You are on page 1of 316

Structural Adhesive Joints in Engineering

Structural Adhesive Joints


in Engineering

ROBERT D. ADAMS
BSc (Eng.) (Lond.), PhD(Cantab.), ACGI, C Eng., FIMechE, FInstP
Reader in Mechanical Engineering, University of Bristol, UK

and

WILLIAM C. WAKE
MSc, PhD, DSc(Lond.), FRSC, FPRI, Hon.DSc (The City University)
Hon. Fellow, The City University, London, UK

ELSEVIER APPLIED SCIENCE PUBLISHERS


LONDON and NEW YORK
ELSEVIER APPLIED SCIENCE PUBLISHERS LID
Crown House, Linton Road, Barking, Essex IGll 8JU, England

Sole Distributor in the USA and Canada


ELSEVIER SCIENCE PUBLISHING CO., INe.
52 Vanderbilt Avenue, New York, NY 10017, USA

British Ubrary Cataloguing in Publication Data

Adams, R. D.
Structural adhesive joints in engineering.
1. Adhesive joints
1. Title II. Wake, William C.
624.1'89'9 TA492.A3
ISBN-13 :978-94-0 10-8977-7 e-ISBN-13 :978-94-009-5616-2

DOI: 10.1007/978-94-009-5616-2

WITH 30 TABLES AND 125 ILLUSTRATIONS

© ELSEVIER APPLIED SCIENCE PUBLISHERS LID 1984


Reprinted 1986

Softcover reprint ofthe hardcover 1st edition 1984

Special regulations for readers in the USA


This publication has been registered with the Copyright Clearance Center
Ine. (CCC), Salem, Massaehusetts. Information ean be obtained from the
CCC about conditions under which photoeopies of parts of this publieation
may be made in the USA. All other copyright questions, incIuding photo-
copying outside of the USA, should be referred to the publisher.

All rights reserved. No part of this publieation may be reprodueed, stored


in a retrieval system, or transmitted in any form or by any means,
electronic, mechanical, photocopying, recording, or otherwise, without the
prior written permission of the publisher.
Preface

The intention of this book is that it should contain everything an


engineer needs to know to be able to design and produce adhesively
bonded joints which are required to carry significant loads. The advan-
tages and disadvantages of bonding are given, together with a sufficient
understanding of the necessary mechanics and chemistry to enable the
designer to make a sound engineering judgement in any particular
case.
The stresses in joints are discussed extensively so that the engineer
can get sufficient philosophy or feel for them, or can delve more deeply
into the mathematics to obtain quantitative solutions even with elasto-
plastic behaviour. A critical description is given of standard methods of
testing adhesives, both destructively and non-destructively. The essen-
tial chemistry of adhesives and the importance of surface preparation
are described and guidance is given for adhesive selection by me ans of
check lists. For many applications, there will not be a unique adhesive
which alone is suitable, and factors such as cost, convenience, produc-
tion considerations or familiarity may be decisive. A list of applications
is given as examples.
The authors wish to increase the confidence of engineers using
adhesive bonding in load-bearing applications by the information and
experience presented. With increasing experience of adhesives en-
gineering, design will become more elegant as weH as more fitted to its
products.

R. D. ADAMS
W.C.WAKE
Acknowledgements

Acknowledgements are necessarily and gladly made by the authors for


help, advice and criticism to Dr J. A. Harris of Bristol University, Dr
A. J. Kinloch of the Ministry of Defence (PERME), Waltham Abbey,
and Dr P. Poole of the Royal Aircraft Establishment, Farnborough.
Former colleagues and students have contributed in the past to the
work on which this book is largely based, in particular Drs N. A.
Peppiatt, J. Coppendale and A. D. Crocombe.
The publisher joins with the authors in thanking those individuals
and Journals listed below for permission to use diagrams and photo-
graphs which have either appeared elsewhere or are modifications of
published material:
Adhäsion (Berlin) for Figs 99-101, 110 and 111 from papers by the
late Prof. Dr-Ing. A. Matting.
K. W. Allen and J. Adhesion for Figs 107 and 108.
Dr W. Althof and Metall (Berlin) for Figs 116, 119 and 120.
Dr H. A. Burgman and J. Appl. Polym. Sei. for Fig. 118.
Dr J. Cotter and J. Adhesion for Fig. 115.
Prof. Dr-Ing. F. Eichhorn and Adhäsion for Fig. 114.
Engineering Sciences Data Unit Publication, ESDU 79016 for Fig.
33.
Dr L. J. Hart-Smith for Figs 32, 34-36 and the Douglas Aircraft
Co., for Figs 37 and 38.
The Institution of Civil Engineers for Figs 54 and 56.
J. Adhesion for Figs 27-30, 69 and 87.
J. Strain Anal. for Figs 10, 13-23 and 63-68.
Dr H. Schonhorn and J. Appl. Polym. Sei. for Fig. 92.
Dr M. E. R. Shanahan and The City University for Fig. 98.
Dr B. Wargotz and J. Adhesion for Fig. 102.
Figures 1 and 124 are Crown Copyright and Fig. 109 is based on data
supplied by the Royal Aircraft Establishment, Farnborough.
vi
Contents

Preface v

Acknowledgements . vi

Chapter 1. INTRODUCrION 1
Joint Configurations: Lap-shear Joints, Butt Joints,
Fillets. Metals and Other Constructional Materials.
The Decision to Use Adhesive Bonding. The Balance
of Advantages and Disadvantages

Chapter 2. THE NATURE AND MAGNITUDE OF


STRESSES IN ADHESIVE JOINTS 14
Introduction: Reality, Methods of Mathematical
Analysis. The Single Lap Joint: Linear Elastic
Analysis, Volkersen's Analysis, The Analysis of Go-
land and Reissner, Effect of Bending in a Double-lap
Joint, Volkersen's Second Theory, Later Work. The
Single-lap Joint-End Effects: Reduction of Stress
Concentrations. The Single-Iap Joint-Elasto-plastic
Analysis. The Effect of Adherend Shape-Scarfed,
Bevelled and Stepped Adherends. Composite Ma-
terialls. Tubular Joints. Butt Joints. The Use of Joints
in Design: Lap Joints, Tubular Joints, T -joints,
Corner Joints, Butt Joints, Stiffeners, Doublers,
Assembly

Chapter 3. STANDARD MECHANICAL TEST PROCE-


DURES . . . . . . . . . . . . . . . . . . 115
Destructive Testing: Tests with Thin Sheet
Adherends, Tests for Properties of Adhesives. Non-
vii
viii CONTENrS

destructive Testing: Nature of Defects, Tests Carried


Out Before Bonding, Post-bonding and In-service
Testing

Chapter 4. THE GENERAL PROPERTIES OF


POLYMERIC ADHESIVES . . . . . 143
Polymer Structures: Unsaturation. Mixed Adhesives.
Properties and Temperature: The Glass Transition
Temperature, Decomposition Temperature, Melting
Temperature, The Deformation of Adhesive Poly-
mers by Stress, Viscoelasticity, The Modulus of an
Adhesive, Poisson's Ratio, Strength Properties of
Adhesive Polymers, Yie1ding Stresses of Polymers,
Failure Modes After Yie1ding, Creep, Failure without
Yielding-Brittle Fracture, Crazing, Coefficient of
Thermal Expansion, Resistance to Deterioration

Chapter 5. FACTORS INFLUENCING THE CHOICE


OF ADHESIVE . . . . . . . . . . . . . . 175
Interaction with Substrate. Structural Adhesives for
Metals: Check-list for Structural Metal Adhesives
Used at Temperatures up to 70°C, The Advantage of
Supported Filmic Adhesives (Tapes'), Unsupported
Films, Liquids and Pastes, Infiuence of Metal of
Adherend, High Temperature Metal-Metal Adhe-
sion. Structural Adhesives for Wood: Check-list for
the Use of Structural Wood Adhesives. Structural
Adhesives for Mixed Constructions: Metal-Wood
Structures, Metal-reinforced Plastics Structures.
Choice of Adhesives for Semi-structural Use: Check-
list for Adhesives for Semi-structural Use

Chapter 6. SURFACE PREPARATION 218


Metals. Wood. Concrete. Glass or Carbon-fibre
Reinforced Plastics. Shot, Sand or Grit Blasting.
Solvent Degreasing or Wiping. Chemical Etching:
Aluminium, Ferrous Metals, Titanium, Other Metals.
Priming Layers: Primers as Coupling Agents

Chapter 7. SERVICE LIFE 237


The Creep of Adhesive Joints. Time-to-failure
(Under Static Loading). Cyc1es-to-failure: Influence
CONTENTS IX

of Temperature, Infl.uence of Test Frequency, In-


f1.uence of Amplitude, Infl.uence of Moisture. Effects
of Temperature Change on Joint Strength. Service
Life as Indicated by Climatic Exposure Trials

Chapter 8. APPLICATIONS 271


Aircraft, Anchorages. Bridges. Carriages. Cars.
Decking. Furniture. Glass Reinforced Plastics. Heli-
copters. He1icopter Blades. Hovercraft. Lamp Posts.
Magnets. PABST. Rollers. Segmental Construction.
Ski Constructions. Telephone Kiosks. Yachts

Reterences . 283

Appendix: Standard American and UK Specijications tor


Adhesion Tests. . . . . . . . . . . . . . 293

Author Index 299

Subject Index 303


Chapter 1

Introduction

An engineer designs or builds machines and structures. This book is


about the design and building of structures using a particular method
of construction chosen from the several, diverse methods available.
Structures built from metallic components can have these components
assembled by bolting, riveting or other forms of mechanical fastening,
by welding, by the related processes of brazing or soldering, or an
organic adhesive can be used to bond the components to each other.
One of the few remaining alternatives is to machine the structure from
a solid block as astatue is sculpted from a monolith. Training and
experience combine to teach the engineer to marry design and method
of assembly when considering the various me ans of mechanical fasten-
ing, distinguishing them not only from each other but also from the
alternatives of welding or brazing. His training is less able to cope with
the use of adhesives, with essential choices between the many types
available and with the design approach appropriate to structures
assembled with them. Subsequent chapters of this book have been
written to assist the training of engineers in this extension of their
expertise, to give practising engineers an aide-memoire of design
criteria with adhesive properties and as a bonus, to help the adhesives
technologist in assessing the needs of the engineering industry.
Apart from structural uses, adhesives are used for surfacing either by
the adhesives material itself providing the finished surface or by its use
for attaching ceramic tiles or brick slips. They are also used as sealants
for curtain walls and for sealing the seams of liquid containers from the
small tin can to the integral fuel tanks of large aircraft. Neither the
surfacing nor the sealing use of adhesives will feature, other than
trivially, in these pages. Also excluded is the use of adhesives sub-
1
2 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

stances as aggregate binders in 'resin concretes', which may be re-


garded as structural material in their own right and have been recently
discussed in the context of adhesion (Hewlett and Shaw, 1977). Glass
and carbon fibre reinforced plastics are important structural materials,
the strength of which depends on the performance of the polymer
matrix as an adhesive but, once again, it is the material rather than its
manufacture which is the concern of the structural engineer. Fastening
or bonding structural components made of GRP or CFRP is a focus of
attention and features in Chapters 3 and 5-8.
This brief catalogue of inclusions and exclusions may be briefly
summed up by stating that the engineer, in building structures, is
concerned about the transmission and accommodation of stresses. This
book discusses the behaviour of adhesives, and joints made with them,
when stresses of significant magnitude are imposed, and how such
joints are best made.

JOINT CONFIGURATIONS

Joint design is dependent upon the nature of the materials to be joined


as weIl as on the method of joining. The first cast iran bridge, at
Ironbridge, Salop, was made by casting massive half-arches and as-
sembling the structure with mortise and tenon joints, with fit-on parts
and tapered pegs. This mode of construction was the only one known
to the designer, a mode used for timber structures. The only conces-
sion to the new structural material was in the use of large scale, single
piece, curved members. Rolled bars of T - and L-sections did not
become available until after the Napoleonic wars and the first I-beams,
for light loading only, were made in Paris in 1847 (Hamilton, 1958).
Heavy I-beams only became available in 1860 after the invention of
mild steel. Fastening was originally by slotting a circular cylinder, cast
as an integral part of the beam, over a cylindrical column, cast with a
retaining lip, but fish plates and bolting were introduced at an early
stage.
The design of joints in wooden structures similarly underwent
changes through the centuries when structural iron became available.
Mortised joints, at best half the strength of the wooden tie-beams
transmitting a puH, were avoided in large structures. In the eighteenth
century iron work was introduced, and forged straps secured by gib
and cotter, enabled loads to be suspended via wooden tie bars (Hamil-
ton, 1958). Tie bars were also lengthened by various designs of scarfed
joints secured with fish plates and bolts. These impravements in joint
INTRODUCTION 3

design enabled much lighter timbers to be employed with advantages


to other aspects of the design.
A comment, tinged with admiration for the appearance and achieve-
ment, which is frequently heard in buildings or when standing by old
machinery such as beam engines used for pumping, is that the size of
the members employed is the result of 'gross over-engineering'. The
implication is that greater stresses could be borne by the same design.
This is rarely so, partly because of the materials properties but mainly
because of the joint design where members meet and stress is distri-
buted between them.
The load-bearing structural adhesive joint is relatively new-a half
century is the measure of its time span. Carpenters' glue, used for
many centuries, was an adjunct to prevent looseness in a mechanical
joint, be it mortise and tenon, slot and peg or trenail. In any case it
could only be used on furniture protected from extremes of tempera-
ture or humidity.
It will be realized from this historieal divagation that the jointing of
materials has been subjected to a long process of development with the
structural joint as the most recent innovation. It will also be ap-
preciated that there is a dose relation between the materials to be
joined, the nature of the stresses to be transmitted and the method of
jointing adopted.
It is convenient now to define an adhesive as a polymerie material
whieh, when applied to surfaces, can join them together and resist
separation. A structural adhesive is one used when the load required to
cause separation is substantial such that the adhesive provides for the
major strength and stiffness of the structure. The structural members
of the joint, which are joined together by the adhesive, are the
adherends, a word first used by de Bruyne (1939).
Several important features about adhesive joints are capable of
elucidation from Fig. 1. The joints it illustrates were prepared by R. A.
G. Knight, now retired but formerly Head of the Composite Wood
Section at the Forest Products Research Laboratory (now the Princes
Risborough Laboratory), to have the same safe working load of 15 kN.
The outer members were timbers 3~ x 1 in and the inner 3~ x 2 in. The
overlap lengths were calculated from empirical rules based on experi-
ence with each type of fastening. The following details apply:

A. Overlap 5 in, resorcinol-formaldehyde adhesive joint.


B. Overlap 9 in, four ~ in bolts.
e. Overlap 11 in, two 21 in split ring connectors and one 1in bolt.
4 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

FIG. 1. Joints of equivalent load capacity with various fastenings (see text)
(Crown copyright).

D. Overlap 12~ in, twenty-four 4 in 7 gauge nails driven in double


shear into drilled holes.
The adhesive used is fully resistant to weathering and should outlast
the wooden adherends. By contrast, the nails will certainly rust unless
protected by paintwork which is renewed every few years. The other
mechanical fastenings will rust but are unlikely to fail from this cause
though corrosion products could be deleterious for the timber. The
total weights of structures with such joints will be greater with the
mechanical fastenings both because of their weight and because of the
increased overlap. There are some dis advantages to the adhesive joint
but these all occur at the stage of manufacture. The major one is the
INTRODUCTION 5

need for holding the members during a setting or curing time, secon-
darily the skill and knowledge required, although trivial, is less wide-
spread than the simple mechanical skills of drilling, bolting, etc. The
weight increases and the overlap lengths show that the additional
stresses introduced by mechanical fastenings have a penalty which is
avoided by the use of an adhesive.

