986,
Edinburgh.
87. A. Johnson and J. Morris, Plast. Film Sheeting, 4(1):50 (1988); in CAS,
Adhesives, 25 (1988), 109:191649v.
88. J.P. Keally and R.E. Zenk (Minnesota Mining and Manufacturing Co.,
St.Paul, MN, USA), Canadian Patent, 1224.678/19.07.1982 (USP. 399350).
89. R. Jordan, Coating, (3) 51 (1986).
90. Adhes. Age, (7) 36 (1986).
91. H.D. Brooks, J.Y. Kelly, P.H. Madison, C.D. Thacher and T.E. Long,
‘‘Synthesis and characterization of Acrylamide Containing Polymers for
Adhesive Applications,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 28,
2001, Williamsburg, VA, p.150.
92. C. Kemmenater and G. Bader, Adhäsion, (11) 487 (1968).
93. H. Liese, Adhäsion, (3) 110 (1966).
94. Coating, (6) 188 (1969).
95. Adhäsion, (3) 73 (1978).
96. A.K. Schultz and N. Kofira, ‘‘The Use of Non-Migrating Surfactants in
Pressure-Sensitive Adhesive Applications,’’ Proc. 24th Annual Meeting of Adh.
Soc., Febr.25, 2001, Williamsburg, VA, p.163.
97. J. Ouyang, S. Jacobson, L. Shen and S. Reedell, ‘‘Characterization of Acrylic
PUR-Based Waterborne PSAs,’’ Proc. 24th Annual Meeting of Adh. Soc.,
Febr.25, 2001, Williamsburg, VA, p.236.
98. C.W. Koch and A.N. Abbott, Rubber Age, (82) 471 (1957).
99. Adhes. Age, (10) 26 (1977).
100. D.J. Yarusso, J. Ma and R.J. Rivard, ‘‘Properties of Polyisoprene Based PSAs
Crosslinked by Electron Beam Radiation,’’ Proc. of the 22nd Annual Meeting
of the Adhesion Society, Panama City Beach, FL, Febr. 21, 1999, p.72.
101. K.L. Ullman and R.P. Sweet, ‘‘Silicone PSAs and Rheological Testing,’’ Proc.
of the 22nd Annual Meeting of the Adhesion Society, Panama City Beach, FL,
February 21–24, 1999, p.410.
102. J. Johnston, ‘‘Alternate Methods for the Basic Physical Testing for PSA Tapes
and Proposals for the Future,’’ Proc. of the 22nd Annual Meeting of the
Adhesion Society, Panama City Beach, FL, Febr. 21, 1999, p.327.
103. K. Fukuzawa and T. Uekita, ‘‘The Mechanism of Peel Adhesion,’’ Proc. of
the 22nd Annual Meeting of the Adhesion Society, Panama City Beach, FL,
Febr. 21, 1999, p.69.
104. K.C. Sehgal, ‘‘Fundamental and Practical Aspects of Adhesive Testing,’’
SME Society of Manufacturing Engineers, Adhesives ’85 Conf. Papers,
Sept.10, 1985, Atlanta, GA, USA.
105. U. Zorll, Adhäsion, (3) 69 (1976).
106. H. Wiest, Adhäsion, (4) 146 (1966).
704 Chapter 10

107. Lihua Li, C. Macosko, G.L. Corba, A. Pocius and M. Tirrell, ‘‘Interfacial
Energy and Adhesion Between Acrylic PSA and Release Coatings,’’ Proc.
24th Annual Meeting of Adh. Soc., Febr.28, 2001, Williamsburg, VA, p.270.
108. J. Johnston, Adhes. Age, (11) 30 (1983).
109. G.R. Hamed and W. Preechatiwong, ‘‘Peel Adhesion of Uncrosslinked
Styrene Butadiene Rubber Bonded to PET,’’ Proc. of the 23rd Annual Meeting
of the Adhesion Society, Myrtle Beach, SC, Febr.20, 2000, p.511.
110. A.J. Kinloch, B.R.K. Black, H. Hadavinia, M. Paraschi and J.G. Williams,
Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001, Williamsburg, VA, p. 44.
111. D.R. Moore and J.G. Williams, in Fracture Mechanics Testing Methods
Polymers, Adhesives and Composites, Eds. D.R. Moore, A. Pavan and J.G.
Williams, Elsevier Science Publishers, Amsterdam, 2000.
112. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p.285.
113. J.M. Piau, G. Ravilly and C. Verdier, ‘‘Experimental and Theoretical
Investigations of Model Adhesives Peeling,’’ Proc. 24th Annual Meeting of
Adh. Soc., Febr.28, 2001, Williamsburg, VA, p.287.
114. G.R. Hamed and C.H. Hsieh, Rubber Chem. Technol., 59, 883 (1986).
115. E.P. Chang, I.F. Wang, and M. Ziemelis, ‘‘Excimer Fluorescence Method for
Determining Cure of Release Coating,’’ Proc. of the 22nd Annual Meeting of
the Adhesion Society, Panama City Beach, FL, Febr. 21, 1999, p.421.
116. Coating, (1) 8 (1985).
117. R. Muny, Labels and Labelling, (1) 36 (1999).
118. V. Gordon, T.M. Leay, M.J. Owen, M.S. Owen, S.V. Perz, J.L. Stasser, J.S.
Tonge, M.K. Chaudhury and K.A. Vorvolakos, ‘‘Resin–Polymer Interaction
in Silicone Release Coating,’’ Proc. of the 23rd Annual Meeting of the Adhesion
Society, Myrtle Beach, SC, Febr.20, 2000, p.39.
119. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
‘‘Rheology in the Release of Silicone Coatings,’’ in Book of Abstracts, 215th
ACS National Meeting, Dallas, March, 29 (1998).
120. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
‘‘Silicone Release Coatings: An Examination of the Release Mechanism,’’
Adhes. Age, (11) 35 (1998).
121. G.V. Gordon, R.L. Tabler, S.V. Perz, J.L. Stasser, M.J. Owen and J.S. Tonge,
‘‘Release Force Control: The Bottom Line Õ . Synergism Between Adhesive
and Liner,’’ Proc. of the Pressure-Sensitive Tape Council, Technical Seminar,
Washington DC, May 5, 1999.
122. M. Gerace, Adhes. Age, (8) 85 (1983).
123. I. Benedek, ‘‘Bond Failure in Pressure-Sensitive Removable Thin Plastic Film
Laminates,’’ Proc. of the 22nd Annual Meeting of the Adhesion Society,
Panama City Beach, FL, Febr. 21, 1999, p.418.
124. A. Aymonier, E. Papon, J.J. Villenave and P. Tordjeman, ‘‘Direct Relation
Between Copolymerization Process and Tack Properties of Model PSA’s,’’
Proc. 24th Annual Meeting of Adh. Soc., Febr.28, 2001, Williamsburg,
VA, p.280.
Test Methods 705

125. A.A. Chalykh, A.E. Chalykh, V. Yu Stepanenko and M.M. Feldstein,


‘‘Viscoelastic Deformations and the Strength of PSA Joints Under Peeling,’’
Proc. of the 23rd Annual Meeting of the Adhesion Society, Myrtle Beach, SC,
Febr.20, 2000, p.252.
126. R. Sweet and K. Ulman, CRS Symposium, 1984; in ‘‘Silicone PSAs and
Rheological Testing,’’ Proc. of the 22nd Annual Meeting of the Adhesion
Society, Panama City Beach, FL, Febr. 21, 1999, p.410.
127. Coating, (11) 316 (1984).
128. D.J. St.Clair and J.T. Harlan, Adhes. Age, (12) 18 (1975).
129. B. Zhao, E. Miasek and R. Pelto, ‘‘Peeling Tapes From Paper,’’ Proc .24th
Annual Meeting of Adh. Soc., Febr.25, 2001, Williamsburg, VA, p.364.
130. J.J. Bikermann and W. Whitney, Tappi, 46 (7) 420 (1963).
131. T. Yamauchi, T. Cho, R. Imamura and K. Murakami, ‘‘Peeling Behavior of
Dehesive Tape From Paper,’’ Nordic Pulp Paper J., (4), 128 (1988).
132. R. Pelton, W.Chen, H. Li and M. Engel, ‘‘The Link Between Paper Properties
and PSA Adhesion,’’ Proc. of the 23rd Annual Meeting of the Adhesion
Society, Myrtle Beach, SC, Febr.20, 2000, p.120.
133. Adhäsion, (10) 399 (1965).
134. P. Dunckley, Adhäsion, (11) 19 (1989).
135. S. Trenor, A. Suggs and B. Love, ‘‘An Examination of How Skin Surfactants
Influence a Model PIB PSA for Transdermal Drug Delivery,’’ Proc. 24th
Annual Meeting of Adh. Soc., Feb.25, 2001, Williamsburg, VA, p.144.
136. J.C. Conti, E.R. Strope, R.D. Gregory and P.A. Mills, ‘‘Cyclic Peel
Evaluation of Sterilized Medical Packaging,’’ Proc. of the 22nd Annual
Meeting of the Adhesion Society, Panama City Beach, FL, Febr. 21,
1999, p.116.
137. J.C. Conti, E.R. Strope and E. Jones, Proc. of the 20th Annual Meeting of the
Adhesion Society, 1997, p. 425.
138. J.C. Conti, E.R. Strope, E. Jones and D. Rohde, Proc. of the 21st Annual
Meeting of the Adhesion Society, 1998, p. 418.
139. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p.281.
140. Y. Lai and D.A. Dillard, Internat. J. Solids Structures, 34, 509, 1997.
141. E.O. Brien, K. Doyle, T.C. Ward, S. Gio and D. Dillard, ‘‘Adhesion of Model
Epoxy Bonded to Glass Subjected to Chemical Stress Measured by the Shaft
Loaded Blister Test,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr.25, 2001,
Williamsburg, VA, p.475.
142. E.P.O‘ Brien and T.C. Ward, ‘‘Characterization of Thin Films :Shaft Loaded
Blister Test,’’ Proc. of the 23rd Annual Meeting of the Adhesion Society, Myrtle
Beach, SC, Febr.20, 2000, p.202.
143. E.O. Brien, T.C. Ward, S. Gio and D. Dillard, ‘‘Characterizing the Adhesion
of Pressure-Sensitive Tapes Using the Shaft Loaded Blister Test’’, Proc. 24th
Annual Meeting of Adh. Soc., Feb.25–28, 2001, Williamsburg, VA, p.113.
144. K.T. Wan and Y.W. Mai, Internat. J. of Fracture, 74, 181 (1995).
145. Gent A.N., J. Adhes., 23, 115 (1987).
706 Chapter 10