Lap-shear Joints
Figure 1 shows a configuration known as a double-Iap shear. A
single-Iap shear is the more usual form of joint. In this, two adherends
of equal thickness with a single interface replace the two equal
members which balance a central member of double thickness in the
double-Iap joint. Both configurations are used for test purposes rather
than featuring explicitly in actual constructions. They do, however,
simulate constructional features and the extent to which this is so will
be apparent in the subsequent chapters of this book. They are impor-
tant because the best use of an organic adhesive is to resist a shearing
stress, simple or torsional. Designing with adhesives, as is indicated in
Chapter 2, consists ina large part in arranging that stress is resisted in
shear rather than by direct tension or cleavage. With all adhesives, a
far larger stress can be borne in shear than in tension.

~ _______~______~____~______~3
FIG.2. End-to-end jointing converted by straps into shear.

The simplest example of the conversion of direct tension to lap shear


is provided by the replacement of end-to-end jointing by the strapped
configuration of Fig. 2, a method which is, of course, exactly like that
used in constructions involving bolting or riveting but unlike that used
in welding. For the strapped joint, the lap shear is a good testing
simulation. Similarly, the bonding of stringers on to the skins of
aircraft bodies is a use for which the single-Iap joint serves as a good
model though the stiffness of the assembly cannot be deduced directly
from the lap shear modulus nor its buckling force from the lap shear
strength.

Butt Joints
The butt joint, in which the two components are pulled apart by forces
acting along the axis of the joint, is again used as a test procedure,
6 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

FIG.3. Filleting in the construction of honeycomb.

usually with cylindrical though sometimes with square adherends. As


already emphasized, it is a configuration best avoided in designing
structures but, as a test procedure, is useful in assessing the perfor-
mance of joints subjected to cleavage. The butt joint may be subjected
to torsional shearing stress instead of direct tension and is then a
reasonable type of joint if only torsional loads are applied, the
maximum stress being determined by the diameter of the cylinders.
Non-axial loading of a butt joint leads directly to cleavage stress;
this differs from direct tension only in the concentration of stress at
one side of the specimen. All joint configurations can be regarded as
elaborations of these two types of joint and they are dealt with in
detail later in this book.

Fillets
Also considered in greater detail is the important part played by fillets
of adhesive at the free edges of joints. In particular, honeycomb
panelling as a structural material would not function properly if the
adhesive used did not, on melting and wetting the metals, creep up the
honeycomb to form a fillet with the base as shown in Fig. 3. The joint
thus formed has a much more favourable stress distribution than it
would if only the edge were bonded to the base plates. The rheological
property of the liquid adhesive whereby it both forms and maintains a
fillet during the curing operation is not shown by all materials and the
polymer used as the adhesive may require additives to induce this
behaviour.

METALS AND OmER CONSTRUCTIONAL


MATERIALS
The discussion so far has assumed either indifference to the material of
construction or that the material is mild steel or aluminium. All metals
INTRODUCTION 7

can be fastened by adhesives but surface preparation is, with very


limited exceptions, always necessary and for some metals the prepara-
tion demands are more exacting than for others. Such matters are
discussed in detail in Chapter 6.
Wood has traditionally been joined with carpenters' glue but it is not
widely realized that the function of the glue is essentiaHy gap-filling
rather than load-bearing. Traditional joinery with wood always im-
poses the principal stresses at wood-to-wood surfaces, the glue be-
tween them being in compression. The stresses imposed on the glue-
line were such as to impart stifIness and rigidity to the structure. A
traditionally joined wooden chair would still function as achair in the
absence of glue but it would lack rigidity to the point of looseness of
the members and could easily be dismembered. Similar functionality
but with a lack of rigidity would occur in a chest of drawers with the
drawer sides dovetailed but not glued to the front. Those engineering
uses of timber outside the field of domestic furniture have involved,
and increasingly so, the use of timber of length greater than is
obtainable in the single piece. Jointing to withstand tension and shear
thus becomes necessary. The earliest way of spreading the stress and of
bearing a major part of it in shear was the use of the scarf joint. The
multiple scarf or finger joint is the most efficient way provided that the
fingers are accurately cut by machinery designed for the purpose.
Glass and carbon fibre reinforced laminates (GFRP and CFRP) are
normaHy fabricated to the shape desired without the need for joining
separate units to form the finished product. This situation arises from
the use of these relatively expensive materials in special products, from
the comparative ease of making them in complicated shapes and
sections as weH as from the absence in commerce of standard construc-
tional units corresponding to I-beams and angle iron or the extruded
sections of the softer metals. The further development of the pultrusion
technique could weH remedy the lack of the last item of this list if a
demand were seen for some given sections of desirable general utility.
Occasion does arise when GFRP or CFRP needs to be joined to metal
and for this purpose adhesives are commonly used because the non-
metallic component can be tailored accurately to face to the metal
surface with shaping to allow correct contouring of the stress. The use
of mechanical fastening through the thickness of the material would
reintroduce the high stress concentrations which the contouring is
designed to eliminate. The transfer of stress from metal to GFRP or
CFRP panels exemplifies the general property that joints between
8 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

unlike materials are conveniently made with adhesives, particularly


where contact potentials may give rise to corrosion currents or the
materials themselves deteriorate in contact, e.g. oak and mild steel or
CFRP and some met als which corrode if bolted or riveted. The
adhesive helps to provide electrical insulation.

THE DECISION TO USE ADHESIVE BONDING

There are a number of general features of design which may determine


whether or not adhesive bonding is a suitable construction technique:
(i) Does the structure need to be dismantled tor repair, maintenance
or inspection?
An adhesively bonded structure will most probably not be capable of
being dismantled and re-assembled, at least without considerable
cleaning and repetition of surface preparation. Since this can imply
removal of material, parts may no longer be to tolerance or even fit.
Sub-assemblies manufactured with adhesives instead of bolting and
threading may be economically designed as throw-away parts not
needing ever to be dismantled.
(ii) What is to determine the service life of the structure?
In many cases, this will be mechanical failure or near-failure. If the
adhesively bonded joint is continuously stressed in shear it may be
expected to creep (see Chapter 7) and failure may be judged not when
the creep ends in stress rupture but when the strain reaches a certain
figure. This may be determined by interference with other components
or may simply be laid down as an arbitrary safety limit. Joints loaded
in shear at a reasonable stress and with comparatively small cleavage
components to the stress should not fail catastrophically without some
creep being apparent though this will depend on the ductility of the
particular adhesive and on the temperature. Sudden catastrophic fail-
ure either on impact or through fatigue is usually associated with high
cleaVage stress. If stress rupture after repeated applications of stress is
the expected failure mode, then the rate of application of stress
becomes more important with adhesive bonding than with mechanical
fastening.
Stress rupture after some stated number of cycles of loading is a
frequent failure criterion used in the aircraft industry. Figure 4, repro-
duced from Argyris (1962), shows typical fatigue curves for riveted
INTRODUCTION 9

...g' 0·6
.c

"
<-
t: 0.5
!
"'
E 0.4
'Z
:;

.
'" 0·3
'<-"
1;; 0.2
l(

"'
1: 0.1
Riwted

Numbllr of CYClflS

FIG. 4. Fatigue strength of riveted versus adhesively bonded joints (after


Argyris, 1962).

and adhesively bonded aluminium plates, indicating the superiority of


the latter. Choice of adhesive and care in surface preparation are
essential prerequisites for this superior performance.
Where vibration is transmitted through joints, mechanically fastened
components can sometimes show fretting due to very small amplitude
relative movement normal to their mating surfaces: whilst adhesive
bonding not only eliminates fretting it can also serve to damp or
attenuate the vibration. Vibration can also loosen bolted assemblies
and thread-Iocking adhesives are used to prevent this.
(iii) How is assembly to be effected?
In most mechanical methods of assembly the faying surfaces are laid
together and can be moved relative to each other until alignment of
drilled holes enables fastening to be effected. Holding the parts in a jig
may be required if drilling and bolting or riveting are done together
but immediately the joint is made it is fully load bearing and the jig
can be removed. Welding may also require jigging but the joint
becomes self-supporting during the time necessary to remove the jig.
This comparative ease of assembly which characterizes mechanical
fastening, welding and brazing is absent from adhesively bonded
assemblies. In this context, it is convenient to divide adhesives into
three c1asses:
Contact adhesives achieve an instant bond on impacting the two
10 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

surfaces. This precludes movement of the parts with respect to each


other from the moment of contact. Jigging is frequently necessary to
ensure correct placement of the parts in a single decisive movement.
The strength of the bond does however take time to develop and
although component members may be self supporting after 10-30 min,
full bond strength may take 24 h to develop. Similar considerations of
placement apply to hot melt adhesives although they become load-
bearing much more quickly. They are, however, little used in structural
engineering because of their tendency to 'cold flow' and their sensitiv-
ity to temperature at both ends of a comparatively small range.
Hot setting adhesives achieve high structural strength by forming a
cross-linked, three-dimensional, molecular network by chemical reac-
tion. Normally this is accomplished at elevated temperatures, but in
many cases the reaction is to proceed at room temperature though, of
course, taking a much longer time to reach completion. During this
curing reaction, the adhesive becomes, or is already, liquid and able
easily to shear. The faying surface must therefore be held until the
re action has proceeded sufficiently for the parts to be self supporting.
SL~h curing is frequently accomplished between the plattens of a
heated hydraulic press. Indeed, for some adhesives, hydraulic pressure
is an additional requirement since water vapour or other material is
evolved during curing and must be retained in solution. The period of
occupancy of the press can sometimes be shortened by removing the
components when it is safe to handle them and completing the chemi-
cal cross-linking in an oven either still contained in a jig or merely
supported in the oven atmosphere. The operation is frequently called
post-curing.
Cross-linking adhesives which can cure at room temperature may
take 2-3 days to achieve 50% of the ultimate strength. Moderate heat
has, in these cases, a good accelerating influence and an oven at, say
60°C may enable a good cure to be reached within a day.
Anaerobic and cyanoacrylate adhesives are both monomeric materi-
als which polymerize rapidly within the glue-line once the latter is
formed. Setting can be very rapid indeed with cyanoacrylates and this
implies that rapid and correct location of the adherends is necessary.
Anaerobics also set more rapidly but movement of the surfaces is
possible after polymerization has commenced. However, such move-
ment is inimicable to the highest bond strengths but, because of its
possibility, jigging is desirable. In the most common use of such
adhesives, namely thread locking, the application is, in effect, its own
jig.
INTRODUCTION 11

Both types of adhesive yield bonds which, depending on formula-


tion, achieve load-bearing strength quite quickly but continue to
increase in strength over several days.
It will be seen that all classes of adhesives, therefore, involve the
necessity for fastening by jigs if there is any possibility of movement
during placement or curing. With some adhesives it is a question of
instant and exact placement, with others maintaining immobility during
aperiod of curing which may be quite lengthy and lastly, with certain
specialized applications, placement being achieved within the second
but held for many minutes. In all these situations the ultimate bearing
load is only achieved with time. In these particulars, advantage is
wholly with mechanical means of fastening and the method of assem-
bly is therefore an important aspect of design if adhesive bonding is
considered feasible or desirable on other grounds.
(iv) In what environment is the assembly to operate?
This criterion for design is one which can operate against the choice of
an adhesive for joining components of an assembly. A hot, mo ist
environment will cause the corrosion of mild steel, some aluminium
alloys and die metal castings. The effect of this corrosion may be gravely
enhanced by electrochemical effects arising from the juncture of differ-
ent metals or alloys or the same met als partly enveloped in oil and
partly exposed to the atmosphere. This corrosion will normally be
apparent on simple inspection and its severity fairly easily assessed.
Many adhesives are sensitive to similar environments but the effect of
the imbibed moisture, being located at the interface of the bond, is not
apparent until failure has occurred. It has been realized in recent years
(see, for example, Cotter, 1977) that the combined effect of moisture
and sustained stress is particularly damaging for some structural adhe-
sives, even where the joint design is wholly favourable to the use of
adhesives. The need to consider the operating environment is almost
entirely a question of economics. It may involve designing the joint at
a lower stress level but, more usually, it will involve a more expensive
surface preparation with possibly the use of a primer (see Chapter 6)
and the use of a more expensive adhesive undoubtedly involving the
curing of a cross-linking adhesive by heat. In spite of this pessimistic
note, it must be added that suitable structural adhesives are known to
perform satisfactorily under the surface of the sea and, an even more
demanding situation, in the zone of interaction of sea and air. In this
zone wave-induced stresses exist to provide an added cyclic component
to the steady imposed load whilst oxygen and sunlight are added to the
chemical stresses.
12 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TUE BALANCE OF ADVANTAGES AND


DISADVANTAGES

If the assumption is made that it is always possible to design the


assembly so that adhesive joints take their main stress in shear, where
they perform best, and mechanical fasteners such as bolts take their
main stress in tension, where they show to best advantage, it is possible
to compare the advantage and disadvantage of adhesion bonding
without dealing explicitly with those of mechanical or other methods of
fastening. The implied comparisons are then always those of the best
joint configuration conceivable for each type of fastening and not a com-
parison of identical joints.
Feature Comment
Applied The stress will be more uniformly distributed and
stress stress concentrations at a lower level than with mechani-
cal fastening, though the maximum local stress which can
be applied will be lower. In a situation where the loaded
surface cannot be increased, adhesive bonding may be
inappropriate.
Service life Superior life under conditions of cyclic stress would be
expected in comparison with, for example, riveted
joints. Temperature limitations exist and, with metal
adherends, these temperatures will be below the upper
limits for the safe use of the meta!. Some adhesives are
sensitive to high humidity, particularly when continu-
ously stressed.
Preparation The machining of parts for assembly, as by drilling and
of tapping holes or using riveting or welding equipment,
adherends are all avoided. In their place, hydraulic presses, jigs and
ovens or autocIaves are variously required depending on
the size of the components and the adhesive to be used.
In large scale production, dispensingjapplication equip-
ment will be necessary.
Surface preparation will, in general, be necessary and
although mechanical abrasion or solvent degreasing
sometimes suffices it will often involve chemical etches
followed by primers.
Economics There is no alternative but to undertake a cost analysis
of the alternative designs. In general, designs involving
INTRODUCTION 13

adhesive bonding tend, because of the more uniform


stress distribution and the absence of the shear weight of
the mechanical fasteners, to be lighter and more
economical of materials. Structural adhesives are expen-
sive but so are nuts and bolts, threads, etc. Surface
preparation costs may be greater or lesser than machin-
ing costs in given circumstances.
Chapter 2

The Nature and Magnitude of Stresses


in Adhesive Joints

INTRODUCTION

Any rational basis of structural engineering design must be based on


the ability to predict the loads and hence stresses which are likely to be
encountered in practice. The loading system is often prescribed by the
function, but the skill of an engineer is to use the best available
materials and design techniques to arrive at a suitable and cost-
effective solution. Centuries of practice can lead to many design skills
becoming second-nature to the experienced engineer. However, ad-
vances in technology together with new and more demanding environ-
ments have resulted in a strong emphasis being placed in modern
engineering on the need to quantify the structural loads and stresses.
Pressures come not only from market-oriented competition which
requires a cost-effective product, but also from legislation which may
seek to blame in the courts of law the designer who was careless and
allowed his structure to fai!. Thus it is with adhesive joints, as with any
fastening method.
Figure 5 shows some typical classifications of joint which are com-
monly found in current engineering practice. Any joint occurring in
practice is designed to carry a given set of loads. Most of the adherends
(the technical term for the materials to be joined) are loaded in
tension. The subsequent loads on the adhesive are then a function of
the geometry of the joint.
In Fig. 5(a) is shown the single-Iap joint. This is one of the most
commonly occurring joints and is the configuration most often used for
testing adhesives. However, the stress state is complex. As shown, the
loads in the single lap are not colinear. A bending moment must
14
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 15

~- _ _ _ _ _---,I -

c:::::==~~,.-_---,l-

_ _ _ _ _--'1-

.-------=------~)--

~----~=-------~I--

-+1-=1-;:-:: 23--~-~-tT I
il Tubular lap }
~

jl Peel /

FIG.5. Some common engineering adhesive joints.

therefore exist and the joint will rotate as shown in Fig. 6(b). Clearly,
the adhesive layer will no longer be solely in shear, but will have
tearing stresses at the ends of the joint. Further, the adherends are no
longer in simple tension, but are bent. It is quite possible that both the
adhesive and the adherend may have become plastic, particularly in
the highly-stressed regions. Indeed, the commonly used high-strength
16 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

P - 1_ . _ Lc------ Bending moment Mo =PÖ /2


1-·- ·_·+-P
(al Undeformed joint

M =KPÖ
P-f . _ _ _ ~/,B,"d;" ~='; -2-
K<1

(bl Deformed joint

FIG. 6. lllustrating a way of representing the Goland and Reissner bending


moment factor geometrically.

aluminium alloy adherends used in the single-Iap joint test are often
found to have gross plastic deformation, owing to rotation, after
failure.
It used to be said that by using a double-lap joint, as shown in Figs 1
or 5(b), the bending effect could be eliminated. Whilst it is true that
there is no gross rotation of the joint, it is no more than a back-to-back
arrangement of two single-Iap joints, each of which has its own system
non-colinearity in the region of the joint! (The exaggerated deftections
shown in Fig. 11 (p. 27) nonetheless illustrate what happens. There is a
tensile tearing stress where the thick adherend is loaded, and a com-
pressive stress at the other end.)
The other configurations of joints shown in Fig. 5 are largely
designed to improve load-transfer so as to minimize stress concentra-
tions and peel. Peel is the hated enemy of the joint designer. Whereas
the adhesive in the single- or double-Iap joint is, certainly at low loads,
largely in shear, that in Fig. 5(j) is experiencing principally transverse,
tearing loads which we call peel. Indeed, the configuration shown in
Fig. 5(j) is that used as the basis of a number of so-called peel tests.
But let us even at this early stage take breath and look back. Surely,
with a few simplifications and assumptions we can tackle these joints,
especially if we forget about rotation and keep everything elastic. So
we can, but are the answers realistic? The purpose of this chapter is
therefore to look at the various methods available for analysing joints
(and some of the well-known results) and to indicate to a designer the
best way forward. In the end, we need to be able to predict the strength
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 17

of a joint when it is subjected to a given set of loads. It is therefore


necessary to understand those factors which bear heavily on the
answer. Basically, these factors are the loading system, the magnitude
of the loads, the structural geometry, and the mechanical properties of
the adhesive and the adherends.
However, let us be under no illusions. It is no easier to predict the
strength of a joint from first principles than it is to predict the fatigue
life of a steel bar by knowing its crystal structure and the strength of
the crystals. Any theoretical predictions must be justified by experi-
ment. There has to be a limit, or rather a collection of limits, set for a
given situation. For instance, for a joint which is often used, testing
and experience may be better than any analysis. However, an expen-
sive, one-off design in a critical component would repay a careful
analysis before metal is cut.

Reality
The purpose of this section is to outline some of the complexities of
the problem. To start with, the joint does not consist of simple,
separate, elastic materials with a clear, mathematical geometry. The
surface of the adherend is rough and usually has an oxide or similar
layer. The thickness and elastic modulus of this layer are often indeter-
minable. In addition, a primer is sometimes applied to the surface of
the adherend for reasons of chemical compatibility, protection during
storage, handling, and so on. The thickness and elastic modulus of the
primer are again variables to be allowed for before the adhesive proper
is reached. If the primer is similar to the adhesive, then a uniform state
is reached more quickly.
But what of the adhesive? Are its properties the same in the thin
film form (0·1 mm or so thick) as they are in bulk? For many years, it
was believed that the elastic moduli increased with reducing glue-line
thickness. Unfortunately, this was based on an inadequate interpreta-
tion of a careful experiment! Recent work has shown that the film and

TABLE 1
TYPICAL ADHEREND AND ADHESIVE PARAMETERS
THROUGH AN INTERFACE

Bulk adherend ;;;:0·5rnrn


Surface oxide 4-300 x 10-6 rnrn
Primer 0·02-0·1 mm
Bulk adhesive layer 0·1-0·5mm
18 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

bulk properties are similar for the same eure condition (Adams and
Coppendale, 1977).
In Table 1 a typical set of adherend and adhesive parameters are
given. The values given are ranges from various sourees.

Methods of Mathematical Analysis


It has been pointed out above that it is necessary to carry out some
form of mathematical analysis so that strength predictions may be
made or, at least, can be extrapolated from experimental data. A
variety of possibilities exists.
The method most attractive to a mathematician is to set up aseries
of differential equations to describe the state of stress and strain in a
joint. Then, by using stress functions or other methods, closed-form
algebraic solutions may be obtained. In the simple case where the
adhesive and the adherends are all elastic, it may be possible to devise
a solution for given boundary conditions. These methods become
increasingly more difficult to use as non-linearities arise. Causes of
non-linearity vary from joint rotation to material plasticity. However,
by making various simplifications, it is possible to produce solutions
which can readily be used to vary geometrie and material parameters
to cover a range of cases. Thus, once the solution has been obtained, a
parametric study may be carried out with great efficiency, provided the
limits of the simplifications made are borne in mind.
Modern digital computers have led to the use of a variety of
numerical techniques for solving problems of mechanics, fluids, ther-
modynamics and so on. In one of these methods applied to adhesive
joints, the structure is split into aseries of small parts (called finite
elements), each of which obeys the prescribed material behaviour and
each of which interacts regularly with its neighbours in terms of force
continuity and displacement compatibility. The finite-element techni-
que (FET) is very powerful. It can be used to accommodate awkward
shapes plus material and geometrie non-linearities, which can throw a
mathematician into a panic! It can be efficiently used in that small
elements need only be used where there are large stress gradients.
However, the big problem is that each solution applies only to a given
set of parameters. If it is necessary to vary joint length, adhesive
thickness, adhesive modulus, and so on, a new computer run is required
each time. While it is possible to use mesh generation programs to
alter the geometry, and while changing moduli is very simple, the cost
in computing time for a parametrie study is not insignificant. Because
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 19

very large stress gradients, approaching singularities, can exist in


typical joints, it is necessary to use very fine meshes. Thus, the
computing power required is large and the cost of a parametric study
unattractive, particularly when non-linear behaviour is included. How-
ever, for investigations into the mechanics of real joints, there is no
real substitute for finite-element techniques. Also, where one-off joints
are concerned, it is probably much more efficient than trying to evolve
a closed-form solution-at least you know that you can get an answer
in the end, which is not true of the purely algebraic method.
It is now opportune to illustrate some of the points made above by
considering examples of simple joints and how these can be treated.
There is no point in trying to re-invent the wheel. We will therefore
look critically at the existing literature and theories on adhesive joints
rather than present an apparently all-pervading solution without dis-
cussing alternatives. First we will study the simple lap joint with
parallel adherends. All parameters will be considered elastic, but not
necessarily linear: we will then allow for some unexpected effects at
the ends of the bond-line and, finally, consider the case where the
adhesive and/or the adherends may become plastic. Modifications of
the simple lap joint, together with butt, tube, peel and other geomet-
ries, will also be considered.

THE SINGLE LAP JOINT

Linear Elastie Analysis


The lap joint, in which two sheets are joined together with an overlay,
is one of the most common joints encountered in practice. Joints of
this type made from thin aluminium sheet 25 mm wide, 12 mm long
(1 in by 0·5 in) and 1· 6 mm thick have long been used for quality
control. The joint is easy to make and the results are sensitive to both
adhesive quality and adherend surface preparation.
The simplest analysis considers the adherends to be rigid and the
adhesive to deform only in shear. This is shown in Fig. 7(a). If the
width of the joint is b, the length 1, and the load P, then the shear stress
'T is given by:

and the adherend tensile stress will decrease linearly to zero over the
20 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

P -1,-- ------\;r~::,~i
:
\"'.\\s.\
B:

~'~l~'~'~i________~'--p
-1-

(al

A A B

(bl

FIG. 7. Exaggerated deformations in loaded single-lap joint: (a) with rigid


adherends; (h) with elastic adherends.

joint length from A to B. In Fig. 7(b) is shown a similar joint but in


which the adherends are now elastic. For the upper adherend, the
tensile stress is a maximum at A and falls to zero at B. Thus, the
tensile strain at A is larger than that at Band this strain must
progressively reduce over the length 1. The converse is true for the
lower adherend. Thus, assuming continuity of the adhesive/adherend
interface, the uniformly sheared parallelograms of adhesive shown in
Fig. 7(a) become distorted to the shapes given in Fig. 7(b). This
phenomenon is called differential shear. Essentially, this is what Vol-
kersen (1938) analysed. It is instructive to study Volkersen's analysis
first and then to add to it the subsequent, more refined analyses, and to
see what effect they have.

Volkersen's Analysis
In Volkersen's shear lag analysis it is assumed that the adhesive
deforms only in shear, while the adherend deforms only in tension, as
illustrated in Fig. 8. Consider the relative displacements of the upper
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 21

-
p

11

(al

(bI

FIG. 8. Single-Iap joint analysed by Volkersen (1938): (a) undeformed


geometry; (b) section through deformed joint showing assumed forces.

and lower adherends 8x along the overlap. Then:

8x =8 0 - r Bldx+ L/2
L/2
x
r B2 dx x
(1)

where 80 is the relative displacement for x = -1/2 and Bl and B2 are the
distributions of direct strains in adherends 1 and 2.
For a unit width and applied load P for the upper adherend we
have:

Bl=E1
h
r TxdX]
[p_ 1l/2 x
(2)

and for the lower:

1
B2=-E IX Tx dx (3)
t2 -1/2
22 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

where E = the adherend Young's modulus. Für the adhesive we have:

(4)

where G = the adhesive shear modulus.


By substituting eqns (2), (3) and (4) into eqn (1) and differentiating
twice, we obtain an equation of the form:

which has a solution:


Tx = Al cosh wx + A 2 sinh wx
where Al and A 2 are arbitrary constants, defined by the boundary
conditions. Proceeding in non-dimensional terms, i.e.:

where Tm is the averaged applied shear. With appropriate boundary


conditions we obtain the adhesive shear stress distribution:
_ w cosh wX (t/J - w sinh wX1)
T ="2 sinh w/2 + t/J + 1 "2 cosh w/2
where
w2 = (1 + t/J)</>
t/J = t l /t2
Gl 2
</>=-
Et l t3
X=x/l
and

For the case of adherends of equal thickness, t l = t2 = t. Then t/J = 1 and


w=.J(2</».
The maximum adhesive shear stress occurs at the ends of the joint:

i max = ~~ coth ~~
It is worth noting that für large overlap lengths:
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 23

and the absolute value of the maximum shear stress, T mrno is propor-
tional to:

which is independent of overlap length.


The theory developed by Volkersen takes no account of two impor-
tant factors. First, the directions of the two forces P in Figs 7 and 8(a)
are not colinear. There will therefore be a bending moment applied to
the joint in addition to the in-plane tension. Secondly the adherends
bend, allowing the joint to rotate as shown in Fig. 6(b).
The rotation alters the direction of the load line in the region of the
overlap, giving rise to a geometrically non-linear problem, since the
joint displacements are no longer proportional to the applied load.

The Analysis 01 Goland and Reissner


Goland and Reissner (1944) took this effect into account by using a
bending moment factor, k, which relates the bending moment on the
adherend at the end of the overlap, Mo, to the in-plane loading, by the
relationship :

where P is the applied load and t is the adherend thickness (the


glue-line thickness was neglected). If the load on the joint is very
small, no rotation of the overlap takes place, and the line of action of
the load is as shown in Fig. 6(a), passing elose to the edge of the
adherends at the ends of the overlap. In this case, therefore, Mo = Pt/2
and k = 1·0. As the load is increased, the overlap rotates bringing the
line of action of the load eloser to the centre-line of the adherends, as
shown in Fig. 6(b) and thus reducing the value of the bending moment
factor.
The second theoretical approximation of Goland and Reissner is
applicable to metal-to-metal adhesive bonded joints, whereas the first
approximation assumes the joint to the monolithic. By treating the
adhesive layer as an infinite number of shear springs and an infinite
number of tensionjcompression springs in the y direction (a description
first explicitly used by Cornell, 1953), soluble differential equations,
assuming plane strain, to describe both the shear stress (T3J and the
normal stress (U3y) distributions can be derived.
24 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Limitations of the Second Approximation of Goland and Reissner


The range of validity of the second approximation is given by Goland
and Reissner to be:

where E, G and t are the Young's modulus, shear modulus and


thickness and the subscripts 1 and 3 refer to the adherends and
adhesive respectively.
The upper bound is governed by strain energy considerations in that
the theory neglects the work of the stresses O"ly and Tlx in the
adherends compared with the work of the stresses 0"3y and T3x in the
adhesive layer. A derivation of these bounds is given by Sneddon
(1961).
For a typical lap joint, t 1 = t2 = 1·62 mm, t3 = 0·25 mm, GI = G 2 =
25·9 GN.m- 2 , EI = E 2 = 69 GN.m- 2 , G 3 = 1·72 GN.m- 2 , E 3 = 4·82
GN.m- 2 , so:

t1 G 3 = 0.43· t1E 3 = 0.45


t3 G 1 ' t3 E 1
and the Goland and Reissner theory is not strict1y applicable. How-
ever, Lubkin and Reissner (1956) considered that the results of photo-
elasticity experiments presented by Cornell (1953) show that the
bounds of the theory are conservative, but they have given no indica-
tion of the true bounds.
These bounds will apply to any theory which makes similar assump-
tions to the Goland and Reissner theory in neglecting the shear
deformations and tensile stresses across the adherends. They do not,
however, apply to the three-dimensional theory presented later (due to
Adams and Peppiatt, 1973), because the shear stress in the adherends
is assumed to vary linearly across the thickness. This theory neglects
the existence of tensile stresses within the adhesive layer.
Benson (1969) has indicated more fundamental limitations in the
Goland and Reissner theory in that the derivation of the bending
moment factor, k, allows for the rotation of the overlap region, but the
internal stresses in the joint are derived assuming no rotation. The
solution for the adhesive shear and the tensile stresses are only strict1y
accurate for k = 1·0, i.e. for very low loads. It should be pointed out
that the derivations relating the bending moment factor to the load
assume that the joint is monolithic, Le. that the glue-line thickness is
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 25

zero, whereas the second approximation applies to joints with a finite


glue-line thickness.