146. A. Paiva, M.D. Foster, A.J. Crosby, and K. Shull, ‘‘Studying Changes in
Surface Adhesion of PSAs with AFM and Spherical Indenter Test,’’ Proc. of
the 23rd Annual Meeting of the Adhesion Society, Myrtle Beach, SC, Febr.20,
2000, p.43.
147. Adhes. Age, (11) 28 (1983).
148. Glossary of Terms used in Pressure Sensitive Tapes Industry, PSTC, Glenview,
IL, USA, 1974.
149. O. Hahn, M. Schlimmer and D. Ruttert, Adhäsion, (12) 9 (1988).
150. D.J. James and H.C. Holyoke, Adhes. Age, (4) 23 (1984).
151. Adhäsion, (1/2) 27 (1987).
152. L.J. Smith, Adhes. Age, (4) 28 (1987).
153. C.M. Chum, M.C. Ling and R.R. Vargas (Avery Int.Co., USA), EPA
1225792/18.08.1987.
154. R. Lombardi, Paper, Film and Foil Conv., (3) 74 (1988).
155. A. Sustic and B. Fellow, Adhes. Age, (11) 17 (1991).
156. Adhes. Age, (12) 25 (1977).
157. Wingtack, Technical Bulletin, Goodrich, Akron.
158. R. Hinterwaldner, Coating, (3) 73 (1985).
159. J.A. Schlademan, ‘‘The Role of Tackifier Compatibility and Molecular
Weight on Bulk Properties of PSAs,’’ Proc. of the 22nd Annual Meeting of the
Adhesion Society, Panama City Beach, FL, Febr. 21, 1999, p.75.
160. A. Moser and J.G. Dillard, Private Comm.; in Ref. 156.
161. L.A. Sobieski and T.J. Tangney, Adhes. Age, (12) 23 (1988).
162. Adhes. Age, (5) 24 (1987).
163. R.D. Adams, R. Thomas, F.J. Guild and L.F. Vaughan, ‘‘Thick Adherend
Shear Tests,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001,
Williamsburg, VA, p.59.
164. D.J. Yarusso, J. Ma and R.J. Rivard, ‘‘Properties of Polyisoprene-Based
PSAs Crosslinked by Electron Beam Radiation,’’ Proc. of the 22nd
Annual Meeting of the Adhesion Society, Panama City Beach, FL, Febr. 21,
1999, p. 72.
165. K.E. van Holde and J.W. Williams, J. Polymer Sci., 11, 243 (1953).
166. Tackifier Resins Data Sheets, Akzo-Eisele & Hoffmann, Mannheim,
Germany, 1987.
167. Water Based PSA, Data Sheets, Rhone-Poulenc, 1990.
168. Die Herstellung von Haftklebstoffen, Tl. 2.2; 15d, Nov. 1979, BASF,
Ludwigshafen, Germany.
169. Coating, (7) 20 (1984).
170. E.G. Ewing and J.C. Erickson, Tappi J., (6) 158 (1988).
171. Adhes. Age, (10) 24 (1977).
172. Coating, (7) 186 (1984).
173. Technical Information, LAT 037/Aug 82; LAT 002/Aug 87; LAT 051/May 87;
LAT 040/Aug. 82, Doverstrand Ltd., Harrow (UK).
174. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, Chapter 5.
Test Methods 707

175. Z. Czech, Adhäsion, (11) 26 (1994).


176. Allg. Papier-Rundschau, (18) 572 (1987).
177. N. Willenbacher, ‘‘High Resolution Dynamic Mechanical Analysis (DMA)
on Thin Films,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001,
Williamsburg, p.263.
178. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 4.1.
179. M. Ziemelis and E.P. Chang, ‘‘Curable Fluorescent Oxygeno-Polysiloxane
Compositions,’’ U.S.Pat., 5,545,830,13.08.1986; in 115.
180. S. Price and J.B. Nathan Jr., Adhes. Age, (9) 37 (1974).
181. S. Milton and Ch. Max, Adhes. Age, (1) 12 (1983).
182. D. Becker and J. Braun, Kunststoff Handbuch, Bd.1, K. Hanser, München-
Wien, 1990, p. 931.
183. D.J.P. Harrison, J.F. Johnson and J.F. Yates, Polymer Eng. Sci., (14) 865
(1982).
184. R. Hinterwaldner, Adhäsion, (3) 14 (1985).
185. R.F. Grossmann, Adhes. Age, (12) 41 (1969).
186. Adhes. Age, (3) 36 (1986).
187. Technical Service Report, 6110, Firestone, August, 1986.
188. J.F. Kauffmann, Adhäsion, (10) 163 (1981).
189. S. Mitton and C. Mak, Adhes. Age, (1) 42 (1983).
190. Resins, Narez Technical Bulletin (Spain), 1990.
191. Tappi J., 67 (9) 104 (1983).
192. R. Mudge, ‘‘Ethylene-Vinylacetate Based, Waterbased PSA,’’ in TECH 12,
Advances in Pressure Sensitive Tape Technology, Technical Seminar Proc.,
Itasca, IL, USA, May 1989.
193. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, Chapter 4.1.2.
194. P.H. Gamlen and R.M. Lang (ICI), Chemicals and Polymers, Research and
Technology, p. 8.
195. Adhäsion, (1) 11 (1974).
196. Mobil Plastics Technical Bulletin, MA 657/06/90.
197. C.W. Drechsler, Coating, (1)10 (1985).
198. I. Benedek, Development and Manufacture of Pressure-Sensitive Products,
Marcel Dekker, New York, 1999, p.544.
199. I. Benedek, Pressure-Sensitive Formulation, VSP, Utrecht, 2000, p.81.
200. K. Goller, Adhäsion, (4) 101 (1974).
201. K. Goller, Adhäsion, (4) 126 (1973).
202. K. Goller, Adhäsion, (7) 266 (1973).
203. F. Altenfeld and D. Breker, Semi Structural Bonding with High Performance
Pressure Sensitive Tapes, 3M Deutschland, Neuss, Germany, 2nd ed.,
Febr.1993, p.278.
204. A. Dobman and J. Planje, Papier u. Kunststofjverarb., (1) 37 (1986).
205. A.L. Bull, Thermoplastic Rubbers, Technical Manual, Shell Elastomers,
TR 8.12, p. 13.
206. I.J. Davis (National Starch, Chem. Co., Bridgewater, USA), US Pat.,
4.728.572/ 01.03.1988.
708 Chapter 10

207. R. Schuman and B. Josephs (Dennison Manuf. Co., USA), PCT/US86/02304/


25.08.1986.
208. G.W. Horst Lehmann, and H.A. Julius Curts (Beiersdorf A.G., Hamburg,
Germany), US Pat., 4038454/26.07.1977.
209. C.M. Miller and H.W. Barnes, ‘‘Factors Affecting Water Resistance of Latex-
Based PSAs,’’ Proc. 24th Annual Meeting of Adh. Soc., Febr. 25, 2001,
Williamsburg, VA, p.153.
210. A. Kenneth, J.R. Stockwell and J. Walker, (Allied Colloids Ltd., Low Moor,
UK), EP 0147 067/28.11.1983.
211. US Pat., No., 3691140/12.09.1972; in Hiroyashu Miyasaka, Yasuaki Kitazaki,
Tetsuaki Matsuda and Junichi Kobayashi, Nichiban Co., Ltd. Tokio, Japan,
Offenlegungsschrift, DE 3544868 A1/18.12.1985.
212. Coating, (3) 65 (1974).
213. Adhes. Age, (11) 40 (1988).
Abbreviations and Acronyms

1 COMPOUNDS

AA acrylic acid
A-CPE amorphous copolyethylene
AC acrylic
AN acrylnitrile
APAO amorphous polyalphaolefin
APO amorphous polyolefin
APP atactic polypropylene
B butadiene
BOPP biaxially oriented polypropylene
BPO benzoyl peroxide
BuAc butyl acrylate
CMC carboxymethyl cellulose
CPP cast polypropylene
CRA control release agent
CSBR carboxylated styrene-butadiene rubber
DEA diethanol amine
DOP dioctyl phthalate
EA ethyl acrylate
EB ethylene butylene
EPC epichlorohydrine
EHA ethyl hexyl acrylate
EMA ethyl methacrylate
EPDM ethylene propylene diene monomer

709
710 Abbreviations and Acronyms

EPR ethylene propylene rubber


EPVC emulsion PVC
EVAc ethylene vinyl acetate
GR-S a special grade of rubber
HDPE high density polyethylene
H-PVC hard PVC
LDPE low density polyethylene
M maleinate
MAA methacrylic acid
MEK methyl ethyl ketone
MIBK methyl isobutyl ketone
MMA methyl methacrylate
NR natural rubber
OPP oriented polypropylene
PA polyamide
PB polybutylene
PBuAc polybutyl acrylate
PC polycarbonate
PE polyethylene
PEB poly(ethylene butylene)
PEHA polyethyl hexyl acrylate
PET polyethylene terephthalate
PIB polyisobutylene
PP polypropylene
PS polystyrene
PUR polyurethane
PVA polyvinyl alcohol
PVAc polyvinyl acetate
PVC polyvinyl chloride
PVE polyvinyl ether
PVP polyvinyl pyrrolidone
RR rubber/resin
S styrene
SBR styrene-butadiene rubber
SBS styrene-butadiene-styrene
SEBS styrene-(ethylene/butene)-styrene
SI styrene-isoprene
SIS styrene-isoprene-styrene
STS stainless steel
VAc vinyl acetate
VAc/E vinyl acetate/ethylene
VE vinyl ether
Abbreviations and Acronyms 711

2 TERMS

A area or a constant
aT temperature dependent shift factor
Ao reference area
AFERA Association des Fabricants Europeens de Rubans
Auto-Adhesifs
ASTM American Society for the Testing of Materials
B constant
BGA Bundesgesundheitsamt
BWA bonding wet adhesion
BWB German military procurement office
BWB-TL German military norm
C constant or cuttability or cohesion
Cw coating weight
CC carbon-carbon
CD cross direction
CH carbon-hydrogen
CI cuttability index
CMD cross-machine direction
DIN German standard
DMA dynamic mechanical analysis
DP degree of polymerization
DSC differential scanning calorimetry
DWA debonding wet adhesion
E modulus
EB electron beam
EDP electronic data processing
EN European norms
EPSMA European Pressure Sensitive Manufacturers Association
eV electron volt
EV viscosity test method
FDA Food and Drug Administration
FINAT Federation Internationale des Fabricants Transformateurs
d’Adhesifs et Thermocallants sur Papiers et Autres Supports
FIPAGO Federation Internationale des Fabricants de Papiers Gommes
FTM FINAT Test Method
G0 storage modulus
G00 loss modulus
GID gear-in-die
HC hydrocarbon
HD high density
712 Abbreviations and Acronyms

HLB hydrophilic-lipophilic balance


HM hot-melt
HS hot shear
HMA hot-melt adhesive
HMPSA hot-melt PSA
HSG hot shear gradient
IR infrared
ISO International Standards Organization
kp kilopond
LD low density
LEFM linear elastic fracture mechanics
LVT known Brookfield viscosity method
MD machine direction
MFI melt flow index
MFR melt flow rate
MFT minimum film-forming temperature
MW molecular weight
MWD molecular weight distribution
MP melting point
mPa.s milliPascal second
MV viscosity test conditions
NC viscosity test conditions
PN Williams plasticity number
Ppt parts per thousand
PS pressure-sensitive or polystyrene
PSA pressure-sensitive adhesive
PSP pressure-sensitive product
PSTC Pressure-Sensitive Tape Council
Pts parts
PTS Papiertechnische Stiftung
R universal gas constant
RB rolling ball
RBT rolling ball tack
RCT rolling cylinder tack
RF radiofrequency
RH relative humidity
RT room temperature
RVT known Brookfield viscosity method
SAA surface active agent
SAFT shear adhesion failure temperature
SB solvent-based
SF solvent-free
Abbreviations and Acronyms 713

SH shear
SL solventless
SLBT shaft loaded blister test
SP softening point
SPS slight panel staining
TAST thick adhesive shear test
TD transverse direction
Tg glass transition temperature
TMA thermo-mechanical analysis
TLMI Tag and Label Manufacturers Institute
TPE thermoplastic elastomer
UV ultraviolet
WA work of adhesion
WB water-based
WBA wet bonding adhesion
Index