Results from the Second Approximation of Goland and Reissner


Kutscha and Hofer (1969) indicated that the solution for the normal
stress, 0"3y, was incorrect1y written in the original paper. The solution
has been checked, and the correct result, as is also given by Sneddon
(1961), is:

O"t 2
0"3y = c 2 R 3 R 2 A
[( 2 k,
2- ),\x,\x
Ak cosh A cos A cosh C cos C

+ (R1A 2 ~- Ak' sinh A sin A) sinh~ sin ~]


where
c= 1/2
A =~(6E3t)!t
t Et3

k'=k~(3(1-v2) ;y
R 1 = cosh A sin A+sinh A cos A
R 2 = sinh A cos A- cosh A sin A
R 3 = (sinh 2A +sin 211.)/2
v = Poisson's ratio

Aseries of parametrie studies of the Goland and Reissner solutions


was carried out by Kutscha and Hofer (1969) and they noted that both
the load on the joint and its width are not explicitly factorable from the
functions for shear and normal stress (i.e. the applied adherend stress
is not factorable). This is because of the change in value of the bending
moment factor as the load is increased.
The influence of the applied load on a 25·4 mm (1 in) wide lap joint
is shown in Fig. 9. The maximum tensile stresses in the adhesive per
unit load of both the original and corrected solutions, and the max-
imum adhesive shear stress per unit load are shown. The maximum
stresses occur at the ends of the overlap. It can be seen that the
maximum stress per unit load in the joint decreases as the load is
increased. The greatest decrease occurs in the value of the corrected
26 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

2i---===:::::::-~d~3~Y(""II,not)

..
'tI
o _
-"
2 'E 10
E
ii:b
:J

~
"<- ~
Vi 5 Typlcal 5trongth rang~
for standard 127mm
long JOInts

0'L-~1~0--------~1~OO~------~1~OOO~--~~~~~-~~~
Appl,ad load (N)

FIG. 9. Maximum shear and normal stresses, T3x, 0"3y, plotted against applied
load according to the second approximation of Goland and Reissner.

~
E
z bO
ror OrIgInal
$olul,on
~

..
'0

~ 50
/
z
>< Corrected
...'" / $oIUIIOn
L 40 -
2
i
:>
s1"' I(
:;
"
.."
;0
40~~
L...
~.. 30
e
L

...
Perctntaqe rrror
L
-;Ji

..
C
!!l
10 '"c:
!!
....~ '0
..~
c
t:- 00
i
2
,
3
t
/)
0..

Young's modutus 01 adhll2'sivQ (GN/m Z )

FrG. 10. Original and corrected solutions of Goland and Reissner for the
maximum transverse tensile stresses in the adhesive layer of a single-Iap joint
plotted against Young's modulus of the adhesive (from Adams and Peppiatt,
1974).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 27

maximum tensile stress per unit load. The photoelastic work of McLaren
and MacInnes (1958) confirms qualitatively that the maximum
stress per unit load in a single lap joint reduces as the load on the joint
is increased.
Figure 9 also shows that the original solution for U3y given iu the
paper by Goland and Reissner is considerably in error at higher joint
loads. Figure 10 shows the two solutions for the maximum tensile
stress for a joint load of 4· 5 kN, together with the percentage error of
the originally published solution, for a range of adhesive Young's
moduli. It can be seen that the original prediction is considerably in
error, especially for low modulus adhesives.

Etlect of Bending in 31 Donble-Iap Joint


Although there is no net bending moment on a symmetrical double-lap
joint, as there is with a single-lap joint, because the load is applied
through the adhesive to the adherend plates away from their neutral
axes, the double-lap joint experiences internal bending, as shown
diagrammatically in Fig. 11 (see also Fig. 6). In a symmetrical double-
lap, the centre adherend experiences no net bending moment, but the
outer adherends bend, giving rise to tensile stresses across the adhesive
layer at the end of the overlap where they are not loaded, and
compressive stresses at the end where they are loaded as shown in Fig.
11.

Double lap. loaded almost to failure


Adhesive transverse
tension normal stress

t-·_·_·-
2P-'·_·_·;t==t _3--P
1._ ..::J---P

FIG. 11. Bending moments induced in the outer adherends of a double-lap


joint.
28 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Volkersen's Second Theory


Treating the adhesive in the same way as Goland and Reissner,
Volkersen (1965) has set up soluble differential equations to describe
the tensile stresses (a3y) and the shear stresses ('T3x) in the adhesive
layer of a double-Iap joint. Volkersen's second theoretical model for
the double-Iap joint neglects the same adherend stresses as does the
second theory of Goland and Reissner, so it has the same bounds of
validity. The bending in the double-Iap joint does not cause rotations
of the overlap region, and so the adhesive stress per unit load is not
dependent on the load applied. Thus, the applied load is explicitly
factorable from the solution functions for the shear and normal stresses
in the adhesive layer.

Later Work
The classical early work of Volkersen, Goland and Reissner and other
earlier workers was limited because the peel and shear stresses were
assumed constant across the adhesive thickness, the shear was a
maximum (and not zero) at the overlap end and the shear deformation
of the adherends was neglected. Because the end face of the adhesive
is a free surface, there can be no shear stress on it. Thus, by the law of
complementary shears, the T x shear stress at the joint end must also be
zero. Volkersen's later work (1965) did include an allowance for end
effects (by putting 'T3x = 0 at the end of the bond-line) and adherend
bending, but Peppiatt (1974) found that there were errors in the paper
and it was impossible to derive the solutions. By neglecting the
adherend stresses caused by bending, Peppiatt showed that it was
possible to derive a solution for the adhesive shear stress and the
transverse (peei) stress. A similar solution was also derived by Benson
(1969).
Recently, several authors, notably Renton and Vinson (1975) and
Allman (1977) have produced analyses where the adherends have been
modelled to account for bending, shear and normal stresses. They have
also set the adhesive shear stress ('T3x) to zero at the overlap ends. In
addition, Allman has allowed for a linear variation of the peel stress
across the adhesive thickness, although his adhesive shear stress is
constant through the thickness.
Let us now look at another aspect of stresses in joints which is often
neglected. It has been shown by Adams and Peppiatt (1973) that there
exist significant stresses across the width of an adhesive joint. By using
reflective photoelastic analysis of a lap joint, Hahn (1960) showed that
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 29

the shear stresses in the adhesive were highest at the corners. In this
experiment, the adherends were allowed to bend, and the high
stresses at the corner were thought to be caused by the anticlastic
bending of the adherends. Basing his ideas on this supposition,
Kutscha (1964) drew a qualitative picture of the distribution of longi-
tudinal shear stress in the adhesive of a single-lap joint.
Adams and Peppiatt (1973) considered the existence of shear stres-
ses in the adhesive layer and direct stresses in the adherends acting at
right-angles to the direction of the applied load, these stresses being
caused by Poisson's ratio strains in the adherends. A simple physical
argument may be used to indicate the existence of these transverse
stresses. If the adherends are in uniform tension up to the joint, there
should be a uniform contraction both of thickness and of width
because of Poisson's ratio. However, there is no load in the other
adherend at this end section and so there can be no lateral contraction
(see Fig. 12). If the adhesive had zero shear stiffness, this situation
could exist but, since this is hardly likely, the load-carrying adherend is
restrained from contracting fully and lateral tensile stresses are de-
veloped. Similarly, the other adherend is subjected to lateral compres-
sive stresses. For a symmetrical joint there will be no lateral stresses
half-way along the joint.
Adams and Peppiatt neglected the effects of bending (so that their
results are more applicable to double- rather than single-lap joints)
and treated the adhesive as an infinite number of shear springs. Thus,
tearing and peeling stresses, together with longitudinal normal stresses
in the adhesive were also ignored. However, they took into considera-
tion adherend shears, using the approach developed by Demarkles
(1955). This was necessary since, for many practical joints, Goland and

FIG. 12. Deformation in single-Iap joint due to lateral straining.


30 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

\
\
\

FIG. 13. Lap joint displacements in x direction (from Adams and Peppiatt,
1973).

Reissner's criterion for neglecting adherend shear strains:

01G2 :%:O'1
°3 G I

is not applicable.
From compatibility considerations in the x-y plane (Fig. 13) we
have:
03 a'Y3x au
ix 2x
---+--+--=10
au
-10 Ix (5)
ax ax ax 2x

From equilibrium considerations in the x-y plane (Fig. 14) we have:

(6)

(7)

(8)

From equilibrium considerations in the y-z plane, z being the coordi-


THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 31

--
-- 2L
O"zx-IL.. _ _ _ _ _ I-·,·"'·"
._7'_2X_ _ _ _....

FIG. 14. Stress equilibrium and directions in lap joint (from Adams and
Peppiatt, 1973).

nate in the width direction:


(9)
The shear displacements of the adherends U 1x and U 2x , may be
found by the method of Demarkles (1955) by which:

TIX=T3X(1- ;J
T2x = T3x(1- ;J
Thus:
U - T3x 8 1 (10)
lx- 2G 1

U - T3x 82 (11)
2x- 2G2

Equations (5)-(11), together with Hooke's law, then give:

(12)
32 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

where
2G l G 2G 3(El ()l + E 2()2)
K =--------~~~~~~~------
a E l ()lE2()i()1 G G + ()2G l G 3 +2()3 G l G
2 3 2)

2Gl G 2 G 3 (V2 E l ()l + V 1E 2 ()2)


K b =-------------------------------
Ei ()lE2()i()1 G 2 G 3 + ()2Gl G 3 + 2()3Gl G 2 )
C = -2PG1G2G3
a Mi ()2 E i() 1G 2 G 3 + ()2 G l G 3 +2()3Gl G 2 )

By consideration of the y-z plane in a similar manner the equation:

(13)

is obtained where:
2V2PG1G2G3
Cb=----------~~~-=~----------
M l ()2E i()1 G 2 G 3 + ()2 G l G 3 +2()3 G l G 2 )

Equations (12) and (13) form a pair of simultaneous partial-


differential equations. The boundary conditions of this system are:
P
atx =0, Ul x = b()l' U2x =0

P
at x = 1, U2x = b()2' Ulx = 0

b
at z = ±2' Ulz = 0, U2x =0

No dosed analytical solution for eqns (12) and (13) is known: a


solution in series may be derived but it would be cumbersome.
However, an approximate analytical solution, which is exact at the
boundaries, or a finite-difference solution may be obtained. For the
approximate analytical solution it is necessary to assume that Ulx is
constant with z and to neglect the term KbUlz which is small compared
with KaUlx.
Equation (12) can then be written as:
a2Ulx
-2- = KaUl x + Ca
aX
This is essentially the equation obtained by Volkersen (1938), Sazhin
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 33

(1964), and Demarkles (1955), the solution being:


_~[ _ (_ ) (l-t/J(l-coshal))sinhaX]
Ulx - b 1 t/J 1 cosh ax . h I (14)
Sl SIll a
where:

From eqn (2):

_ Pa [(1- t/J( 1 - cosh al)) cosh ax ./, sinh ax] (15)


'T3x - b . h I
SIll a
'f'

Equations (14) and (15) are exact for the boundaries z = ±bj2
because here U lz = 0.
Equation (13) then becomes:

giving the following solution for the normal transverse stresses:

and for the transverse-shear stresses in the adhesive:

sinh az
(ab)
SlS2a(V1UlxE2 - V2U2xEl)
(17)
'T3z =
(SlEl + S2 E 2) cosh 2

When x = 0, Ul x = PjbS l and U2x = 0, thus eqns (16) and (17) are exact
at the boundaries x = 0, and x = l.
The (complete) eqns (12) and (13) were also solved by a finite-
difference method, and the results are compared with the approximate
analytical solution in Figs 15 and 16.
The results for a 25·4mmx25·4mm (1 inx 1 in) lap joint show that
the form of the adherend tensile stress and the adhesive shear stress (in
the longitudinal direction) were much as would be expected from a
Volkersen (1938) type solution (Fig. 15). However, the transverse
stresses show direct (tension or compression) and shear stress maxima
at the ends of the joint (see Figs 16(a) and (b)).
34 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

16

14
\.
12
~
~
~x
10 P
!!:
)(

'"
1''''
'"
a.' '-6
)( ",'\
b 4 I
\.

0
0 OA 0·5 0·6 0-7
x
Tri
- - - Approximate analytical solution (exact at z = ±b/2).
o Finite-difference solution at joint centre-line (z = 0).
- -- Approximate analytical solution neglecting adherend
shears.

FIG. 15. Tensile stress in adherend and shear stress in adhesive (each per unit
load) plotted against x for a 1 in x 1 in single-Iap joint (from Adams and
Peppiatt, 1973).

Thus it is possible, by using closed-form analyses of varying com-


plexity, to predict the stresses in simple lap joints. (This approach is
termed continuum mechanics.) In many instances, such solutions may
be deemed acceptable. However, two problems still remain to be
solved if it is required to predict the strength of real joints. These may
be summarized as end effects and material non-linearity (adhesive and
adherend plasticity). We wi11look first at end effects for linear elastic
systems.

THE SINGLE-LAP JOINT-END EFFECTS

One common result from all the closed-form analyses, whether they be
simple or complex, is that the maximum adhesive stresses always occur
near the end of the bond-line.
The closed-form algebraic lap joint analyses, which have been
discussed so far, have all assumed that the adhesive layer ends in a
~
tTI

1·6 ~
c:
x.I·Oin x=O
"IZ/P I· ~
:-;;;=z-
3 >-Z
x=o·o ~~
!\~~8 t:I

OAI ~
0·1 ~
--==l ::J
0 c:
z I t:I
"üSin -0-4 tTI
o
~
-0·8 CI"l

-Zrx.I.Oin ~
-I·Z, 'x. 1·0 in CI"l
x-o CI"l
a -3 tTI
CI"l
b -1·6
Z
FIG. 16. (a) Transverse direct stress per unit load (U1Z/P) in the adherend plotted against z for a 1 in x 1 in single-Iap joint >-t:I
at different positions along the length of the joint: - - appraximate analytical solution (exact at x = 0, 1 in); 0 ::r:
finite-difference solution; (b) transverse shear stress per unit load ('T3z/P) in the adhesive plotted against z for a 1 in x 1 in tTI
r:/)

single-Iap joint at different positions along the length of the joint (fram Adams and Peppiatt, 1973).
~
.....
o
~
w
U1
36 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

AdheSive
layer

AdheSive layer wlth


~-~-square edge farmmg
free surface
Adh ere nd
p
a /

FIG. 17. Diagrammatic lap joints to show adhesive layers with (a) square edge;
(b) spew filIet (from Adams and Peppiatt, 1974).

square edge as is shown in Fig. 17(a). Coker (1912), experimentally


(using photoelasticity), and Inglis (1923) analytically, have shown that
a rectangular plate with shear loading on two opposite sides experi-
ences high tensile and compressive stresses at its corners, the mag-
nitude of these being about four times the applied shear stress, and the
direction being at right angles to the sides on which the shear load is
applied. These transverse direct stresses arise because the direct and
shear stresses acting on the free surface must be zero. If the adhesive
layer is assumed to have a square edge, it would be expected that
similar tensile and compressive stresses must occur in the corners of
this layer, because of the free surface.
However, real structural adhesive joints do not have a square edge
but are formed with a fillet of adhesive spew (see Fig. 17(b)), which is
squeezed out under pressure while the joint is being manufactured.
The assumption that the adhesive layer has a square edge is thus
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 37

unlikely to be realistic. Mylonas (1954) has investigated the stresses


indueed at the end of an adhesive layer for a number of adhesive edge
shapes using photoelastic teehniques and has shown that the position
of the maximum stress is dependent on the edge shape. The adherends
in his model were rigid, and none of the edge shapes studied were
typical of the normal spew fillet, where the adhesive ftows round and
bonds to the end of the adherend.
It was shown above that the existing closed-form solutions predict
that the highest stresses should be near the ends of the joint. However,
they do not take into account the inftuence on these stresses of the
spew fillet which is formed at the ends, and so it is in just these regions
of maximum stress and where failure is bound to oecur that the
assumed boundary conditions of the previous theories are the least
representative of reality.
Sinee it is unlikely that a closed-form analytical solution will be
found which ean be used to predict the inftuenee of the adhesive spew
on the stresses in adhesive joints, the solution to the problem must be
found in numerical techniques. The finite-element method, as de-
seribed for example by Zienkiewicz (1971), is now a well-established
me ans for mathematically modelling stress analysis problems. Its great
advantage lies in the fact that the stresses in a body of almost any
geometrical shape under load ean be determined. The method is
therefore well-suited for analysing an adhesive joint with a spew fillet.
The method also avoids the approximations of the closed-form
theories, presented earlier, in neglecting the strain energy of certain
stresses within the joint, and thus enabling more accurate answers to
be found outside the bounds of the Goland and Reissner criteria.
One of the authors (R. D. Adams) and his co-workers (N. A.
Peppiatt, J. Coppendale, A. D. Crocombe, J. A. Harris and Z. Chen)
have been some of the main investigators in this critical field and this is
frequently referred to in this book. They have used finite-element
techniques (FET) almost exclusively. Wooley and Carver (1971) were
among the first to apply the FET to a single-Iap joint. They carried out
a parametrie study, investigating the effects of adhesive modulus,
overlap length, and adhesive thickness, but still modelled the adhesive
as having a square end. Furthermore, no attempt was made to refine
the mesh in the regions at the overlap ends. Nonetheless, good
correlation was found with closed-form analyses. In the finite-element
model of Harrison and Harrison (1972) the adhesive was taken to be a
square-edged layer between rigid adherends and so the effects of spew,
38 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

bending and differential straining were not considered. Wang et al.


(1976) also carried out a parametric investigation, using a greater
degree of mesh refinement than Wooley and Carver, and showed that
the stresses can vary considerably across the adhesive thickness. They
also inc1uded a spew fillet, but this was restricted to the height of the
adhesive layer.
In their early work, Adams and Peppiatt (1974) used constant strain,
two-dimensional, triangular elements which give the stress at the
centroid of the element. Since joints tend to be wide compared with
the thickness, the problem was considered to be one of plane strain.
This assumption should be satisfactory for the adhesive layer, but less
so for the adherend. The spew fillet was approximated to a 45°
triangular fillet of varying size. The restraints used in their models are
shown in Fig. 18 and a typical mesh in Fig. 19.
It is difficult to carry out experimental stress analysis on the adhesive
layer in a typical joint. Adams and Peppiatt (1974) therefore con-
structed a model consisting of silicone rubber cast between two, rigid,
steel adherends. This model was based on the earlier experimental
stress analyses by Adams et al. (1973) using hard rubber for the
adherends and soft (foam) rubber for the adhesive. The model is
shown in its deformed state in Fig. 20(a), and good agreement is shown
between the predicted and actual deflections. Figure 20(b) shows the
principal stress pattern obtained by the finite-element analysis. The
length and direction of the lines represent respectively the magnitude
and direction of the principal stresses at the centroid of each finite
element. A bar at the end of the line implies a negative principal stress,
i.e. compressive. It is evident that the presence of the fillet causes the

r -__________-L/~1======~----------~I==:p
~ L A A A A A 1'-
_ _ _ _ _ _ _ _ --..,. _ _ _ _ -..
JT--- - - - - - - - - - - ,
L _____ ________ J
(a)

(b)

FIG. 18. Constraints for finite-element models of lap joints: (a) full-Iength
double-Iap joint; (b) single-Iap joint (from Adams and Peppiatt, 1974).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 39

This rt910" IS sl'1awn


ifnlarqfd brlo.

FIG. 19. Finite-element mesh for fulliength lap joint with spew (from Adams
and Peppiatt, 1974).

stress pattern to differ significantly frorn the pattern at the end with no
fillet. At the points Al and A 2 , the high tensile and compressive
stresses predicted by Inglis (1923) are shown, the absolute magnitude
of the largest element al principal stresses being at least 3·6 times the
shear stress in the rubber between the plates. This is of similar size to
the value predicted by Inglis who says that the normal stress is more
than 4 times as large as the applied shear stress. It should be noted that
the rubber away from the ends of the steel plates is in pure shear, as is
shown by the equal and opposite principal stresses in the elements in
this region. The stresses in the fillet are predominantly tensile, the
maximum principal tensile stress being at least 3·5 times the shear
stress in the rubber between the plates.
The stress pattern at the end of a square-edged adhesive layer in a
lap joint is shown in Fig. 21. This is again a plot of principal stresses,
the interpretation of which was given in the discussion of the rubber
model. The highest tensile stress exists at the corner of the adhesive
40 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

FIG.20(a). Comparison between ca1culated and experimental displacements of


the silicone rubber model. (The black crosses are the finite-element predictions
of the intersections of the grid lines of the model.) (from Adams and Peppiatt,
1974.)

FIG. 20(b). Principal stress pattern for silicone rubber model showing end
effects (from Adams and Peppiatt, 1974).

adjacent to the loaded adherend and represents a stress concentration


of at least 10 times the applied shear stress on the joint. It should be
noted that, because constant-stress elements were used and the stress
gradients are high, it is impossible to determine the stresses just below
the surface very accurately without using infinitesimally small ele-
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 41

/
,/

- , /
/

FIG. 21. Finite-element prediction of the principal stress pattern at the end of
a square-edged adhesive layer (from Adams and Peppiatt, 1974).
ments: the stresses acting on the surface are, of course, zero. The effect
of bending of the outer adherends modifies the stress distribution from
that obtained in the simple rubber model, the absolute value of the
largest principal stress at this end being about 4 times the absolute
value of the largest principal stress at the other side of the adhesive
layer (compare points Al and A 2 in Fig. 20(b).
The effect of the spew on the stress pattern is shown in Fig. 22,

/ /

//
.,

FIG. 22. Finite-element prediction of the principal stress pattern at the end of
an adhesive layer with 0·5 mm spew (from Adams and Peppiatt, 1974).
42 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

which is at the tension end of a double-lap joint. The spew is


represented by a triangular fillet 0·5 mm high. It can be seen that,
because of the predominance of the major principal stress, the adhe-
sive at the ends of the adhesive layer and in the spew fillet is essentially
subjected to a tensile load at about 45° to the axis of loading. The
highest stresses occur within the spew at the corner of the unloaded
adherend, the presence of the 90° corner introducing a stress-
concentration effect. As the maximum stress occurs within the spew and
not at or near the adhesive surface, it is unlikely that the approxima-
tion to the spew shape by the triangular fillet has a significant effect on
the stress distribution.
The stress pattern shown in Fig. 22 suggests that the area of transfer
of load between the adherends is effectively lengthened. Figure 23
shows the average shear stress distributions along the adhesive layer
obtained for joints with varying sizes of spew. The distributions are
shown dotted outside the overlap length to show an area of load
transfer within the spew. As the loading in the spew is predominantly
tensile, a meaningful distribution of shear stress is difficult to obtain.
The maximum shear stress obtained from the largest spew size,
which extends completely across the ends of the adherends, is 70% of
that obtained for the square-ended adhesive layer. A spew of 1·0 mm
height gives a shear stress reduction of 15%. It can be seen that, when
the effects of spew are taken into account, the maximum shear stress
obtained is significantly reduced.
It has been observed that, in the spew of aluminium to aluminium
lap joints bonded with low-ductility adhesives, cracks are formed
approximately at right-angles to the directions of the maximum princi-
pal stresses predicted by the finite-element analysis. In general, these
cracks run dose to the corners of the adherends. The region where
cracks are formed in the spew is indicated in Fig. 22. These observa-
tions give weight to the view presented here that failure in a lap joint is
initiated by the high tensile stresses within the spew. The cohesive
failure of the adhesive occurs in this manner in normal, well-bonded
joints. (In adhesives terminology, 'adhesive failure' means the destruc-
tion of the bond between adherend and adhesive; 'cohesive failure'
means a fracture wholly within the adhesive material.) If, however, the
boundary of the adherend and adhesive is weak, the spew is not
cracked but is pulled away from the loaded adherend surface by the
tensile stresses in the spew.
Figure 24 shows a typical crack in the spew fillet of a double-Iap
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 43

--e- Q SQUARE EDGE


~b 0'25mm SPEW (EQUALS GLUE-LINE THICKNESS)
-e-c 0'50mm SPEW
-a-d 1'0 mm SPEW
6'0
-*-e FULL DEPTH SPEW

5'0
,
1E \
~\
b 4'0 1\ \
X
'0
~~
nJ
Q 1\ \\
1I \
1\ \ \
1\ \ \
II \ \
II \ \
II \ \

"0
:I \ \
II \ \
I I
O·OL-._.....I_ _--1_ _....L._ _......L._ _...L-_ _L.....I-J....LlL--.L....--'...L-_-"
00 "0 2'0 3'0 4,0 5'0 6'0 7,0 8'0
Distance trom centre ot overlap (mm)

FIG. 23. Influence of spew size on shear-stress distribution at tension end of


double-lap joint (from Adams and Peppiatt, 1974).
44 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Outer adherend

Central adherend

Fro. 24. Typical crack on loading double-lap joint made with spew fillets
(joint has been cut and polished).

joint. The load on this joint was released as soon as the crack occurred
so that it did not propagate. Acoustic emission monitoring was used to
indicate the initial failure. Adams et al. (1978a) also showed that the
crack starts at the corner and runs out to the free surface. They did this
by inserting a thin metal shim in the spew fillet to act as a crack
stopper. The result is shown in Fig. 25.
When a lap joint fails completely, the initial crack in the fillet is
turned to run along (or dose to) the adhesive-adherend interface. It
meets a similar crack running in the opposite direction and we have the
familiar result shown schematically in Fig. 26.

Central adherend

Outer adherend

FIG.25. Crack sirnilar to that shown in Fig. 24 but stopped by thin metal shim.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 45

FIG. 26. Diagram of failure surfaces of single-Iap joint.

It should be noted that complete removal of the spew by machining


would be difficult without machining either of the adherends, so that
some spew, similar in size to the glue-line thickness, is likely to be left
at the end of any joint. Moreover, machining may initiate cracks in the
adhesive, especially if it is one of the more brittle, high-temperature
adhesives. Thus an adhesive layer with a square edge is not only
undesirable but is unlikely, and even difficult to obtain in practice.

Reduction of Stress Concentrations


One of the results of the finite-element analysis is that the highest
principal stresses in the adhesive layer of a lap joint occur not only
near the ends of the bond-line but, more precisely, in the region of the
corner of the unloaded adherend. This suggests that it may be possible
to increase joint strength by radiusing the corner of the adherend.
Adams and Peppiatt (1974) analysed a 0'4mm radius on the unloaded
adherend corner and the maximum principal stress in the spew was
shown to be much less than that obtained with a right-angle corner
(see Table 2).
Standard 1 in (25·4 mm) by 0·5 in (12·7 mm) lap joints were pre-
pared with L73 aluminium alloy adherends, half of which had a
0·4 mm radius hand-reamed on the adherend corner. The surfaces
were prepared by a standard etching procedure and bonded with
AF130 adhesive. The joints were then tested in tension in accordance
with ASTM D1002-64. The strengths obtained are given in Table 2
together with the finite-element predictions. No significant increase in
strength was obtained, the small improvement (less than 3%) with
rounded corners certainly not being of the size predicted by the
46 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 2
THE EFFEcr OF ADHEREND CORNER SHAPE ON PREDlcrED AND AcruAL LAP
JOINT STRENGTHS

Adherend corner shape


Square 0·025mm 0·4 mm radius
chamfer

Predicted strength* 0·56 0·81 1·0


Actual strength* 0·97 1·0

* Relative to strength with 0·4 mm radius.

finite-element model. The reason for this is that etching the adherends
produces a radius on the adherend corner (see, for example, Fig. 25),
even on the nominally rectangular edges. The finite-element model
was therefore modified so that the adherend had a small chamfer
(0·025 mmxO·025 mm) at its corner. The maximum predicted adhe-
sive principal stress was now much doser to that of the joint with the
rounded adherend (Table 2).
Recently, Crocombe and Adams (1981b) have studied how a spew
fillet can affect the adhesive stress distribution over a range of material
and geometric properties. Also, they investigated the stress distribution
ac ross the adhesive thickness, a variable which has been assumed
constant in most other analyses (but notably not by Allman (1977) for
peel stresses). They used a more advanced system of finite elements
than had been used by Adams and Peppiatt, choosing plane-strain,
two-dimensional' rectangular, quadratic, isoparametric elements (see
e.g. Zienkiewicz (1971) for further details). The mesh was progres-
sively refined until a stable stress distribution was obtained. A typical
mesh and boundary conditions are shown in Fig. 27. In this mesh,
there are four elements across the adhesive thickness in the overlap
region. They used low loads so that the results are as for a single lap
with the bending moment factor of Goland and Reissner k = 1, or for a
double lap. Figure 28 shows how the adhesive principal, peel and shear
stresses vary across the adhesive thickness at different distances from
the overlap end. Except within a few adhesive thicknesses from the
end, the stresses are essentiallY uniform: the principal and peel stresses
smooth out more rapidly than the shears. A possible explanation for
this stress variation across the adhesive thickness near the overlap end
is the discontinuity caused by the unloaded adherend. Figure 29 shows
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 47

2C

)
(

~ A
\
C
)

FIG. 27. Single-Iap joint geometry and finite-element mesh (from Crocombe
and Adams, 1981b).

the same stresses plotted against distance from the overlap end along
three planes in the adhesive parallel to its length, on the unloaded
adherend-adhesive interface, on the adhesive central plane and on the
loaded adherend-adhesive interface. Again, the variation in the stres-
ses increases towards the overlap end. The maximum value of the
shear stress is seen to occur just within the overlap. The peak in the
stress distributions appears to shift from the overlap end further into
the spew as the distance of the plane of the stresses from the unloaded
adherend increases. This is probably because load transfer in the lap
joint does not occur perpendicularly across the adhesive thickness. The
maximum loads are likely to travel from the loaded adherend before
the overlap and across the adhesive to the corner of the unloaded
adherend; the positions of the peak stresses may therefore correspond
with this load line. The positions of the maximum of all the adhesive
stresses occur under the overlap, the maximum principal stress being at
the unloaded adherend corner, acting at about 45° to the longitudinal
axis of the joint (a feature first noted by Adams and Peppiatt (1974)),
the maximum peel stress being at the overlap end just within the
adhesive layer, and the maximum shear stress being on the adhesive-
adherend interface at a small distance from the overlap end: these
00
"'"
sl~'
.4~ \.
~h .2 .2