[Acrylics]
Acrylics pure, 177
adhesion/cohesion balance of, radiation cured,
181 (see also Adhesion/ crosslinking of, 195 (see also
cohesion balance) Crosslinking)
dependence of, on polymeriza- tackified (see Tackification)
tion technology, 180 tackifying ability of, 38
adhesion towards polar surfaces water-based, 214 (see also Water-
of, 180, 266 based dispersions)
advantages of, 177 Additive, 447
cohesion of, 330 choice of, dependence on coating
comparison of (see also machine, 482, 538
Comparison of PSA) Adhesion
vs. CSBR, 177 build up, 266, 289
vs. EVAc, 178 -cohesion balance, 12, 26, 235,
compatibility of, 11, 434 320, 330
converting costs of, 341 (see also failure energy of, 37 (see
Converting properties) Separation energy)
copolymers of, 178 dependence of, on debonding, 269
vinyl based, 181, 182 failure mode, 269
vinyl acetate, 182 interfacial, 61, 76, 79, 270, 283,
crosslinking of, 192, 194, 314, 332 289
(see also Crosslinking) on polyethylene, 167 (see also
die-cuttability of, 177 (see also Peel)
Cuttability) test methods of, special, 666, 667
electron-beam cured, 196, 214 wet, 79, 96, 492, 497, 500, 790
(see also Electron-beam work of, 132, 133
curing) Adhesive, 1
peel of (see Peel) anchorage of, 27, 29, 50, 57, 70,
plasticizer resistance of, 177 72, 340
(see also Plasticizer dependence on adhesion/
resistance) cohesion, 291

715
716 Index

[Adhesive] [Adhesive]
dependence on coating influence of on the
technology, 57 formulation, 211
improvement of with plasticity of, 28, 30, 38, 63
solvents, 533 removable, 8, 15, 18, 19, 27, 30, 78
parameters of, 409 (see also Removable
for removable adhesives, 291 adhesive)
break, test of (see Test methods)
dependence of, Agent,
on peeling rate, 18 antimicrobial, 429, 478
on temperature, 18 of controlled release, 201, 202
build up of, on the machine, 45, 50 (see also Release
characteristics of, 235 coating)
balance of, 430 of crosslinking, 192, 519 (see
dependence on coating tech- also Crosslinking)
nology, 472 (see also in water-based PSA, 519 (see
Coating technology) also Water-based
dependence on coating weight, dispersions)
253 (see also Coating of detackifying, 501, 509, 512 (see
weight) also Peel, Cuttability)
global evaluation of, 677, 680 of formulating, 104 (see also
rheological evaluation of, 681 Formulating)
coating of, discontinuous, 305 of hygroscopicity, 539
(see also Peel) of lowering the viscosity, 533
coating machine of, 560 (see also protective, 520
Coating machine) solubilizing, 536 (see also
converting properties of, 25 Water resistance)
degradation of (see Aging) of stabilizing in water-based
failure of (see Mode of failure) dispersions, 520 (see
flow out, 400 (see also Cuttability) also Water-based
formulating of, 426 (see also dispersions)
Formulating) of wetting (see Wetting agents)
hold out of (see Coating weight) Aging
influence on converting resistance, 334, 339
properties, 320 (see also test of, conditions of, 470, 499
Converting properties) Air
influence on end use properties, brush, 568, 585, 587, 591
320 knife, 564, 588, 565 (see also
modification of, 190 (see also Coating device)
Formulation) Amorphous polyalphaolefins
on PVC label, 490 (see also for migration resistance, 508
Label) (see also Migration)
permanent, 11, 15, 18, 27, 38, Antioxidants, 166, 192, 207, 312, 466,
57, 65, 78 502 (see also Aging)
physical state of, for aqueous dispersions, 502, 518
Index 717

[Antioxidants] [Carboxylated styrene-butadiene]


for electron beam cured Tg values for, 96 (see also
elastomers, 518 Glass transition
for hot-melts, 502, 510 temperature)
level of, 503 Carrier
for natural rubber, 504 bulk characteristics of,
for SBR latex, 166 influence on adhesive
for thermoplastic elastomers, 510 properties, 58
Approval, 640 surface properties of, 71
BGA, 536, 537, 640, 643 influence on,
FDA, 536, 537, 640, 643 anchorage of PSA, 89
Arrhenius’s low, 34 peel, 281
removability, 78
Ball tack test (see Rolling ball) rheology, 70
BGA compliance (see Approval) thickness of, critical, 279
Biodegradable polymers, 610 Cavitation, 131, 133 (see also
(see also Recycling) Debonding)
Blade coaters (see Coaters) Gent’s criterion, 134
Bleeding, 71, 148, 149, 151, 155, 255 Cell depth, 258 (see also
(see also Migration) Gravure cylinder)
dependence of, Cellulose,
on coating weight, 148 (see derivatives of, as face stock
also Coating weight) material, 365, 385
on application viscosity, 150 Chemical composition,
Block copolymers (see also of formulated adhesive (see
Segregated elastomers) also Formulating)
radial, 171 influence on peel, 265 (see
star shaped, 171 also Peel)
Bonding wet adhesion, 375 (see also of PSA, 161
Adhesion) dependence on coating weight
Break down edge (see Stationary roll) (see Coating weight)
Butylene rubber, 165 dependence on end use, 216
(see also Polybutene) (see also End use
properties)
Carboxylated styrene-butadiene dependence on formulation,
latex, 13, 22, 167, 246 216 (see also
advantages of, 167 Formulation)
compatibility of, 434 dependence on physical status,
crosslinking of, 195 (see also 161, 211
Crosslinking) dependence on solid state
molecular weight of, 13, 22, components, 221
81, 91 dependence on synthesis, 205
shear values of, 322 (see also factors of, 205
Shear strength) choice of additives, 209
styrene content of, 96 choice of monomers, 206
718 Index

[Chemical composition] [Coating]


choice of polymerization device of, 565
procedure, 207 choice of, 565
polymerization, 205 of laboratory samples, 645, 693
synthesis, 205 Meyer bar (see Meyer bar)
for direct coated PSAs, 221 parameters influencing the
for transfer coated PSAs, 221 usability of, 567
of release coatings, 199 (see requirements on, 585
also Release coating) with stationary roll (see
Chloroprene latex (see also Stationary roll)
Neoprene) as shear for water-based PSA (see
modifier, 434 also Water-based
Coagulum dispersions)
build up (see also Adhesive direct, 50, 73, 74, 221, 566
build up) of film (see also Film coating)
on machine, 357, 456 metering device for,
level of, 357, 636 50, 74
test of, 357 (see also Test method) geometry of, choice of, 585
Coalescence, 109, 356 in line, 560, 561 (see also
dependence of Radiation curing,
on glass transition temperature, Siliconizing)
110 machine of, 561, 562
on viscosity, 110 for labelstock, 49
Coatability, 25, 40, 45, 48, 70, 80 for silicones, 561 (see
dependence of also Siliconizing)
on foaming, 535 (see also Foam) for solvent-based PSA, 43 (see
on tackification, 455 (see also also Solvent-based
Tackification) PSA)
of diluted systems, 354 for water-based PSA (see
of HMPSA, 359 Water-based
of radiation-cured systems dispersions)
of solvent-based PSA, 354 on reverse roll, 50 (see also
of water-based PSA, 353 Reverse roll)
Coater rheology of, 40, 41, 44
of cartridge style, 566, 585 (see of HMPSA, 374
also Coating machine) of solvent-based PSA, 42
choice of, 585 speed of, 45, 562 (see also
die, 585 (see also Slot die) Running speed)
Duplex, 591 (see also Gear in die) dependence on tackification,
formulation, 514 455
kiss, 565, 578 of HMPSA, 359
knife, 567 (see also Knife) of solvent-based PSA, 354
Park, 591 (see also Slot die) systems, 565 (see also Coating
Coating, 560 device)
defects of, 564, 593 technology of, 560
Index 719

[Coating] [Coating]
influence of, on the influence on shear, 253, 315
formulation, 564 (see also Shear)
influence of, on the adhesive parameters of, 48, 51, 252
properties, 564 of primer, 300
transfer, 44, 46, 71, 73, 221 tolerances of, 47, 258, 578, 590
versatility of, 354, 355, 357 values of, 58, 258, 260, 279, 306
of HMPSA, 359 Cohesion, 156 (see also
weight of, 252, 681 Shear strength)
adjustment of, 306 dependence of
critical, 55 on additives, 315
dependence on adhesive on crosslinking, 193 (see
quality, 270 also Crosslinking)
dependence on chemical on molecular weight, 312
composition, 252 on temperature, 402
dependence on coating regulation of, 480
conditions, 254 Cohesive strength, 21 (see also
dependence on end use, 258 Cohesion; Shear)
dependence on face material, Cold flow, 14, 26, 28, 40, 55, 57, 66, 148
258 dependence of
on face stock porosity, 70 on chemical affinity of face
on gravure cylinder, 258 (see also stock/liner, 407, 409
Gravure cylinder) on coating technology, 57
of hot-melts, 594 (see also on coating weight, 57, 148, 408
Hot-melts on creep modulus, 127, 270
dependence on substrate, 12, on laminate, 404
46, 55, 65, 72 on shear rate, 150
dependence on surfactants, on smoothness of face stock/
252, 255 liner, 405
dependence on viscosity, 262 on viscosity, 150
influence on adhesive Colloids, 159 (see also Protective
properties, 57 colloid)
influence on cold flow, 57 Comonomer, in carboxylated
(see also Cuttability) latex, 166 (see also
influence on converting Monomers)
properties, 57 Comparison of PSA, 147
influence on cuttability, 57 of acrylics and other synthetic
influence on drying, 252, polymer-based
289, 315 elastomers, 321
influence of, on the additive adhesion on polar/nonpolar
choice, 470 surfaces, 326
influence on peel, 253 on different chemical basis, 321
(see also Peel) in film application, 325
influence on tack, 253 flow of, vs. plastics, 151
(see also Tack) formulation of, 148, 210
720 Index

[Comparison of PSA] [Contact]


mechanical resistance of, vs. physics, 129
plastics, 156 surface of, reduction of, 305
in permanent/removable paper Convertability, 26, 80 (see also
label application, 325 Converting properties)
relaxation of, vs. plastics, 155 dependence of
rheology of, on adhesive properties, 354, 359
vs. plastics, 151 on adhesive physical state, 354
vs. other adhesives, 159 on coating technology, 359
on rubber-based vs. on coating weight, 56
acrylic-based, 321 on end use properties, 360
on solvent-based acrylics vs. on laminate, 26, 80
solvent-based rubber on solid state components of
PSAs, 244 the laminate, 361
on solvent-based acrylics vs. Converting properties, 25, 320,
water-based acrylics, 546 353, 484
on solvent-based/water-based/ dependence of
hot-melt, 543, 546 on coating weight (see
synthesis of, 254 Coating weight)
on water-based acrylics vs. other on glass transition
water-based PSAs, 244 temperature, 106
with thermoplastics and rubber, on laminate, 80
148 on tackifier (see Tackification)
Compatibility, 106 on viscoelastic properties, 25
of the resin, 434 (see also Resin) Corrugated board, shear test on, 308
dependence on melting point, Cratering, 74, 531 (see also Test
436, 307 methods)
dependence on molecular Creep, 46, 28, 305 (see also
weight, 116, 462 Cold flow)
with PSA dispersions, 431 compliance of, 5, 15, 110, 132,
Composite structure, 363 274, 307
of PSA layer, 41, 166, 315, 454, 481 dependence of
influence on adhesive proper- on anchorage, 401 (see also
ties, 108 (see also Anchorage)
Adhesive properties) on adhesive nature, 400, 416
Contact test of, 677
angle of, 47, 48, 74, 363, 521 values of, 127
(see also Wetting out) Crosslinking
measurement of (see Test chemical, of CSBR, 117, 192,
methods) influence of
cement of (see Primer) on the peel (see Peel)
hindrance of, 489 on the shear (see also Shear)
mechanics, of natural rubber, 117
for elastics materials, 129 (see also Natural rubber)
for viscoelastic materials, 130 physical, 117, 172, 314
Index 721