~
c::
~
::c
X- Tb/2 ~
~
.2 .1 .1
Si
CI)
X-2Tb
~
.....
o
~
Z
X· 2Tb ~
o .05 .10 .15 .20 0 .05
~ .10 15 .200 05 .10 .15 .20
Y/mm 2J_N~ Y/mm Y/mm ~
, ~
Cl

FIG. 28. Variation of adhesive principal (O'prin), peel (0',), and shear (7'xy) stresses across the adhesive thickness (y) at
various distances (x) from the overlap ends (from Crocombe and Adams, 1981b).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 49

"-re
l51e
02

01

0'2 o 02 04 -02 o 02 04 -02 o 02 04


x/mm X/mm X/mm

FIG. 29. Variation of adhesive principal (Uprin), peel (U y ) , and shear (TxY)
stresses with distance from the overlap end (x) at various longitudinal planes
through the adhesive: +unloaded adherend interface; 0 adhesive mid-plane;
x loaded adherend interface (from Crocombe and Adams, 1981b).

points are labelled A, Band C respectively in Fig. 27, and do not vary
widely over the range of configurations analysed.
Comparison was made with the analysis by Allman (1977) by
averaging the finite-element stresses across the adhesive thickness.
Results for adhesive peel and shear stresses are shown in Fig. 30.
Several points should be made here. First, when a square-ended
adhesive layer was modelled, the results agreed closely with Allman's.
However, when a full-depth spew filIet was used, both stresses were

05 03

~~04
OE 03 ~~02
02

01 01

0
8
'01 00 2 3 4 5 6 7 8
x/mm

FIG. 30. Variation of averaged adhesive peel (uy) and shear (Txy) stresses with
distance (x) from the centre of the overlap: - - Allman; +finite element with
square end; V finite element with spew fillet (from Crocombe and Adams,
1981b).
50 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

considerably reduced. This is partly because the fillet carries a signific-


ant proportion of the total load. However, another discrepancy results
from Allman's assumption that the peel stress varies linearly across the
glue-line thickness, whereas Fig. 28 shows it to be more nearly an
exponential variation. The linear approximation gives higher averaged
stresses than are actually present.
However, when a finite-element analysis is made of a square-ended
adhesive layer (as analysed by Allman), the comparison worsens mar-
kedly. Figure 31 shows the magnitude of the principal stresses pre-
dicted for a square-ended adhesive layer and one with a full-depth
spew fillet. It can be seen that a high, tensile principal stress is
predicted at the loaded adherend surface in Fig. 31(a) and small
compressive stress at the corner. Averaging these stresses gives us 59
units. Any failure criterion based on average stresses with square-ended
layers would therefore be in error by about 100%! It is also in the

Loaded adherend
12' 34 31
38 53 36

14 43 36

-6 41 38

(a) Unloilded adherend


/

265 20

Ju5 23
Unloaded
(b) adherend
/

FIG. 31. Finite-element principal adhesive stresses around the end of the
overlap for: (a) square-edged joint; (b) joint with fun depth spew tillet,
carrying the same load (arbitrary units).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 51

wrong direction as far as safety is concerned. However, for the same


joint geometry but with a full depth spew fillet (as shown in Fig. 31(b))
the large stress gradient has largely disappeared and the average
principal stress is only 28·6 units! A similar calculation with a smaller
fillet (one-third full height) gave an average of 40·75 units (35·25 and
46·25 at the corner). The increase in stress was due in part to the
reduced load-carrying capacity of the smaller spew fillet. But the
important point is that, even in this latter case, the maximum value of
the principal stress is only 37·3% of that predicted for the square-
ended layer, and its position has shifted across the glue-line to the
corner of the unloaded adherend.

THE SINGLE LAP JOINT-ELASTO-PLASTIC ANALYSIS

In the earlier part of this chapter, it was shown that increasingly


complicated mathematics is required if the stress situation in a single
adhesive lap joint is to be determined. Even so, the law of diminishing
returns applies, together with the irony that end effects, particularly
where a spew fillet is involved, are the most difficult to model accu-
rately while this is the most critical region since failure almost always
occurs here.
But yet there is one more irony to be endured! Modern adhesives,
particularly those such as the rubber-modified epoxies, have a large
plastic strain to failure. Thus, we need now to consider what happens
to the stress distribution when the adhesive can yield. Further, these
new adhesives are so strong that the adherends too may be caused to
yield. In fact, even with the old, brittle adhesives, the adherends in
single-Iap joints often yielded plastically in bending. Physically, two
opposite effects occur when the adherends yield. In Fig. 7(b) the
effect of adherend differential straining was shown to cause adhesive
shear stress peaks towards the bond-line end. If the adherends yield at
their loaded end, the differential straining is enhanced and so the
adhesive stresses will be increased, leading to premature failure. How-
ever, if the adherends are stressed to yield, they will more easily rotate
under the effect of the non-colinear applied loads. This causes the
Goland and Reissner joint factor k to decrease more than if the
adherends remained elastic, thus reducing the stresses. We will there-
fore need to study not only adhesive plasticity but also adherend
plasticity.
52 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

There have been two basic approaches to studying the stresses in lap
joints when plasticity occurs. The first of these to be considered is
based on the use of continuum mechanics, while the second uses
finite-element techniques.
The main advocate of continuum mechanics is L. J. Hart-Smith who
has produced an enormous amount of work on this subject on the
P ABST programme (Primary Adhesively Bonded Structure Technol-
ogy) under contract to the USAF Flight Dynamics Laboratory. This
method is a development of the shear-lag analysis of Volkersen (1938)
and the two theories of Goland and Reissner (1944). While Hart-
Smith has published many papers on the analysis of joints (see, for
example, 1973a, b, c; 1978a, b; 1980a, b) the reader is referred to his
excellent review (1981). The design philosophy behind Hart-Smith's
work is that the adhesive should never be the weak link. Thus, if peel
stresses are likely to occur, they should be alleviated by tapering the
adherends (scarfing) or by locally thickening the adhesive layer. Hart-
Smith's continuum mechanics approach has the advantage over the
finite-element technique that it allows a parametric investigation con-
cerning the effects of glue-line thickness, joint length, and so on,
together with adherend and adhesive mechanical properties, to be
carried out at low cost.
The first difficulty is how to characterize the adhesive. Hart-Smith
choose an elastic-plastic model (see Fig. 32) such that the total area
under the stress-strain curve was equal to that under the true stress-
strain curve. If the maximum stress is less than yield, the true elastic
curve may be used, while for a peak stress intermediate between yield
and failure a different, and more accurate, model is chosen. The
bi-linear model is eIoser to the true adhesive characteristic over its
entire range of loads, so that a single model can be used for calculating
the stresses without having first to ca1culate these to establish which
intermediate elastic-plastic model should be used! It is argued that the
model or models to be adopted for a given situation will be a
compromise between precision and convenience.
Hart -Smith has developed computer pro grams for analysing various
joints, double-lap and single-lap, with equal or dissimilar adherends,
for parallel, stepped, scarf and double-straps. Similar programs are
available from ESDU (Engineering Sciences Data Unit, 251-9,
Regent Street, London) for elastic and elastic-plastic calculations. The
ESDU programs were developed from the work of P. Grant (1976,
1978) of British Aerospace and are again based on the work by
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 53

/
ADHESIVE /
I
SHEAR ,.........."1.......... . ' .... ".
STRESS ! / . - .~_. ELASTIC·PLASTIC MODEL
._.j.. . /' - . TRUE CHARACTERISTIC
r /
; ! / BILINEAR MODEL
T.I ! / ELASTIC·PLASTlC MODEL
f / FOR PARTIAL LOAD LEVEL
! I
i /
! I
! /
! I
! I

!/~ NOTE: AREA UNDER STRESS-STRAIN CURVES IS THE SAME.

\
!/ Gelpl

!~"
:/ Gel

le
ADHESIVE SHEAR STRAIN 'Y

FIG. 32. Adhesive shear stress-strain curves and mathematical models (after
Hart-Smith, 1981).

y
FIG. 33. Theoretical adhesive stress-strain curve (ESDU/Grant model) (from
ESDU 79016, 1979).
54 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Volkersen (1938) and Goland and Reissner (1944). In the


ESDUjGrant method, the adhesive is modelIed as shown in Fig. 33
with the following conditions:
If 'Y < 'Ye, then T= 'YGe
T=Te+(~)
a+ß
where a = 'YGe - Te and ß = Tmax - Te·
In efIect, Hart -Smith equates the yield stress and the failure stress,
and says that failure occurs when the adhesive reaches its limiting
shear strain. This is illustrated in Fig. 34 which shows the adhesive
shear stresses and strains for a double-Iap joint as the applied load is
progressively increased. It should be noted that the shear strain dis-
tribution is not simply a multiple of the low-Ioad case since the
assumption of a limiting (plastic) shear stress in the adhesive layer
causes a distortion to the elastic shear-Iag theory.
Figure 35 shows the efIect of lengthening the overlap in a double-Iap
joint which has been loaded so that plastic deformation of the adhesive
has occurred. It can be seen that there is still a large, low-stress region

JOINT GEOMET ....

.. O~EIIVE IHEAR ITREIS

CitlTICAl SI-IEA_ snAIN


-- fAllUIi:ES

FIG. 34. Development of shear stress and shear strain distribution in a double-
lap joint with increasing load (after Hart-Smith, 1981).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 55

ADHESIVE BOND STRESS;


DISTRIBUTION

A. SHORT OVERLAP

JOINT CROSS SECTION

l- P

B. INTERMEDIATE OVERLAP

;- ; : }-
P

TI~
C. LONG OVERLAP

FIG. 35. Influence of lap length on bond stress distribution lafter Hart-Smith,
1981).

in the middle of the joint. Far from accepting that this is structurally
inefficient, the PABST/Hart-Smith design philosophy claims that if is
essential to overcome the effects of creep at the ends if low cycle
creep/fatigue loading is encountered. Figure 36 summarizes Hart-
Smith's design criteria for double-Iap joints. The overlap is designed for
the worst operating condition which is usually when the adhesive has
beeil softened by moisture absorption and temperature. For this worst
case, the plastic zones must be long enough to carry the ultimate load,
the elastic region in the middle must be big enough to prevent creep,
and the minimum operating stress (elastic minimum) must be no
greater than 10% of the adhesive shear strength.
For single-Iap joints, where peel stresses are significant, especially at
the ends of the overlap, the P ABST philosophy is to reduce theI11 by
tapering the adherends and by increasing the overlap. For instance, the
ratio of overlap length to adherend thickness in a typical laboratory
56 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

I a~ :
~~ l
r G

1E' lt~ ~
t uu1t
2 Tp -1 ~
I Tp ADHESIVE SHEAR
STRESS DISTRIBUTION

H
• PLASTIC ZONES LONG ENOUGH FOR ULTIMATE LOAD
• ELASTIC TROUGH WIDE ENOUGH TO PREVENT CREEP
AT MIDDLE
• CHECK FOR ADEQUATE STRENGTH
FIG.36. Design of double-Iap bonded joints (after Hart-Smith, 1981).

lap-shear test (0·5 in long x 0·063 in thick; 12·5 mm x 1·6 mm) is about
8 : 1 and such joints usually experience adherend yielding prior to bond
failure. In the PABST programme, a ratio of 80: 1 was chosen: weight
for weight, these joints were only 10% weak~r than a double-Iap joint,
and much easier to manufacture and inspect.
It should be borne in mind that the PABSTjHart-Smith and the
ESDUjGrant design philosophies were developed for a specific appli-
cation, that of aircraft construction, and may not be generally applica-
ble to other aspects of the engineering usage of adhesives. However, it
does represent a major and a successful input to the design of adhesive
lap joints.
As mentioned already, the basic Hart-Smith approach entails neg-
lecting the normal or 'peel' stresses acting across the glue-line. How-
ever, in practice these may be the main contributors to joint failure,
even in double-Iap joints. Hart-Smith has recognized this possibility
and in one of his analyses (1972) combined elastic peel stresses with
plastic shear stresses. Thus he gives, for a double-Iap joint, the 'peak
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 57

peel stress', O"p, as a function of the 'peak shear stress', Tp , as:


O"p = [3E~(1- V2)to]~ (18)
Tp EoTl
where to is the adherend thickness, TI is the glueline thickness, E~ is the
effective transverse adhesive Young's modulus, and E o is the adherend
Young's modulus.
This is illustrated in Fig. 37. From this it is argued that, for
sufficiently thin adherends, the 'peeling' stresses are not important.
This approach may be applied to the failure of composite joints, where
failure is due to the transverse stresses in the laminate as shown in Fig.
38, which also shows how profiling the adherend can reduce the peel
stresses, change the failure mode, and hence increase joint strength.
When we consider non-linear material properties by a closed-form
analysis such as Hart-Smith's, the limitation is how tractable is a
realistic mathematical model of the stress-strain curve within an alge-
braic solution. With the finite-element techniques developed for adhe-
sive joints by Adams and his co-workers, the limit becomes that of
computing power. The high elastic stress and strain gradients at the
ends of the adhesive layer need to be accommodated by three or four
8-node quadrilateral elements across the thickness. However, consid-
eration of non-linear material behaviour requires a much larger com-
puting effort on any given element. Thus, it becomes necessary to

...
~10'0
Critical loeation
f;;

-~~
~ 5·0
"tJ t.~? Adhesive
1:" <ll<lm"nt
~ 2·0
P e
tf

'"~ "\ E'c EO,V


'"
-.
.
~
1·0

1ii 0.5
"ii
GI

'"
g
"tJ
0·2

'"~ 0·1 . I ... 1 I . . I .•.. 1 I . , I , .. "


Il. 0·01 0·02 0·05 0·1 0·2 0·5 1·0 2·0r::5'0 E,1~~
Non-dimensionallzed geometrie thiekness paramet<lr C(1-V2 ) E~.o

FIG.37. Peel stresses in double-Iap bonded joints (after Hart-Smith, 1972).


58 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

:- - --< • - •••s;;;. .
J -_v--~ 1t~-[-=========.7=:::;ir.,...---J
Charactllrlstic failurCl mode
for thick composite adherends
:: }
L..-_ _ _ __ _

Alternative peClI strClss


relief techniqulls

Bond pllel stresses Raduced bond peel strellslls

Associatlld shllar stresses Highllr aVllrage shllar stress

FIG. 38. Relief of peel stress failure of thick composite bonded joints (after
Hart-Srnith, 1972).

reduce the elements, perhaps to only one across the adhesive thick-
ness. Apart from the computing, it is necessary first to define yield (of
the adhesive usually but it can also be the adherend) and then to adopt
a suitable failure criterion.
Let us first consider the yield behaviour of an adhesive. Adams et al.
(1978a) analysed double-Iap joints in which two adhesives of different
strengths and strains to failure were used. It is, of course, widely
accepted that the yield behaviour of many polymers, including epoxy
resins, is dependent on both the hydrostatic and deviatoric stress
components. A consequence of this phenomenon is the difference
between the yield stress in uniaxial tension and compression. For
epoxy resins, Sultan and McGarry (1973) obtained ratios of compres-
sive to tensile yield stresses of 1·28 and 1'35: Pick and Wronski (1976)
obtained a ratio of 1·33. Bowden and Jukes (1972) showed that the
ratio of compressive to tensile yield stress increased with the amount of
plasticizer present in an epoxy. Tests on bulk specimens of AY103 and
MY750 have given ratios of 1· 27 and 1·14 respectively. This be-
haviour has been incorporated into the analysis by assuming a
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 59

paraboloidal yield criterion of the form:


(UI-U2?+(U2- U3?+(U3- Ul?+2(lucl-UT)(Ul +U2+ U 3)
= 21uc l UT (19)
where Ub U2 and U3 are a combination of principal stresses causing
yield and IUcl and UT are the absolute values of the uniaxial compres-
sive and tensile yield stresses. This type of yield criterion was proposed
by Raghava et al. (1973) and was shown by them to apply to several
amorphous polymers over a wide range of stress states. It should be
noted that when \uc ! = UT the paraboloidal yield criterion reduces to
the more familiar von Mises cylindrical criterion.
Tensile stress-strain curves were obtained from tests on bulk* speci-
mens of two Ciba-Geigy adhesives, A Y103 and MY750. A Y103, a
plasticized epoxy, was mixed with 17 pphw (parts per hundred by
weight) of HY956 and cured for 1 hat 100°C. MY750, an unmodified
epoxy, was mixed with 85 pphw of HY906 and 2 pphw of DY062 and
cured for 3 h at 80°C followed by 4 h at 125°C. The specimens were
machined from cast blocks which had been subjected to the same cure
schedule as the lap joint specimens. The gauge lengths were carefully
polished with metal polish to remove any surface scratches. For the
purpose of representing the stress-strain curves in a convenient form
for the computer program, the stress in the non-linear region was
expressed as a third-order polynomial function of the plastic strain
component. Typical tensile stress-strain curves for the two adhesives
and their polynomial approximations are shown in Fig. 39. Typical
strains at failure were 5% for the AY103 and 3% for the MY750
while the MY750 was clearly the stronger adhesive.
The finite-element analysis predicted that yield would first occur at
the sampling point nearest the corner of the outer adherend and the
highest strains were predicted in the adhesive at the adherend corner.
Figure 40 shows the spread of the zones of plastic deformation at the
unloaded end of the outer adherend as the load was increased. Figure
41 shows the relationship between the applied load and the maximum
principal strain predicted at the adherend corner, for both adhesives in
double-Iap joints having an overlap length of 12·7 illill and an adhesive
thickness of 0·13 illill. It should be noted that, at a maximum strain of
3%, the AY103 joint is supporting less load than the MY750 with a

* These two epoxies were chosen for this investigation because they could be
readily cast into large blocks from which bulk specimens could be machined.
60 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

90

80

70
-'"
/.
~
AY103

60

111
111
W
...111
Cl: 50

w "
lL
.Jl:
iii
z
...w 40

30

20

10

o 0.01 0.02 0.03 0.04 005


STRAIN

FIG. 39. Tensile stress-strain curves of adhesives: - - stress-strain curves


from bulk specimens; - - - - approximations to stress-strain curves (from
Adams, Coppendale and Peppiatt, 1978a).
FIG. 40. Incremental zones of plastic deformation at tension end of double-Iap
joint. The contour lines for each increment enc10se the sampling points at
which plastic flow takes place. (Increment 1 corresponds to 8·81 kN load;
subsequent increments at 2·34 kN intervals.) AY103 adhesive; 12·7 mm over-
lap; 0·13 mm adhesive thickness (from Adams, Coppendale and Peppiatt,
1978a).

20

15

0
«
9
....
Z
.,Ö 10

a.Z
«~
.J

LII
.J

g
CD

"'AX. PRINCIPAL STRAIN IN AOHESIVE AT AOHERENO


CORNER

FIG. 41. Double-Iap joint load versus maximum principal strain in adhesive;
12·7mm overlap; O·13mm adhesive thickness (from Adams, Coppendale and
Peppiatt, 1978a).
62 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

similar strain. However, if the AY103 is able to withstand a 5% strain


before failure, the load carrying capacity of the joint is 27% higher
than that of the MY750 joint at a 3% strain. This is despite the fact
that the tensile strength of A Y103 at astrain of 5% is less than the
tensile strength of MY750 at astrain of 3%. It should be remembered
that the strain at failure of an adhesive in the highly stressed region of
a lap joint will not necessarily be equal to the strain at failure of a bulk
specimen subjected to a uniaxial tensile stress. However, in the ab-
sence of other data, these values should give a qualitative estimation of
the failure strains of the two adhesives in a lap joint.
Figure 42 shows the predicted variation of joint strengths with
overlap length of double-Iap joints assuming astrain at failure of 5%
for AY103 and 3% for MY750. The failure loads from the double-lap
joint tests are also shown for comparison. The shape of the curves is
similar to those obtained using the linear elastic analysis. The predicted
strength increases with overlap length up to about 15 mm and then
remains constant. However, unlike the linear elastic analysis, the
non-linear analysis has predicted the A Y103 joints to be stronger than
the MY750 joints, which is in broad agreement with the test results.
The prediction for the AY103 joints was very good but the failure
loads of the more brittle MY750 joints were over-estimated.
Figure 43 shows the predicted lap joint strengths for a 12·7 mm
overlap and various adhesive thicknesses. The non-linear analysis

t!=
25

x AY103 prediction
~20 (non-linear)
z
~
~ __ -_ x____~:q~~~~c~~J-
.... ~ (non-linear) I
li/'
x ~~f:-----
T
--i--------l-
+
I ,,"" + 1
' /~
'XT" MY750 prl2diction (linl2ar)
"1 AY103 prediction (ljnl2ar)

t
o 5 10 15 20 25 30 35 40
OVl2rlap I12n9th (mm)

FIG. 42. Double-lap Jomt strengths compared with predictions from non-
linear and linear analysis: I-e-l AYI03; f---e---j MY750 (from Adams,
Coppendale and Peppiatt, 1978a).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 63

AY103
____
25 X- prediction
(non-linear)

~i-_---- -*-
x
MY750
- prediction
~20
z (non-linear)
'" _--
_- T
1 MY750
'"'0.15 _x- ..- _ -*.- - prediction
- T _ - - - :
.
c:
--~-
(linear)

. . .1<-.~
AY103
L
- - -. •
1i1
~ 10
: l prediction
(linear)
--f 1
:§.
Co
.!!
~ 5
:E
"o
CI

o 0·1 0'2 0·3 0·4


Adheslve thickness (mm)

FIG. 43. Double-Iap joint strengths compared with predictions from non-
linear and linear analysis: I-e-; A Yl03; t- - e - -1 MY75'O (from Adams, Cop-
pendale and Peppiatt, 1978a).

predicts the same increase of strength with adhesive thickness that was
obtained with the linear elastic analysis.
Although failure is caused by tensile stresses near the adherend
corner, most of the load is transferred by the adhesive in the overlap
region. It is instructive, therefore, to examine the effect of non-linear
behaviour on the shear stress distributions in the overlap region. Tbe
predicted shear stress distribution for an AY103 lap joint just before
failure is compared with the elastic shear stress distribution in Fig. 44.
The highest shear stress near the end of the outer adherend has been
reduced to 83% of the value predicted by the elastic analysis.
Tbe prediction of the failure loads of lap joints therefore requires an
accurate knowledge of the adhesive behaviour in the highly stressed
region of the joint. It also requires an estimation of the strain that the
adhesive can withstand before failure occurs, or some other failure
criterion.
Recently, Harris and Adams (1982) have produced results for lap
joints which allow for rotation of the adherends (large displacement
analysis), and significant plastic deformation of both the adherends and
the adhesive.
The standard single-Iap joint test piece geometry has been analysed
by Harris and Adams (1983) for aluminium adherends, and the various
non-linearities included. With the material properties maintained as
64 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

1.5

1.25
r", \
C/l
C/lC/l
,, .
C/lw
wO: 1.0 --~,_._---~---
0:1- /..
I-C/l
C/lo:
o:ct /
ct W /
.
w r /
rC/l
0.75 /
.,
C/l0
w!!:! l
,
/
>...J .."
'.. --.- ---
-a.
C/la.
'I!ct
02 0.5
ctct
W
~

0.25

o 2 4 6 8 10 12
DISTANCE ALONG OVERLAP
mm

j !"mm
~SI_._
FIG. 44. Adhesive shear stress distribution in A Y103 double-Iap joint; - 0 -
elastic solution; -- X--non-linear solution (19'35 kN load) (from Adams,
Coppendale and Peppiatt, 1978a).

linearly elastic, the effect of the large displacement joint rotations on


the adhesive maximum principal stress distributions is shown in Fig.
45, for aseries of applied loads. The stress distributions have been
normalized by dividing by the average applied shear stress (applied
load divided by overlap area) and so indicate the magnitude of the
stress concentrations along the overlap. As indicated in the previous
discussion, when the applied load is increased, joint rotation occurs,
resulting in a reduction in the peak stress concentration at the edge of
the overlap, and a greater proportion of the load being transferred
within the low stress region of the overlap. The peaks in these
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 65

B C_-----~~======~======~-=======~==~

6·0

15kN
, \ 5kN 0.01 kN
~ -:...--=-=--=--~-- - - =--==:!
-1 o 234 5 6
Distance along overlap (mm)

FIG. 45. Variation of normalized maximum principal stress distribution along


the adhesive layer with applied load for large deflections.

distributions are an indication of the degree of non-linearity in the


single-lap joint test. Thus, to increase joint strength from 5-15 kN,
adhesive strength need only be increased by a factor of 2·6 rather than
3 if elastic adhesive properties are assumed.
The effects of adherend yielding were investigated by modelling the
adherends with elasto-plastic properties. The adhesive maximum prin-
cipal stress distributions are shown in Fig. 46 for a case where the
adherend properties correspond to a relatively low strength alloy (a
0·2% proof stress of 110 MPa) and the adhesive is linearly elastic. At a
very low load of 0·01 kN the distribution is identical to that in Fig. 45,
since the adherends are still elastic. Under the action of tension and
bending, the adherends begin to yield at an applied load of approxi-
mately 1·5 kN. At 3 kN, the adherend plastic deformation has had two
effects on the adhesive stresses. Firstly, it has led to a reduction in the
peak stress concentration at the end of the overlap, at point A, over
and above that for the elastic case, as a result of the enhancement of
66 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

the joint rotation. Secondly, the concentration of adhesive stress at the


edge of the spew, point B, has increased, due to maximum adherend
deformation occurring adjacent to this point producing a localized
increase in the differential shear effect. At an applied load of 6 kN,
further plastic deformation in the adherends has taken place, the peak
at point A is further reduced but, more significantly, the peak at B has
dramatically increased, and the stresses at this point are now the
highest in the adhesive. Furthermore, the concentration of stress at
point B is greater than that at point A in the elastic case in Fig. 45 for
the same applied load. It may be concluded, therefore, that when
adherend plastic deformation takes place, then joint strength will be
reduced and, at the same time, failure will no longer initiate from point
A, but from point B. The two types of failure observed in practice,
designated as types land 11, are illustrated in Fig. 47; type 11 being
indicative of adherend plastic deformation. For a toughened adhesive,
joint strength was found to decrease from 16-8 kN, when the high

Adherend

Adhesive

10·0

8·0
I

~
c Cj6·0 \
>

4.0\\
0....
'i:
'b
nJ
\\ \

~ ,
2·0

-1 o 2 3 4 5 6
Distance along overlap (mm)

FIG.46. Normalized maximum principal stress distributions along the adhesive


layer with adherend yielding, at various applied loads.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 67

Type I mode of fracture

Type Ir mode of fracture

FIG.47. Modes of fracture in lap joints.

strength adherends were replaced by low strength adherends as model-


Ied in the analysis.
Being able to include the geometrical non-linearities in the analysis
of the single-Iap joint, as weIl as the material non-linearities, enables
elasto-plastic adhesive properties to be included and predictions of
joint strength to be made in a similar fashion to that already described
for double-Iap joints. In Fig. 48, the uniaxial tensile stress-strain curve
for two extremes of adhesive behaviour are illustrated. The high
strength, low ductility response corresponds to an unmodified epoxy
(MY750), whilst the lower strength, high ductility response corres-
ponds to the same basic epoxy toughened by the addition of 15 pphw
of CTBN (carboxyl-terminated butadiene-acrylonitrile). By including
these properties in the analysis, single-Iap joints made using these
adhesives, with a variety of adherends were modeIled and, by applying
appropriate failure criteria, the joint strength predicted. For the brittle
68 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

100

x
80

tU y
c.. 60
l:
<I)
<I)
~

.!:; 40
V)

o 0·04 0·06 0·08 0·10 0·12 0·14 0·16


Strain

FIG. 48. Bulk uniaxial tensile stress-strain curves for adhesives X (MY 750)
and Y (CfBN).
I
16
x
14

,/
,/
,- ....
.... -- ___ --x

12 TI /
x/
/
/
~10
/
/
z /
-'"
TI /
I

~_
n /,- -----------r:;:..
xIV
4 ,x/

I I
o 100 200 300
AdhlZrlZnd yilZld strlZngth (MN m- 2)

FIG. 49. Variation of joint strength with adherend yield strength: --0-
experimental; - - - x - - - finite-element predictions.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 69

adhesive, a maximum principal stress criterion was best, based on the


uniaxial tensile response, whilst for the toughened adhesive, a max-
imum principal strain criterion was used. Figure 49 shows good corres-
pondence between the measured and predicted strengths for a range of
adherends using these criteria. Note also the change in failure mode,
when lower strength adherends are used, from initiation within the
spew to initiation at the edge of the spew, i.e. transition from type I to
type 11 failure.
Clearly, there is a complete range of adhesive behaviour in between
the two extremes included here, where the failure criterion may not be
based on a single parameter such as maximum principal stress or
strain, as conditions in the other principal directions mayaiso be
important. Again, a joint may undergo a variety of loading conditions,
resulting in a range of critical conditions in the adhesive. Further
investigation is required into failure criteria for adhesives in order that
the limits. to the strength of simple lap joints, as weH as more complex
bonded structures, may be determined under a variety of loads.