[Crosslinking] [Cuttability]
of rubber derivatives, 192 on face stock smoothness, 80
of SIS, 96 on formulation of tackifier
Curing dispersion, 458
beam, 190, 193, 195 (see on fillers, 315
also Radiation curing; on laminate stiffness, 69, 81, 408
Silicone) on laminate thickness, 431
generator of, 616 on peel, 405, 406, 509, 488
of rubber-based PSA, 190, 193, 615 on shear resistance, 317
dual, 615 on smearing, 66, 307 (see also
electron beam, 40, 100, 103, 195, Smearing)
314, 615 on stress distribution, 68,
advantages/disadvantages of, 69, 70
615 on tack, 406
curing time of, 615 (see on time/temperature dependent
also Siliconizing) rheology of PSA, 405
influence on face stock, 615 dynamic, 400
of silicones, 614 (see also evaluation of, 396, 397, 398
Release coating; of film paper/film laminate, 397
Siliconizing) improvement of, dependence on
post, 201 (see also Silicones) shear, 403, 404
radiation, 193, 194 (see also Radia- parameters of, 70, 400, 396
tion curing; Silicone) static, 396
of silicone, 195 (see also Silicone) of sheet materials, 397 (see also
thermal, 200 (see also Silicones) Sheet label)
ultraviolet, 40, 195, 214, 306 Cutting
(see also Silicones) angle of, 396 (see also Cuttability)
Curl, 56, 392 (see also Lay flat) on flying knife, 27
Cuttability, 58, 395 on guillotine, 27, 400
dependence of process of, 66
on adhesion/cohesion, 401 steps of, 66, 67, 400
on adhesive anchorage, 57, 80, types of, 396
74, 407 Cylinder depth, 576 (see also
on adhesive properties, 403 Gravure depth)
on bulk material, 66
on coating weight, 57, 408 Dahlgrens system, 588 (see also
on cold flow, 57, 407, 302 Coating machine)
on composite structure, 70, Dahlquist’s criterion, 13 (see also
408, 410 compliance)
on creep, 407 Debonding wet adhesion, 375 (see
on cutting process conditions, also Water resistance)
407 Deep freeze
on environmental humidity, adhesives
70, 410, 411 migration of, 14, 108, 291
on face stock modulus, 70 test of, 504
722 Index

[Deep freeze] [Dryer]


labels, 110 ultraviolet, 601
Defoamers, 314, 335, 558 (see of water-based PSA, 373, 374
also Foam) Drying, 252, 358, 359
Detachment of coating, 598
angle, 64 tunnel of, 47, 109, 599
energy, 54, 84 (see also Peel) conditions of, 599, 645
Diblock/triblock ratio, dependence of
influence of, on face stock, 482
on modulus, 16, 236 on carrier, 598
on converting properties, 16 energy requirements of, 600
Die coater, influence of, on infrared, 600, 601
coating quality, 459 influence of, on shrinkage, 373
(see also Gear in die; methods of, 600
Slot die) radiofrequency, 602
Die cuttability (see also Cuttability) dependence on coating weight,
of acrylics, 336 252
of films, 375 dependence on solids content, 482
dependence of dependence on tackification
on hygroscopicity, 472 (see Tackification)
on shear, 307 of solvent-based PSA, 44, 563
on tackification, 473 temperature of, 360, 374, 376, 385
Diluent, reactive, 195, 196 on polypropylene, 376
(see also Radiation on soft PVC, 74, 109, 372
curing) of water-based PSA, 122, 252, 563
Diluting, 357, 530, 533, 534 (see also Water-based
agents (see Formulation) dispersions)
response (see Water-based speed of, 252
dispersions) DSC, 10 (see also Glass transition
Dimensional stability, 377 temperature)
(see also Shrinkage) Dwell time, 16, 31, 55 (see also
of films, 377 Adhesive properties;
parameters of, 384, 388, 392 Test methods)
Direct gravure, 593, 603 values of, 235, 650, 659, 663,
advantages/disadvantages of, 666, 670
454
DMA, 10, 11, 67, 112, 431, Edge
440, 680 lifting of, 366,
Doctor blade, 566, 567 flow (see Oozing)
oscillating, 574 Elastic modulus (see Modulus)
reverse angle, 582 Elastomers, 163, 180 (see also
two chamber system, 585 Raw materials)
Dryer acrylic, 163, 169
electron beam, 601 ethylene-vinyl acetate, 163
flying, 601 natural, 164
Index 723

[Elastomers] [Ethylene vinylacetate copolymers]


synthetic, 165, 166 solids content of, 195
hydrocarbon-based, 166 water-borne, 183 (see also
manufature of, 314 Water-based
thermoplastic (see Thermoplastic dispersions)
elastomers) Eyrings model, 35
uncrosslinked, 125
Emulsifiers, 520 (see also Surfactants) Face stock, 5, 58, 365 (see
migration of, 335 (see also also Laminate)
Migration) choice of, parameters of, 222,
Emulsion properties 365, 367
improvement of, 520 (see cost effectiveness of, comparison
also Formulation) of, 375
End use properties, 489 critical deformability of, 281
dependence of influence of on peel, 281
on glass transition of film, 365, 369, 370
temperature, 106 advantages of, 369
modulus, 151 comparison of vs. paper, 386
on tackifier (see Tackification) modulus of, 386
on viscoelastic properties, profil tolerances of, 374
26, 148 damage of, through radiation,
influence on coating weight 215
(see Coating weight) flexibility of
Energy of separation, 38 (see influence on peel, 61, 66, 276
also Work of adhesion) (see also Peel)
Environmental considerations, 602 influence on rheology, 5, 61
Ethylene copolymers, 162, 182 influence on tack, 241 (see
sequence distribution of, 103 also Tack)
Ethylene maleinate copolymers, geometry of, 80
183, 332 (see also influence on cuttability, 68
Water-based dispersions) influence on peel, 61
cohesion of, 209 influence on pressure-sensitive
Ethylene propylene diene properties, 80
copolymers, 102 (see influence of
also Thermoplastic on adhesive choice, 70
elastomers) on adhesive properties, 60
Ethylene vinylacetate copolymers, on anchorage, 70
182, 183 (see also on coating technology, 71, 73
Water-based on converting properties, 58,
dispersions) 65, 70
compatibility of, 434 on cuttability, 65
glass transition temperature of, on migration, 71
184 on peel, 60
in raw materials, for hot melts, on removability, 61, 78, 80
184 (see also Hot melts) on shear, 66
724 Index

[Face stock] Flagging, 320


on tack, 61 factors of, 375
in laboratory tests, 645 test of (see Test methods)
of low surface energy, 392 Flexibilizers, 304 (see Removability)
modulus of, influence on Flow, test of, in plastics vs. PSA, 151
cuttability, 69 Fluorosilicones, 172
of paper, 381, 386 Foam, 46, 51, 358 (see also
(see also Paper) Water-based dispersions)
plasticity of, influence on dependence of
peel, 62, 276 on coating device, 535
(see also Peel) on running speed, 358
stiffness of, 58 surface tension, 535 (see also
comparison of, 375 Surface tension)
influence on peel, 61, 190 on tackification, 636 (see also
strengthening of, 305 Tackification)
surface of resistance, 493
influence on adhesive testing of (see Test method)
properties, 70 Formulation, 125
influence on anchorage, 80 age of, influence on adhesive
influence on converting properties, 501
properties, 80 components of, other 198
influence on wetting out, ease and costs, of acrylics, 240
73 (see also of adhesive properties, 430
Wetting out) on aging, 499 (see also Aging)
thickness of, 80 for coating properties, 211, 473
influence on peel, 81, 281 for converting properties, 484
FDA compliance (see Approval) dependence of
Fibrillation, 131, 134, 643 on adhesive technology, 505
(see also Contact on coating technology, 482
Physics) on face stock, 482
as engineering phenomenon, 158 for direct coating, 220, 221
Filler, 315, 430, 480 for cuttability, 486 (see also
in peel reduction, 305, 480 Cuttability)
reactive, 514 of deep freeze PSA, 505 (see
influence on shear, 33 also Deep freeze label)
Film for end use properties, 216, 488
coating of, metering device in, 73 for environment, 214, 223
face stock, 369 (see also Face for reel/sheet, 216
stock) permanent, 490
plastic, 78 (see also Face stock) removable, 490
Filter, for water-based PSA, 588, 637 of hot melts, 483, 515, 549
(see also Mechanical (see also Hot melts)
stability) single component, 507
Fish eyes, 74, 536 (see also for migration (see Migration)
Wetting out) opportunities of, 432
Index 725

[Formulation] [Gel content]


of paper laminates/labels, 216, 490 for tackification, 116
of general purpose values of, 314
permanent 490 Glass transition temperature, 9, 75
of removable, 490 (see also Viscoelastic
on peel, 430 properties)
of permanent/removable labels, additivity of, 104
494 adjustment of, 96, 104
of PVC laminates, 495, 496 in synthesis, 104
of special PSAs, 182, 190 in formulating, 104
for porous substrates, 212, 220 with plasticizers, 104 (see
reasons of, 430 also Plasticizing)
for shear, 432 with tackifiers, 104 (see
for solid components of laminate, also Tackification)
221 dependence of
of solvent-based adhesives, 543 on chemical composition/
(see also Solvent-based structure, 96
PSA) on crosslinking, 100
for special uses, 216, 499 on flexible main chain, 98
for tack, 430 on microstructure, 100
technological considerations of, on molecular weight, 95
205, 565 on monomers, 100 (see
for thermal resistance, 176, 211 also Monomer)
in transfer coating, 221 on morphology, 102
of water-based adhesives, 445, 541 on sequence distribution/
dependence on coater length, 101
configuration, 590 on side groups, 99
special features of, 542 on tackifying resin, 97, 435
of water resistance, 218, 462, 503 factors of influencing, 92
Fox–Flory equation, 95, 106 of HMPSA, 106
Fracture energy, 269, 277 (see influence of
also Peel) on adhesive properties, 106
Functionalization, 210 on converting properties,
Fungicide, 565 (see also Formulation) 108
on end use properties, 110
Gear in die, 591, 592, 594 on peel, 108
(see also Slot die) on tack, 108
for HMPSA, advantages of, 598 of acrylic HMPSA, 100
Gel content, 164, 180, 191 314, role of, in the characterization
335, 424 of PSA, 91
interdependence of with molecular of silicone PSA, 108
weight, 122 values of, 93, 94, 95, 98, 104,
of styrene butadiene copolymers, 107, 108, 112, 159,
166 161, 166, 172, 175,
in SBR, 11 180, 184, 304, 317
726 Index

Gordon–Taylors equation, 106 [Hot melt PSA]


Grafting, 210 acrylic, 185, 200, 150, 512
Gravity bleeding (see Migration) viscosity of, 34
Gravure advantages/disadvantages of, 510
angle of, 580 coating of, 594
depth of, choice of, 579 energy consumption for, 510
coater machine of, 560
adhesive viscosity on, 579 speed of, 511 (see also
choice of cylinder for, 581, Running speed)
583 high temperature performance
coating weight of, 581, 582, of, 104
589, 590 radiation cured, 515 (see also
design of, 585 Radiation curing)
dry edge on, 574 recipe for, 543 (see also
direct, 585 (see Direct gravure) Formulation)
indirect, 585 (see also removable, 510 (see also Test
Offset gravure) methods)
induced structures (see also SBS, 126
Striation) screening of, 495
offset (see Offset gravure) stabilizers of, 510
reverse, 50 (see also Reverse tackification of (see Tackification)
gravure) viscosity of, 510, 597
for paper-based laminates, 73 water soluble, 150 (see also
roll coating, 358 Water resistance)
on film, 50, 73 weather resistant, 150
on paper, 73 Hot shear, 402 (see also Shear
on rotation (see Rotogravure) resistance)
Grit (see Coagulum) gradient of, 402
dependence of, on
High release agent, 202, 285 formulation, 402
Hold strength, twenty degree influence of, on cuttability,
(see Shear strength) 473
Holding power, 107, 117, 669 (see also Humidity, rest of
Shear resistance) influence on cuttability, 472
dependence of influence on lay flat, 372
on gel content, 311 (see also Lay flat)
on melting point of the influence on speed, 356
resin, 312 influence on peel, 289, 297
values of, 314 tolerances of, 372
Hookes law, 6, 22, 309, 316 Hydrocarbon-based tackifier, 468 (see
Hot melt compounding, 507 (see also also Resin; Tackifier)
Hot melt PSA) advantages/disadvantages of, 474
requirements for, 506 applications for, 473
Hot melt PSA, 1, 10, 40, 71, 73, dispersion of,
152 compounding of, 472
Index 727

[Hydrocarbon-based tackifier] [Knife]


foam resistance of, 472 (see also rotating roll knife, 565
Foam) rubber blanketed, 567
mechanical stability of, 472
wetting agent level for, 472 Label, 1
influence of application technology of, 26
on the adhesive properties of (see also Labeling)
PSA, 469 coating technology for, 49 (see
on aging of PSA, 472 also Coating machine)
on converting properties of deep freeze, 14 (see also
PSA, 472 Deep freeze)
level of, 472 dispensing of, 291 (see also
price of, 473, 475 Labeling)
raw materials for, 473 end use properties of, 497 (see
replacement of rosin with, 473 also End use)
Hydrogels, film, 369 (see also Film coating)
pressure-sensitive, formulation manufacturing of (see
of, 182 Manufacture)
Hydrophilic–lipophilic balance, 522 multilayer, 365
(see also Surfactants) reel, 17, 364
removable (see Removable label)
Interdependence, coating see through, 216 (see also Film
weight-solids label)
content–viscosity-coating speed, 51 sheet, 17, 364
Interfacial energy, 202 (see also Peel) requirements for, 695
Isobutylene rubber, 165 (see also stiffness of, 397 (see also Face
Polyisobutylene) stock)
Isocyanate (see also Polyisocyanate) surface quality of, 404
crosslinking agent, 490 thickness of, influence on the
adhesive properties, 80
JKR theory, 130, 264 (see also (see also Adhesive
Contact physics) properties)
wash of, 218 (see also
Kiss coating (see Coating device) Test methods;
Knife Water resistance)
of air (see also Air knife; Coating Labeling, 26, 320, 363, 559
device) floating, 596 guns, 26
holding of, 597 Laminate, 363
over roll, 565, 567, 568 build up of, 222, 365
coating weight on, 568 components of, 365
parameters of, 568 solid, 221
running speed of, 568 of film, 287
types of, 567, 568 flexural resistance of, 81
viscosity on, 567 (see also Stiffness)
roller, 567 paper, general purpose,
728 Index

[Laminate] Loss tangent delta peak, 9, 10, 110,


manufacturing of, 72 112 274 (see also
properties of, 332 Viscoelastic properties)
build up of for roll material, 26
build up for sheet material, 26 Manufacture of pressure sensitive
PVC, 517 label, 559
removable (see also Manufacture of PSA, 425
Removable label) hot-melt, 507 (see also Hot melt
roll, 364 PSA)
sheet, 364 raw materials, 425
Laminating, 559 (see also Raw materials)
properties, 108 natural, 425
Latex synthetic, 427
of natural rubber (see simultaneous, of PSA and
Natural rubber) laminate, 614
synthetic, 166 radiation curing of PSAs, 611
Lay flat, 408 (see also Test methods) siliconising trough radiation, 614
dependence of, on humidity, Manufacture of release liner, 616
372, 392 (see also Release liner)
of printed PVC, 336, 372, Mastication, 42, 116, 161, 426, 506
Layer number, 286 (see also Natural rubber)
influence of for hot melts, 512
on the stiffness (see Stiffness) Mass transfer to the web (see
on adhesive properties, 82 Roll coating)
Legginess, 299, 300, 320 Maxwell’s model, 131, 238
(see also Removability) Mechanical stability
Lifting, test of (see Test methods) of water-based PSA, 357
Line screen, 258 (see also Gravure) of water-based resin dispersions,
influence of, on coating Mechano-chemical destruction, 36
weight, 580 (see also Mastication)
Lodges theory, 7, 137 Mechano-optical tack
Loop tack, 133, 249, 648 tester, 242, 656, 657
(see also Tack) Melt flow
test of, 236 (see also Test methods) index of, 152
Loss modulus, 8, 19, 101, 251, rate of, 151
431, 440 (see Memory effect, 17, 155
also Viscoelastic Metering (see also Coating machine)
properties) bar of, 49
dependence of, on choice of, 572
molecular weight device of, 43
distribution, 116 rod of, 570, 572
frequency and temperature, roll of (see Roll coating)
112, 115 Meyer bar, 571 (see also Coating
master curve of, 13 device)
peak temperature of, 11 in film coating, 50, 56
Index 729

Micromechanics, 132 (see also [Modulus]


Contact physics) creep, 127
Migration, 26, 484 (see also Aging; influence on peel, 260, 266
Bleeding; Penetration; dependence on
Test methods) adhesive thickness, 123, 267
of acrylics, 223 chemical composition, 118
dependence of filler, 120, 121
on adhesive characteristics, 485 glass transition temperature, 9,
on coating technology, 49, 486 123
on face stock porosity, 71, 486 material characteristics, 110
on thickener molecular weight, molecular weight, 33, 112, 118
485, 486 other parameters, 120
of CSBR, 223 sequence distribution, 119
of EVAc, 223 side chain length, 119
parameters of, 71, 72 stress rate, 38
resistance to, 335 tackifier resin, 119, 117, 434
on soft PVC, 222, 373 dynamic, 127
test of, 330 factors influencing it, 110
of water-based PSA (see flexural, influence on peel, 126,
Water-based dispersions) 275
Milling, 23, 163, 426 (see also influence of, on stiffness (see
Mastication) Stiffness)
influence of, on viscosity, 161, 426, loss (see Loss modulus)
427 500 role of, for water-based
Minimum film forming temperature, acrylics, 125
30, 49, 74, 104, 108 in PSA characterization, 107
(see also Coalescence) storage (see Storage modulus)
dependence of values of, 101, 112, 115, 126, 249,
on glass transition 318
temperature, 110 Molecular weight, 91, 110
on experimental conditions, 111 adjustment of, 161
Mirror test (see Test methods) of base elastomers, 96
Mixed systems, 472 (see also distribution of, 91, 110
Crosslinking) in solution polymers, 31, 96
Mode of failure (see also Test method) of tackifier resin, 98
for peel, 284, 299, 300, 491, 650 values of, 98
for tack, 469 Molecular weight distribution,
for shear, 669, 671 influence of, on adhesion, 110
Models Mooney viscosity, 13, 22, 426
viscoelastic, 239 (see also Tack) Monomers (see also Chemical
Modulus composition)
adjustement of, 124 for acrylics, 99, 176, 177
in formulation, 125 and vinyl copolymers, 177, 178
in synthesis, 124 choice of, 99, 125, 168, 206
in crosslinking, 126 in crosslinking, 514
730 Index

[Monomers] [Offset gravure]


emulsifier, 494 differential, 577
for ethylene vinylacetate normal, 577
copolymers. 184, vertical, 577
185, 186 Oil (see also Hot melts)
hard, 99, 118, 179, 432 processing, 483, 493
influence of, on adhesion build up, as rheology modifier, 492, 508
265 (see also Adhesion compatibility of, 508
build up) influence on modulus, 509
level of, 116, 179, 184, 185 Oozing, 26, 71, 148 (see also
in plasma treatment, 375 (see Bleeding)
also Bleeding) dependence of, on viscosity, 150
polyfunctional, 180 test method of (see Test methods)
in radiation curing, 181 (see Orchards Equation, 48 (see also
also Radiation curing) Wetting out)
toxicity of, 615
ratio of, 100, 209 Paper
for shear, 179 cuttability of (see Cuttability)
soft, 179, 205 face stock of, 371 (see also
in styrene butadiene rubber, 165 Face stock)
in thickeners, 533 comparison of vs. film, 377
Mullins effect (see Stress softening) labels of (see Labels)
Multiroll coating, 533 modulus of
dependence on humidity,
Natural rubber, 161 121, 366
adhesives parameters of, 126
latex, 104, 110 printing of, 367 (see also
plateau modulus of, 110 (see also Printability)
Plateau modulus) for release liner, 482, 618
for removability, 115 (see also sensitivity of, towards
Removability) humidity, 368
compatibility with resins, 106 stiffness of, 369, 400, 406
(see also Resin) (see also Stiffness)
Neoprene latex Particle size (see also Test methods)
carboxylated, 522 for CSBR, 166
tackification of, 493, 501, 520 or water-based PSA, 45, 48,
Newtonian systems, 6, 22, 38, 292 110, 338
Non-Newtonian systems, 24, 34, 35, Peel
38, 42, 455 adhesion, 3, 55, 261 (see also
Test methods)
Offset gravure, 574, 576(see also adjustment of, 437, 438
Indirect gravure) angle of, 15, 657, 659
advantages of, 577 dependence on face stock
angular, 576 flexibility, 61
coating weight in, 577 normalized, 62, 657
Index 731

[Peel] [Peel]
influence on peel, 285 on peeling rate, 17, 20, 62, 289,
of CSBR, 326 292, 293
cyclical, 298 on primer, 307 (see also
dependence of Primer)
on adherend, 293, 295, 296, on release liner, 285, 293
297 on shape of the adhesive layer,
on adhesive geometry, 265, 270 275
on adhesive molecular on strain rate, 16, 61, 292, 297
weight, 269 on substrate, 7, 54, 294, 295,
on adhesive nature, 265, 266 296
on adhesive thickness, 271, 272 on tackification, 238
on adhesive state, 267 on temperature, 30, 61
on application pressure, 298 on viscoelastic properties, 15
on cavitation, 133 on viscosity, 15, 31
on carrier deformation, 55, energy of, 80
278, 281, 662 of EVAc, 337
on chemical composition factors of, 265
of base elastomer, improvement of, 298
266, 267 by tackifier, 306
on coating technology, 275 influence of,
on coating weight, 18, 55, 62, on other characteristics, 320
78, 271 low temperature (see Deep freeze)
on cohesion, 292, 299 measurement of, 262
on contact surface 241, 253 test rate for, 292
(see Contact surface) modifiers of,
on crosslinking, 268, 432 for water-based PSA, 532
on dwell time, 16, 18, 282 for permanent/removable
on elastic modulus, 15, 60, 61, PSA, 398
238, 281 on PET, 78 (see also Substrate)
on experimental conditions, rate of, 7, 657
288 reduction of
on face stock, 60, 61, 80, 275, by coating weight, 306 (see
284 also Coating weight)
on fibrillation, 133 by fillers, 306, 498 (see also
on filler, 300, 305 Fillers)
on glass transition by flexibilizers (see
temperature, 18 (see Flexibilizers)
also Glass transition by primers (see Primer)
temperature) by stress resistant polymers
on laminate construction, 285 (see Stress resistant
on layer number, 286, 287 polymers)
on layer structure, 285, 286 on polyethylene plate, 661
on peeling angle, 62, 281, on PVC, 661
282, 283 on release liner, 293
732 Index

[Peel] [Photoinitiators]
dependence on peel angle, compatibility of, 30, 474, 477
292, 293 for electron beam cured PSA,
of removable PSA, values of, 492 477, 478
of tackified acrylic PSA, 331 for hot melts, 478
test method of (see also Test (see also Hot melts)
methods) influence of
at high speed (see Test methods) on cohesion, 477
at 90 , 62, 272, 293 on creep, 305 (see also Creep)
at 180o, 273 on glass transition
drum peel, 666 temperature 477
T peel, 659, 661 on PVC, 669
special, 666 on tack, 477
theoretical formula of, 263 on viscosity, 478
values of, 271, 272, 273, 274, 279, level of, 221, 304, 478
306, 307 (see also Formulation)
for UV cured PSA, 309 macromolecular, 432
Peelability (see also Removability) (see also Polybutene)
dependence of, on tackifier level, migration of, 340
274 (see also Migration)
parameters of, 563 oils (see Oil)
Peel force from release liner, 285 sensitivity towards, 477
(see also Test methods) volatile, 104
dependence of for water-based PSA, 489, 492
on peeling angle, 285 Plasticizing, 104, 477
on peeling rate, 285 effect, of comonomers, 100
on test conditions, 285, 288, Plateau modulus, 16, 110, 127
289 (see also Viscoelastic
values of, 600, 665, 698 properties)
pH, 429, 443 dependence of
adjustment of, 538 on formulating additives, 187
influence of, on tackification of (see also
water-based PSA, 520 Formulating)
Phase separation temperature, 171, on sequence distribution, 16
214 (see also Tackifica- on tackifier resin, 125, 477
tion) influence of, on tackifying,
Penetration, 70 (see also Migration) 119 (see also
dependence of, on modulus, 111 Tackification)
Photoinitiators, 197, of SBR, 119
copolymerizable, 195 of SBS, 119
Pipelines, for water-based PSA, Polishing bar (see Coating machine)
540 Polyacrylate rubber, 168, 178
Plasticizer, 30, 31, 104, 162, 188 (see also Acrylics;
477 Raw materials)
choice of, 498 Polyalphaolefin, amorphous, 168
Index 733

Polyamide, face stock, 294, 360 Polyolefin


Polyaziridine, 192, 304 amorphous, 168
(see also Crosslinking) face stock of, 214, 365, 371, 375
Polybutadiene, 215, 216 Polypropylene
Polybuthene (see also atactic, 168
Butylene rubber) face stock of, 78, 370, 376
aging of, 491 comparison vs. polyethylene,
as tackifier, 478, 486, 487 376, 377
Polycarbonate, face stock treatment of, 376
(see Face stock) release liner of, 617
Polyester surface treatment of, 376
as face stock, 283, 365 Polystyrene
(see also Poly- compatibility of, 106
ethyleneterephtalate) face stock of, 365, 385
as raw material for HMPSA, 195 Poly(styrene-butadiene-styrene),
Polyethylene 172, 173
additives (see Formulation) in hot melts (see Hot melts)
face stock of, 365, 370, 372, 375 sequence distribution of, 119
treatment of, 376, 377, 378 influence on cuttability, 119
(see also Treatment) polystyrene content of, 119
release liner of, 376 Poly(styrene-ethene-buthene-styr-
Poly(ethylene-butylene), ene), 171, 172
crosslinking of, 116 aging resistance of, 175
(see also Thermoplastic compatibility of, 503
elastomers) in hot melts, 503 (see also
Polyethyleneterephtalate Hot melts)
as face stock, 283 Poly(styrene-isoprene-styrene), 172,
as release liner, 617 173,
as standard surface face stock, 78, aging of, 312
283, 645, 648 crosslinkable on EB, 103, 104,
Polyisobuthylene, 170, 483 171, 268
aging of, 491 in hot melts, 503 (see also
for removability, 491 Hot melts)
Polyisocyanate, 494 (see also molecular structure of, 100
Crosslinking) polystyrene content of, 172
Polyisoprene, molecular weight of, sequence distribution of, 100, 119
93, 96, 101, 118, 172 tack of, 245 (see also Tack)
Polyken tack, 14, 246 (see also Polytack (see Test methods)
Test methods) Polyurethane, 176, 177
Polymer analogous reaction, 210 crosslinking agent of, 498
Polymer synthesis, 104 Polyvinylacetate, 30, 494
comparison of, for PSA raw dispersions, 92, 98 (see also
materials, 205 Water-based dispersions)
of styrene block copolymers, 171, Polyvinylethers, 182, 493
175 solubility of, 182, 219
734 Index

[Polyvinylethers] [Pressure-sensitive adhesive]


solvents for, 182 history of, 1, 162, 171, 175, 184,
in tackifier, 430, 446 613, 621, 650, 657
resistance of laminate, definition and
to aging, 182, 503 construction of, 364
to plasticizer, 182 -like network, 192
Polyvinylchloride, manufacturing of (see
coating of, 474, 616 Manufacture of PSA)
face stock of, 104, 162, 273, permanent, 176, 297
301 365, 369 physical basis of for the
advantages/disadvantages of, viscoelastic behavior, 89
371, 372 (see also Viscoelastic
comparison of vs. polyolefins, 375 properties)
flagging of, 374 (see also removable, 39, 298 (see also
Flagging) Removable adhesive)
plasticized, 78, 340, 370, 371 silicone-based, 162
properties of, 372 adhesive characteristics of, 162
shrinkage of, 373 single component, 175
(see also Shrinkage) viscoelastic properties of, 2
surface tension of, 372 (see also Viscoelastic
(see also Surface tension) properties)
versatility of, 376 Pressure sensitive laminate
Polyvinylpyrrolidone, thickener, converting characteristics of, 80
486, 533 (see also Convertability)
Polytack, 449 (see also Tack) Pressure-sensitive labels, 1
Postadditives, 522 (see also (see also Labels)
Formulation) Pressure-sensitive tapes, 1
Precursor, (see also Tapes)
polymerizable, 195 Pressure-sensitive protective films, 1
of UV curable hot-melt, 194, 196 (see also Protective
Premetering, 587 (see also films)
Coating device) Pretreatment (see Treatment)
Prepolymers, Primer, 29, 300, 326, 339, 331
of acrylate functionalized, 181 coating weight of, 304
Pressure, of application, 17 influence of
Pressure-sensitive adhesive, 1 on peel, 301
(see also Adhesives) on removability, 301, 494
acrylic (see Acrylic adhesives) on PVC, 301
adhesive properties of, 3 recipes of, 301
chemical composition of, 161 Printability, 59, 300, 367, 375
(see also Chemical dependence of
composition) on antioxidants, 512
definition of, 1 on defoamers, 536
formulating of, 425 (see also on surfactants, 526
Formulating) of paper (see Paper)
Index 735

[Printability] [Recycling]
of polyolefins, 300, 376, 375 constructive, 602
of PVC, 380 technological, 603
Probe tack, 248, 651 (see also of solvents, 603 ( see also Solvents)
Polyken tack; Test by absorbtion, 604, 605, 606
methods) by condensation, 604, 606
fibrillation in, 112, 241 by afterburning, 604
tester of, 650 of PSPs, 607
Protective colloids, 520, 533 film based, 607
Protective films, 16, 190, 320, 426, 619 paper based, 608
adhesive properties of, 3 Reel stability (see Converting
laminating conditions of, 21 properties)
release coating for, 617 Relaxation, 28, 240
Pumps, for water-based PSA, 540 Release
coating, 199
Quick stick, 23, 111, 648 aqueous, 204
(see also Tack) catalyst for, 201
test of (see also Test method) radiation cured, 203
values of, 170 silicone-based, 200, 203
silicone free, 204, 205, 620
Radiation curing, 193, 192, 195 solvent-based, 203
advantages of, 59 agents of, 620
of PSA, 613 controlled, 201, 620
of silicone, 614 differential, 201
electron beam, 615 (see also force
Siliconizing) adjustment of, 490
ultraviolet, 616 (see also aging test of (see Test methods)
Siliconizing ) dependence of on adhesive nature,
Raw materials (see also 410, 490
Manufacture of PSA) dependence on paper thickness
curable, 193 (see also Curing) 615
of film labels, 216 liner, 388, 389
of HMPSA, 211 conditioning of (see Test
of permanent/removable methods)
labels, 216 film-based, 559, 618
of PSA, 162 humidity of, 390
of reel/sheet label stock, 216 manufacturing of (see
of radiation-cured PSAs, 214 Manufacture)
of resins, 487 nature of, 617
of solvent-based PSA, 35, 213 wettability of (see Test methods)
of water-based PSA, 213 Remoisturization, 487, 621
Readhering adhesives, 304, 651 Removable PSA, 17, 55
problems of, 306 elastomers in, 491
Recycling, migration of, 308
of product components, 602 peel values of, 298
736 Index

[Removable PSA] [Resin]


resins in, 28 aging of (see Aging; Test
shear values of, 306 methods)
Removability, 15, 27, 27, 28, 275 compatibility of, 429, 436, 443
(see also Test methods) converting properties of, 468,
criteria of, 27, 297, 300 472
dependence of hydrogenated, 428, 459, 490
on adhesive anchorage, 299 liquid, 430, 476, 477, 490
on creep compliance, 15 manufacture of, 426
on energy dissipation, 298 mixtures of, 439
on face stock flexibility, 299 natural
on viscosity/modulus, 29, 38 manufacture of, 426
formulation for, 30 tackifying formulations of
problems of, 306 nature of, influence on adhe-
of water-based PSA (see sive properties, 430
Water-based dispersions) level of (see Tackifier level)
Repositioning, 17, 485 molecular weight of
Residence time, 490 (see also Hot influence on
melt compounding) modulus, 113
Resin (see also Chemical basis; molten, 453, 484
Tackifier; Tackification) phenolics, 439, 459
acid, 447, 455, 463 polyterpene, 458, 459, 476
acidity of, influence on reactive, 459, 476
anchorage, 469 rosin (see Rosin resin tackifier)
aging stability of, 426 rosin acid (see Rosin resin
aliphatic, 429, 434, 459 tackifier)
aromatic, 429 rosin ester, 443 (see also Rosin
compatibility of, 106, 125 tackifier)
for beam curing, 194 (see for CSBR tackification, 167
also Beam curing) softening point of, 447
choice of, 434 softening point, influence on
colofonium derivatives, 188, adhesive properties of
430, 446, 476 (see also PSA, 432
Rosin) solutions, 454
compatibility of, 458 synthetic, manufacture of, 427
concentration of, 39 (see also tall oil, 431
Tackifier level) terpene phenol, 436
influence on modulus, 119, 115 Resistance
coumarone-indene, 476 to light, 216
cyclo-aliphatic, compatibility of, mechanical, 156
107 to migration (see Migration;
hydrocarbon-based, 314, 434, Test methods)
468 (see also thermal
Hydrocarbon of acrylics, 170
resin-based tackifier) formulating for, 491
Index 737

[Resistance] [Rheology of pressure sensitive


of natural rubber adhesives, 315 laminate]
to plasticizer, 216 on solid components of the
(see also Test methods) laminate, 57
to rewetting, 528 Ribbing (see Striation)
Reverse adhesion test, 281 Ring and ball
(see also Peel) softening point, 440, 486
Reverse gravure test method of, 496 (see also
coating, (see also Coating Test of softening point)
device) Rod coating, 571 (see also Meyer bar)
with closed chamber, 583 Roll
coating weight in, 564, 589 coating, 575 (see also
components of, 645 Rotogravure)
with doctor blade methods of, 576
(see also Doctor blade) coating weight regulation
coating weight in, 582 of, 579
viscosity in, 582 coating weight values of, 579
running speed of, 588, 590, 591 with Doctor blade, 583
for water-based dispersions, 588 of HMPSA, 567, 594
Reverse roll coating, 456, 564, 589 (see of solvent-based PSA, 575, 585
also Dahlgren system) Rolling ball,
rotogravure (see also (see also Test methods)
Rotogravure) dependence of, on softening point
coating weight in, 589 of the resin, 454
for HMPSA, 567 tack, 252 (see also Tack)
viscosity values of, 42, 590 Rolling cylinder (see also Test
running speed in, 589 methods)
viscosity in, 589 comparison of, vs. rolling ball, 643
Rheology of pressure-sensitive Rosin resin tackifier, 459, 442, 444
adhesive, 5 acid, 462
bonded, 44 in acrylics, 464
dispersions, 40, 44 dispersion of (see also Tackifier
solutions, 40, 41, 42, dispersions)
dependence on adherent, 44 wetting agent level in, 466
dependence on polymer, 42 ester, 481
dependence on solvent, 43 compounding of, 482
dependence on technology, 43 cuttability of, 457, 458
special features of, 42 in natural rubber, 449
uncoated, 5 price level of, 469
Rheology of pressure sensitive Rotating drum (see Tack test
laminate, 52 method)
dependence of Rotogravure, 531
on composite structure, 80 direct, 567
on liquid components of for silicones, 615
the laminate, 53 speed of, 531
738 Index

Rubber, natural (see also Elastomer) [Shear modulus]


-based PSA (see Rubber/resin dynamic, 23
adhesives) Shear resistance, 3, 46, 306
manufacture of, 426 of acrylics, tackified, 335
mastication of (see Mastication) of automotive PSA (see
Rubber/resin adhesives, 42, 106, Test methods)
110, 162, 425 (see also dependence of
Solvent-based PSA) on adhesive/cohesive
solvents for, 44 (see also properties, 309
Solvents) on adhesive nature, 310
Running speed, parameters of, 358 on coating weight, 314
on composite status, 314
SAFT (see also Shear resistance; on crosslinking, 126, 246, 313
Test methods) on face stock, 66, 316
dependence of on modulus, 23
on face stock surface, 661, on molecular weight, 312, 313
672 on sequence distribution,
on softening point of the 315, 316
resin, 320 on softening point of the
on tackifing, 449 resin, 453
Sandwich structure, 370 on strain rate, 34
(see also Laminate) on substrate, 316
Segregated elastomers, 171 on tackifying (see
Self-healing, 16 Tackification)
Self reinforcement, 117 (see also on temperature, 25
Natural rubber) on tensile strength, 309
Separation energy, 7, 11, 16, 18, on time, 24
27 (see also Peel on viscoelastic properties, 21
adhesion) on viscosity, 22
dependence of, factors of, 310
on surface tension, 80 improving of, 107, 318
as tack, 236 influence of
Separation rate, 304 (see also Peel on cuttability
adhesion) (see Cuttability)
Sequence distribution, 246, 247 measurement of, 308
Shaft loaded blister test, 63, 283, as Williams plasticity, 309
291, 668 (see also Williams
Shear, on the coating machine, 44, plasticity)
356, 483, 572, 589, of solvent-based PSA, 24
Shear modulus, 8, 23, 44, 126, 249, test method of (see also
284 Test method)
calculation of, 117, 126 dead load hot strength, 308
dependence of, on crosslinking, dynamic, 23, 308
119 (see also hot, 308, 309, 332, 402
Crosslinking) room temperature, 307
Index 739

SAFT, 309 (see also SAFT) [Silicone]


static, 307 water-based, 621
20 hold strength, 307 Siliconizing
values of, 56, 298, 307, 309, 311, coating machines of, 619
322, 323 for solventless systems, 620
Shear thinning, 52 (see also Shear speed of, 620, 621
on the coating machine) by radiation (see also
Shrinkage (see also Plasticizer Radiation curing)
resistance; Test methods) coating speed of, 615, 616
of acrylics, 336 electron beam, 615
dependence of, on printing, 373 ultraviolet, 616
of face stock materials, 336, 372, viscosity in, 614
373 technology of, 39, 616
parameters of, 373 energy consumption of, 620
of PVC, 373, 375 for solvent-based systems, 620
dependence on plasticizer for solventless systems, 620
migration, 373 Slip agents
dependence on relaxation, 373 in polyolefins, 385
dependence on technology, 372 migration of, 386
values of, 336, 683 Slot die, 565, 566, 568, 575
Sieve residue, test of (see Test for HMPSA, 565
methods) advantages/disadvantages
Silicone of, 591
acrylics, 587 Smearing
fluids, 390 influence of viscosity on, 587
defoamers, 535 (see also on cuttability, 307 (see
Formulating) also Cuttability)
pressure-sensitive adhesives, 162 on relaxation, 400
tack of, 246 on tack (see Tack)
peel of, 267 parameters of, 55
release, 199 Smoothing bar, 432, 571, 565
addition cured, 200 (see also Coating
catalyst system of, 201, 618 machine)
coating technology of, 203 (see Softening point
also Siliconizing) of the adhesive, test of (see
coating weight of, 398, 614, 615, Test methods)
616, 618 of the resin, test of (see
crosslinked, 203 Test methods)
emulsions, 204 Solids content
moisture cured, 200 of solvent-based PSA, 44, 548, 586
postcure of, 203, 479, 614 of water-based PSA, 213, 620
radiation cured, 203, improvement of, 454
solvent-based, 203 Solubility
solventless, 46, 200 diagram of, 444
UV cured (see Radiation curing) parameter of, 77, 78
740 Index

Solvent, 354, 506, 536, 538 Storage


hydrocarbon-based, 44 modulus, 8, 10, 15, 110, 112, 443
influence of (see also Viscoelastic
on anchorage (see Anchorage) properties)
on water resistance, 562 influence of, on peel, 397
recovery of, 501, 538, 600 values of, 10
in solvent-based PSA, 501 tanks (see Tanks)
in WBPSA, 354, 437, 540 of water-based PSA (see
Solvent-based PSA Water-based dispersions)
acrylics Stress
advantages/disadvantages of, softening, 156
542, 548 (see also restricting polymers, 300 (see
Acrylics) also Removability)
anchorage of, 77 resistant polymers, 304
coating machine for, 542 Striation, 455, 549, 573
(see also Coating Strike through (see Penetration)
machine) Stringing (see Legginess)
special features of, 252 Strip coating, 591
equipment for, 569 Stripe structure, 357 (see also
molecular weight of, 42, 99 Structure)
roll coating of Structure, textured, 357
running speed of, 587 Styrene butadiene rubber, 166
solids content of, 80, 493, 587 (see also Raw materials)
surface tension of, 42 carboxylated (see Carboxylated
viscosity of, 42 styrene butadiene
Stability rubber)
dimensional (see Shrinkage) compatibility of, 10, 106
thermal, 375 DMA of, 10
test of, 121 latex of, 166, 167
Stabilizers in HMPSA glass temperature of, 167
(see Formulation) shear of, 168
in PVC, 222 solids content of, 168
for ultraviolet, 502, 512 styrene content of, 166
Staining, 434, 459, 547, 671 (see tack of, 119, 436
also Migration; tackified, peel of, 187
Penetration) Styrene copolymers, 167, 172
Stationary roll, 574, 588 (see also (see also Thermoplastic
Coating device) elastomers)
Stiffening molecular weight of, 172
effect of, 7 styrene content of, 172
of face stock, 683 tackifier for, 440 (see also Hot
Stiffness, 275, 281, 286, 320, 365 melts; Tackifying)
of laminate, 80, 432 Substrate, 65
parameters of, 391, 428, 429, 430 elasticity/plasticity balance of
of liner, 415 influence on peel, 65
Index 741

[Substrate] [Surfactants]
for laboratory tests, 648 interaction of, with other layers
surface of, influence on peel, 68 of the laminate, 526
for test, 298 level of, 527, 528, 541 parameters
of peel, 298, 670, 675 of, 528, 530
of tack, 648 migration of, 526 (see
Sulfosuccinates, 524 (see also also Migration)
Surfactants; Wetting plasticizing effect of, 526
out) as thickeners, 531
level of, 525 Synthesis,
solubility of, 524 in line,
thickeners, 468 components for, 189
Supplier, selection of (see oligomers for, 195
Tackifier dispersion) polymers for, 196
Surface influence of,
energy of, 6 on chemical composition, 205
values of, 73, 74,
standard, for loop tack, 236 Tack, 3, 12, 15, 53, 235 (see also
textured, 44 (see also Structure) Test methods)
Surface tension, 41, 48, 74 level of, 243
dependence of application, 249
on viscosity, 48 (see also factors of, 244
Orchard’s equation) in acrylic block copolymers, 440
dynamic, 48, 521, 633 of acrylics, 325
of fluorsilicones, 170 deadeners (see Detackifying
of release coatings, 200 agents)
influence of dependence of
on foam (see Foam) on adhesive nature, 245
on shear, 326 on coating weight, 245, 246,
on wetting, 41, 357 247, 248
of PVC, 386 on creep compliance, 110, 127
of SBR dispersions, 47 on crosslinking, 246 (see
static, 48 also Crosslinking)
values of, 47, 170, 366, 372, 378 on experimental parameters, 13
Surfactants, 48, 207, 218, 520, 527 on face stock, 249
(see also Emulsifier; on fillers, 247
Wetting agent) on gel content, 96, 122, 165,
anionic, 541, 520, 521, 245
choice of, 520 on glass transition tempera-
fluorinated, 526, 528 ture, 242
influence of on loss tan d peak, 14
on drying, 528 (see also on manufacture, 246
Composite structure) on modulus of elasticity, 13, 236
on the properties of PSA on molecular weight, 13, 245
laminate, 527, 528 (see Chemical basis)
742 Index

[Tack] [Tack]
on the properties of PSA as separation energy, 13, 80, 241
on plasticizer, 244 (see on polyethylene, 253
also Plasticizing) relative, 241
on rheology, 24 of rubber resin adhesives, 436
on sequence distribution, test method of, 236
246, 247 as loop tack, 236, 241, 243
on strain rate, 14 (see (see also Loop tack)
also Viscoelastic as rolling ball, 236 (see
properties) also Rolling ball)
on surface active agents, 247 as rolling cylinder, 236
on surface tension, 249 (see also Rolling
on tackifier, 245 (see also cylinder; Test
Tackification) methods)
on tackifier melting values of, 246, 248, 249, 250
point, 246 (see Werle tester of (see Test methods)
also Resin) wet, 219, 508
on test method and conditions, Tackification, 106, 125, 432
249 (see also of acrylics
Test methods) ease of, 233 (see also Tackifying)
on time/temperature, 14, water-based (see Water-based
248, 249 (see also adhesives)
Rheology) for cost reduction, 434
on viscoelastic properties, 11 of CSBR, 117
on viscosity, 12, 13 dependence of, on the physical
on wetting out, 250 state
energy of, 103, 115, 137, 242 of the tackifier, 460
improvement of, 250, 326 of hot melt PSA, 455, 436,
(see also Tackification) 429
index of, 245 limits of, 460, 461
influence on cuttability copolymers,
(see Cuttability) with plasticizer, 473
of acrylics, 243, 244 with resins, 432
comparison of, 244 liquid, 542
of EVAc, 244, 329 molten, 542
of SBR, 244 solution, 454
of uncoated PSA, 245 for shear, 433
measurement of, 236 (see also with solvent-based tackifier, 542
Test methods) special features of, 454
discrepancies of, 651 of water-based PSA, 454, 541
as coefficient of friction, 236 with water-based resin
as cohesion, 241 dispersion, 541
as fracture energy, 242 Tackifier, 162, 188 (see also
as peel, 19, 243 Plasticizer; Resin)
as plasticity, 243 in acrylic block copolymers, 436
Index 743

[Tackifier] Tackifying
choice of, 434, 437 ability of, 434
concentration/level of ease of, 434
(see Tackifier level) response, 516
in CSBR, 436, 436 of acrylics, 324
degradation of, 222 of CSBR, 324, 325
dispersion of, 216 of EVAc, 324, 325
requirements for, 465, 478 of water-based PSA
in electron beam crosslinkable (see Water-based
SIS, 437 dispersions)
in ethylene vinylacetate Tan d, 8, 11, 23, 30, 270 (see
copolymers, 432 also Rheology)
(see also Formulation) Tanks, for water-based PSA, 540
hybrid, 474 Tapes, 1, 10, 24, 78, 221
hydrocarbon resin-based adhesive strength of, 169, 307
(see Resin) aging of, test of, 666, 682
influence of block copolymers for, 120, 170
on converting (see Converting coating technology of, 357, 459,
properties) 564, 588, 613
on end use properties (see coating weight of, 73
End use properties) double sided, test of, 564, 620
on shear (see Shear) health care, 221
in natural rubber, 435, 436 masking, 645
in polyvinylacetate, 439 of polyvinylchloride, 300
rosin resin-based (see Rosin) primer for, 300
Tackifier level, 430 wet adhesive, test of, 657
in acrylics, 323, 431 TAST, 312, 677 (see also Shear
dependence of, on elastomer measurement)
nature, 436 Test methods, 602
in electron beam cured PSA, axis-symmetric, 654 (see also
436, 429 Contact mechanics)
in ethylene/maleinate of accelerated weathering, 682
copolymers, 436 of adhesive nature, 633
in natural rubber, 436, 432 of adhesive properties, 644, 689
in polybutadiene, 436 adhesion at low temperature, 699
in SIS rubber, 436 (see also Deep freeze
influence of labels)
on migration, 542 (see combined, 657
also Migration) evaluation of, 677, 678
on peel, 431, 433 reference material of, 644
on shear, 432, 433 standards for, 644
on tack, 436 specimen preparation for, 644
for screening, 436 of adhesive residue 687 (see also
with liquid resins, 542 Test of removability)
with molten resins, 542 of aging properties, 681, 682
744 Index

[Test methods] [Test methods]


evaluation of adhesive of lifting, 688
of degradation, 681 of liquid adhesive, 629
evaluation of migration, 682, of migration, 683
683 (see also Migration) evaluation of, 682, 683
of PVC face stock, 683 (see also Migration)
of release, 682 (see also of peel strength, 78, 243, 306,
Release liner) 657, 684, 691
standard conditions in, 681 on cardboard, 665
of coagulum (see Coagulum) with constant force, 647
coating conditions for, 646 at high speed, 662
of coating weight, 639, 681 of 90 peel adhesion, 660
of cold flow, 642 plasticized PVC film, 666
of compatibility, 431 on polyethylene plate, 666
of compliance, 642 from release liner, 295, 662
(see also Approval) standard, 661
conditions of, 646 on standard substrate for,
of contact angle, 633 295
of deep freeze labels, 698 of T-peel, 661
(see also Deep freeze) test conditions of, 288
of cold tack, 699 of plasticizer resistance, 683
of low temperature adhesion, 699 of readherability, 688
of static shear load test at of release liner, 17
low temperatures of, 699 of removable labels, 686
of dimensional stability, 683 of removability, 686
drying conditions for, 637, 638 of HMPSA, 687
of cuttability, 440, 457, 685 by water, 695
of edge ooze, 685 of repositionability, 687
of film coating, 686 of resistance to humidity, 692
adhesive properties of, 689, of shear strength, 56, 308, 645
adhesive screening in, 687, automotive, 677
690, 691 dynamic, 674
adhesive versatility in, 689 dynamic lap shear, 673, 675
loss of transparency of, 692 experimental conditions for,
water whitening of, 692 645, 675
wet adhesion of, 692 hot shear, 673, 691
wet anchorage of, 693 product related, 677
of foam, 634 shear adhesion failure
height of, 634 temperature of, 673
of grit content, 637 standard, 671
of HMPSA, 630 statical/room temperature,
of humidity resistance, 687 671
of laminate properties, 643 twenty degree hold, 675
general, 644 Weyerhousers, 697
special, 645 of shrinkage, 583
Index 745

[Test methods] [Test methods]


of sieve residue, 636, 637 screening of, 638
of softening point, 640 solids content of, 639
of solid adhesive, 639 wetting characteristics of,
of solvent-based PSA, 630 633, 689,
standard, 629 stability to, 339
of surface tension, 633 of water resistance, 690, 695
dynamic, 633 of water solubility, 696, 698
of tack, 14, 524 (see also Tack) of water whitening, 496 (see also
Bulls tack, 655 Water whitening)
Druschkes probe tack, 650 of wettability, of release liner, 633
Hammonds probe tack, 650 of wetting angle, 633
Kendalls probe tack, 650 for adhesive extract, 635
loop tack, 650 (see of wetting out, 632
also Loop tack) direct, 632
Polyken tack, 14, 652 indirect, 632
quick stick, 14, 647 vinyl wetting test of, 632
(see also Quick stick) of Williams plasticity, 640
rolling ball, 14, 653 Thermoplastic rubbers, 125, 167, 172
(see also Rolling ball) compatibility of, 125
rolling cylinder, 691 in raw materials for hotmelts,
(see also Rolling 118 (see also Hot melts)
cylinder) Thick adherend shear test, 56
toothed wheel, 655 (see also Test methods)
Werles, 657 Thickeners, 354, 529
Wetzels probe tack, 650 Thickening, dependence of, on
of tensile strength, 650, 651 pH, 529
of viscosity Thixotropy, 357, 529
for hot melts, 630 index of, 66
for solvent-based PSA, 630 Time-temperature superposition
for water-based PSA, 631, principle, 9, 11, 24, 25
654, 689 (see also Transparency, loss of, test of
Thixotropy, index) (see Test methods)
of washing machine resistance, Treatment, of surface
73 chemical, 378
of wash off adhesives, 696 Corona, 78, 372, 373, 619
of water-based PSA, 631 flame, 378
adhesive properties of, 689 fluor, 378
coating properties for, 631 influence on shear, 383
product related, 679 plasma, 376, 377, 378
coating weight of, 639 shelf life of, 376, 377, 378
density of, 639 Two roll gravure coater
drying ability of, 638 (see Reverse gravure)
mechanical stability of, 636 Two sided coating (see Offset
particle size of, 638 gravure)
746 Index

Unwinding resistance, 21 [Viscosity]


Ultraviolet on coating versatility, 49
drying (see Drying) on coating weight, 50
stabilizers of rubber/resin on wetting, 49
adhesives, 236 of polymer solutions, 31, 42
(see Formulation) time/temperature dependence
of, 44
Versatility, of solvent-based PSA of PSA dispersions, 44, 50
(see Solvent-based PSA) factors of, 50
Viscoelastic models, 239 stability of, 358
(see also Tack) values of, 38
Viscoelastic properties, 5 of acrylic HMPSA, 212
dependence of on chemical com- of HMPSA, 34, 212, 221
position/structure, 33 on knife over roll, 221
experimental and environmental on Meyer rod, 221
conditions, 34 in rotogravure, 221
material characteristics, 31 Viscous components, for PSA, 185
time, 36 Voigts model, 265 (see also
factors of, 31, 402 Viscoelastic properties)
influence of on
adhesive properties, 12 Warm melts, 359
applied label, 31 Water-based dispersions, 515
converting properties, 25 of acrylics, 175 (see also
end use properties, 26 Acrylics)
physical basis of, 89 modulus of, 125, 356
Viscoelastomers, 176, molecular weight of, 96
acrylic, 177 special features of, 254
Viscosity, 6 coating machine of, 565, 587
adjusting of, 529 coating weight of, 587
dependence of coating weight tolerance of, 587
on chemical composition, 33 drying of, 358
on molecular weight, 31 formulation of (see Formulation)
on shear rate, 37 properties, 505
on storage time, 258 (see also ready to use, 44
Water-based requirements for, 587
dispersions) shelf life of, 44
on tackifier resin, 34 special additives in, 516
on temperature, 34 viscosity of, values of, 358
of base elastomers, 31 Water resistance, 216, 496
of HMPSA, 221 Water whitening, 223, 496
of PSA dispersions, 44 (see also Test methods)
of solutions, 31, 39 Weathering performance, 459
influence of Web
on adhesive coating weight control of, 565
(see Coating weight) tension of, 560
Index 747

Wet elongation, 392 (see also Window of performance, 9, 10, 432,


Lay flat) 458 (see also Viscoelastic
Wettability properties)
of tackifier dispersions, 456 Williams–Landel–Ferrys equation,
of release liner (see Test methods) 11, 25, 35, 36, 309
Wetting agents, 520, 521 (see also Williams plasticity, 152, 243, 309 (see
Surfactants) also Test methods)
choice of, 524 Wing up (see Flagging)
level of, 472 Wire wound rod, 571 (see also Meyer
Wetting out, 12, 45, 356 bar; Rod coating)
angle of, test of (see Test methods) coating weight on, 572, 573
dependence of running speed of, 573
on face stock, 74 viscosity for, 572
on viscosity, 530 Work of
on pH, 529 adhesion, 7, 77, 79, 80
dynamic, 47, 48 deformation, 8
parameters of, 356 detachment, 61
of solvent-based PSA, 354 Wrap angle, 573
static/dynamic, 47, 356
theoretical basis of, 46 Young modulus, 6, 133, 135, 243,
of water-based dispersions, 46, 267, 272

S-ar putea să vă placă și