THE EFFECT OF ADHEREND SHAPE-SCARFED,


BEVELLED AND STEPPED ADHERENDS

So far, we have only considered parallel-sided adherends in single- and


double-lap joints. It has been shown that the mathematical treatment,
whether it is by closed-form analytical methods or by finite-element
techniques, is difficult if realistic results are to be obtained. For
instance, it is essential to allow for adhesive and adherend plasticity if
joint strength predictions are to be made. But, as shown in Fig. 5,
there are several forms of the lap joint in which the adherends are not
parallel-sided, constant-thickness sheets but can have a variety of
forms. These deviants are an attempt to reduce the high stress and
strain concentrations, which occur at the ends of the simple lap joint,
by modifying the stiffness of the adherends. In these profiled joints, the
load line direction must change and, in addition to the tensile stiffness,
so the shear and bending stiffness of the adherends change.
While profiled adherends can be treated by closed-form techniques,
the mathematics is necessarily more complicated than for parallel-
sided joints. There have been many notable contributions to this field.
Hart-Smith has produced several analyses for tapered and stepped lap
joints (e.g. 1973b), in which the adhesive is considered as elasto-
plastic. Lubkin (1957), Wah (1976) and Webber (1981) have analysed
70 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

scarf joints. Thamm (1976) considered scarf and bevel joints in which
the adherends were not tapered fully to an edge since this is very
difficult to do in practice. He conduded that tapering was of little value
unless a fine edge could be obtained. Even an edge reduced to 10% of
the original thickness provides litde advantage over a parallellap joint
and makes the additional mechanical operation of questionable value.
However, when using the finite-element technique, there is no
problem in dealing with the complexities of tapered and stepped joints
such as face the dosed-form analytical mathematician. Various authors
have tackled the problem. Among these, Barker and Hatt (1973) used
a rather coarse mesh but with a special element to represent the
adhesive. Wright (1978, 1980) used tri angular elements and an elastic-
plastic adhesive. We present in some detail below results obtained by
the research team led by one of the authors (R. D. Adams) for a
variety of joints which are not of constant thickness and which were
analysed using the finite-element technique. (Other results of this work
are given later in the section dealing with adherends made from
composite materials.)
Figure 50 shows the adhesive shear stress distributions for a double-
scarf joint and a double butt-strap joint and compares them with
ordinary single- and double-Iap joints. The double butt-strap joint
gives the lowest stress concentration at its middle, but is then identical
to the ordinary, parallel double-Iap joint at the other end. For this
reason, it is suggested that the straps should be bevelled or scarfed so
that the stress concentration might approach that of the double-scarf
joint.
Essentially, scarfing helps to reduce the stress concentrations in
joints but, as Thamm has said, it has to be done fully to a knife edge if
it is to be effective. It is impossible to illustrate all the variables
possible in a tapered joint; any further detail would be unnecessarily
specific.
However, there is one further dass of joint which should be consi-
dered, the stepped lap joint. Some results for an elastic analysis are
given here and the reader is again referred to the section on bonding
composites. Figure 51 shows two results for apparently identical
adherends. The difference is the longitudinal spacing between the
steps. For a gap of 0·5 mm, the shear stresses are much higher than for
the case when the gap is only 0·15 mm. The reason for this is that, in
the latter case about half of the load is carried direcdy between the
butt faces. At first sight, this seems to be an ideal situation. However,
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 71

1\

hL_ DOUBLE- BUTT-STRAP JOINT


l
J l

I
I
1I \
:~ 2.5 \
I
VI. "'

~ :"'
X

.... \ I
~ 20

I
V'
a: ..

~~ \
-I "~ II

1.0
V
/
V'
\
\
\ DOUBLE-SC,t.RF JOINT
\
\
'-'-'-'-'-'
\
\
05

,
,, -----
SINGLE - L,t.P JOINT

DOUBLE- , DOUBLE-LAP
BUTT-STRAP JOINT
JOnllT

0 5 10 15 20 25
PISTANCE "LONG OYERLAP
mm

FIG. 50. Adhesive shear stress distributions in aluminium-aluminium joints


(from Adams, 1981).
72 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

~j
2-0
---d=0-5mm
'"'" - - - d=0-15mm
"
L
t) 1·5
L

""
.c
'" lOH--+-------/--J-,I.c-\-------!---,.k-{--I.------+-+
lJ
~
(i 1
g- 1

8,05 1
" I
t I
~ I
I
o 5 10 15 2D 25
Distance along overlap (mm)

FIG. 51. Adhesive shear stress distributions m step joints; aluminium


adherends.

the truth is far less comfortable, for this adhesive must therefore be
carrying a local tensile stress, even ignoring stress concentrations, of
the order of half that of the aluminium. Since the adhesive can only
carry a tensile stress of say 60 MPa, while the aluminium may be able
to carry 300-400 MPa, the joint efficiency is clearly lirnited. In fact, a
close-butted stepped lap joint tends to approach the case of an
ordinary butt joint rather than that of the scarf joint. Thus, care should
be taken to ensure that the butt faces be separated by at least 0·5 mm.
If non-linear adhesive behaviour is considered, then yie1d will occur
first at the ends (A and D in Fig. 51) and then at the intermediate steps
(B, C) before it spreads throughout the whole glue-line.
It has also been proposed that the adherends of a lap joint should be
profiled so that the adhesive thickness may be varied along the length
of the overlap while leaving the adherend thickness essentially con-
stant. Adams et ai. (1973) showed that by using a quadratic profile, the
adhesive shear stress may be made approximately constant along the
lap. However, for practical joints, the maximum amount of profiling is
less than 0·1 mm, which is difficult to achieve in practice. A simpler,
linear profile is possible, but again the problems of precision machining
to such fine tolerances outweigh the likely benefits, particularly as
further precision in joint alignment is required.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 73

COMPOSITE MATERIALS

One of the major advantages of adhesive bonding is that it enables


dissimilar materials to be joined, even when one of these is non-
metallic. A major application of bonding is therefore where composite
materials are concerned. Composites take many forms. In aerospace
applications, these materials usually consist of highly aligned layers of
carbon or glass fibres, each oriented to aceommodate the expeeted
loads. Some aerospaee eomposites are woven or stitehed so that the
fibres are not perfeetly aligned. High-quality chemieal plant may be
made from satin-weave glass fibre reinforeed polyester or epoxy resin,
while lower grade eomposites usually consist of random glass fibres in
polyester. Thus, eomposites ean be highly anisotropie in respect of
both stiffness and strength and, although a unidireetional eomposite
may be very strong and stiff in the fibre direetion, its transverse and
shear properties may be very much poorer. Bolts and rivets ean
sometimes be used with eomposites, but it is then often neeessary to
have load-spreading inserts bonded into the strueture. Adhesive bond-
ing is attraetive sinee it allows for a more gentle diffusion of the load
into the strueture, thus reducing the loealized stresses eneountered in
the use of bolts and rivets. A further form of bonding has been finding
converts reeently: this is the so-ealled eo-curing teehnique in whieh the
eomposite is bonded without the use of an adhesive. To do this, the
eomposite is prepared in its uneured (pre-impregnated fibre) form and
heat and pressure applied. As the eomposite eures, the exeess matrix
material (usually a high-performance laminating epoxy) is squeezed
out in a liquid form and eontaets the adjaeent component. When
curing is eomplete, the bond is made with due saving in production
costs.
The teehniques of analysis are essentially the same as when isotropie
adherends are used, although due attention must be paid to the low
longitudinal shear stiffness of unidireetional eomposites. As Demarkles
(1955) showed, even with metallic adherends in whieh the shear
modulus is of the order of 25-30% of Young's modulus, it is neeessary
to take aeeount of adherend shears. With unidirectional eomposites,
this modulus ratio may be as low as 2%, and so the adherend shears
beeome extremely important. The use of lamination techniques in
whieh fibres are plaeed at different angles to the plate axis leads to
redueed longitudinal and inereased shear moduli. However, the trans-
verse modulus (i.e. through the thiekness of the adherend) remains
74 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

low, being only two or three times that of the matrix material (usually
epoxy 01' polyester resin). In addition, the transverse strength is low,
usually being of the same order or less than that of the matrix. The
composite matrix has to meet a variety of requirements, only one of
which is strength. Also, it will have been formulated from the same
basic family of materials as is the adhesive, which has been largely
chosen for its strength and ductility. Thus, if the joint experiences
transverse (peei) loading, there is a strong likelihood that the compo-
site will fail in transverse tension before the adhesive falls. Adhesive
peel stresses should therefore be minimized where composite
adherends are used lest this leads to adherend failure.
Matthews et al. (1982) have recently published an up-to-date review
on the strength of joints in fibre reinforced plastics. They considered
both theoretical and experimental results, although mainly the former.
The reader is referred to this excellent article. In essence, they con-
cluded that it is necessary to consider non-linear adhesive behaviour if
joint strength is to be predicted. Joint strength is improved as the
adherend stiffness increases and the adhesive stiffness decreases, and a
ductile adhesive is always preferable to a brittle one.
It may be useful to the reader to consider as an example the work by
Adams et al. (1978c) and some of their hitherto unpublished results.
They used finite-element methods to examine the stresses in high-
performance composites in symmetricallap joints with parallel, bevel-
led, scarfed and stepped adherends. The composite adherends were
assumed to be linearly elastic type II carbon fibre reinforced epoxy
composites with a 60% fibre volume fraction. The mechanical proper-
ties of this material are given in Table 3.
The adhesive was treated as elastic-plastic with a paraboloidal yield
criterion as described earlier in this chapter. Its basic mechanical
properties are given in Table 4.

TABLE 3
MECHANICAL PROPERTIES OF CFRP ADHERENDS

Longitudinal Young's modulus (EI) 135GN.m-2


Transverse Young's modulus (E,) 7GN.m- 2
Interlaminar (longitudinal) shear modulus (Glt) 4·5GN.m- 2
Longitudinal tensile strength 1550MN.m-2
Transverse tensile strength 50MN.m-2
Interlaminar shear strength llOMN.m- 2
Longitudinal Poisson's ratio (VI') 0·3
Transverse Poisson's ratio (vtt ) 0·3
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 75

TABLE 4
MECHANICAL PROPERTIES OF EPOXY ADHESIVE

Young's modulus 2·8GN.m- z


Poisson's ratio 0·4
Uniaxial tensile yie1d stress (YT ) 6SMN.m- z
Uniaxial compressive yield stress (Yd 84·SMN.m-2

Stress distributions in the adhesive and the adherend were then


produced assurning timt:
(1) the adhesive was perfectly elastic;
(2) the maximum adhesive plastic strain was 10%;
(3) the maximum adhesive plastic strain was 20%.
The adhesive elastic shear stress distributions for single-Iap, double-
lap, double butt-strap and double-scarf joints between unidirectional
type 11 CFRP adherends are shown in Fig. 52. The adhesive stress
concentrations are less than those for joints between aluminium
adherends of the same geometry because of the greater longitudinal
stiffness of the CFRP adherends. However, as the tensile strength of
the CFRP is more than 4 times the yield strength of the aluminium
alloy, the joint* effidendes predicted by the linear elastic analyses are
much lower than those of the aluminium to aluminium joints. The
stress concentration at the tension end of both the double-lap joint and
the double butt-strap joint are the same, although in the double
butt-strap joint, the highest stresses are predicted to occur in the
adhesive at the butt face. As with the aluminium to aluminium joints,
the effect of scarfing the adherends is to reduce the stress concentra-
tion and to increase the joint efficiency. The stress concentration is
reduced to 77% of the parallellap joint value in the case of the single
lap. Comparison with the results for aluminium adherends suggests
that there is less benefit to be obtained from scarfing unidirectional
composite adherends than there is for isotropic materials such as
metals.
As fibre reinforced plastics have a comparatively low strength epoxy
resin matrix, adherend failure is more likely in composite joints than in
metal to metal joints. There are three possible modes of failure in the
composite:
(i) tensile failure in the fibre direction;

* Defined as the ratio of the joint strength to that of the weakest adherend.
DOUBLE - LAP JOINT

DOUBLE BUTT - STRAP JOINT


L
~/::::::::~~~~~~~;:;;~~~::::::~D~O~U~B~LfE-~.~S~CA~R~F~J~O~IN~T~==~.~~~ .
4.0

3.5

3.0

I~
>. 2.5 \ SINGLE - LAP JOINT

""~
,W
",x
",,,,
w
~O
"'!!I
.J
a:n.
..:n.
w..:
x
"'w
<!)
..:
a:
w
~
1.5

._. __ ._--.
DOUBLE- SCARF JOINT_._ V·
I
I
" ,
,
)
0.5 ,
/
DOUBLE
BUTT - STRAP ------
JOINT

o 5 10 15 20 25
DISTANCE ALONG OVERLAP
mm

FIG. 52. Adhesive shear stress distributions in CFRP-CFRP joints.


THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 77

(ii) tensile failure perpendicular to the fibre direction;


(iii) interlaminar shear failure.
The strengths in these three modes were given above for a type II-S
(treated) unidirectional carbon fibre composite. Rotem and Hashin
(1975) suggested that matrix failure depended on the combined effect
of the tensile stress perpendicular to the fibre direction and the

r r~
interlaminar shear stress and that failure occurs when:

(:~ + ((T;2 1
where (T22 = tensile stress perpendicular to fibre direction, (TT =
transverse tensile strength of composite, (T12 = interlaminar shear stress
and T = interlaminar shear strength of composite.
The joint efficiencies calculated from each of these separate failure
modes are given in Table 5 for a double-lap joint. On the basis of a
linear elastic analysis, it appears that adhesive failure is most likely to
occur before failure in the adherends.
The tensile stress «(Tx) distributions in the CFRP adherends of a
double-lap joint are shown in Fig. 53. The highest stress occurs at the
surface of the centre adherend and is 1· 3 times the stress in the
adherend away from the joint, compared with 1·08 times in an
aluminium to aluminium double-lap joint. This is because the compo-
site has a lower shear stiffness than the aluminium. The maximum
TABLE 5
ELASTIC JOINT EFFIClENCIES FOR ADHESIVE AND ADHEREND FAILURE MODES IN
CFRP-CFRP DOUBLE-LAP JOINT
Mode and loeus of failure Joint
effieieney
(%)

Tensile in adhesive at corner of outer adherends 16


Tensile in fibre direction of composite (centre
adherend) 77
Tensile transverse to fibre direction in centre
adherend at tension end 55
Interlaminar shear in centre adherend 81
Combined transverse tensile and interlaminar
shear (matrix failure) in centre adherend 46
Combined transverse tensile and interlaminar
shear (matrix failure) at corners of
outer adherends 41
-....)
00
A B
r ---- I ---7I -----. ------- I'\.
L Cf.-- '\J \
, \ . l ~
C D
~

1.5
i
~
::r:
tn
CI)
I:)'
VI '
, -, , ~
VI ......
w o
«VI
~~ 101-· - - "
.__ .--t;\ - '\"A-- - - - - ,I ~
w « ~
-' I-
- VI
VI ", , Z
Z
wO
\ ,//
I-
W
:::; \ B
......... _________ _ ~
o a.
~._._._._.
z a. 0.5 ----::

« ~
'l!o
--
---::-:'::--'_.'--')
- ... - - - tn
:>;:I
« // ---- -...... ....... Z
Cl
/' ... ...
I ! , • !
01 I "..........-
... 'J ,
25 30 35 40 45 50 55 60
DISTANCE ALONG JOINT
mm
FrG. 53. Adherend tensile stress distributions in CFRP-CFRP double-Iap joint (frorn Adams, 1980).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 79

tensile stress in the outer adherends is in the outer fibres because of


bending effects. In a scarf joint, the maximum adherend tensile stress
(J'x) occurs at the same place but is only 1·11 times the average.
If we now allow for non-linear adhesive behaviour, the high adhe-
sive stress concentrations predicted by the linear elastic analysis will be
relieved to some extent. Figure 54 shows the predicted spread of the
yield zone of adhesive at the tension end of a double-Iap joint as the
load is increased. As would be expected, plastic flow begins near the
adherend corner and the load corresponds to a joint efficiency of 21 %.
Each subsequent load increment represents an increase in joint effi-
ciency of 4·4%. When elastic perfectly-plastic behaviour is assumed
for the adhesive, a maximum strain criterion for failure seems appro-
priate. In Fig. 55 the joint efficiency is plotted against the maximum
principal strain in the adhesive at each end of a double-Iap joint.
Assuming a failure strain for the adhesive of 5%, the analysis predicts
a joint efficiency of 31% for a double-lap joint compared with 16%
predicted by the linear elastic analysis. Similarly, the non-linear
analysis predicts an efficiency of 39% for the double-scarf joint com-
pared with 20% predicted by the linear elastic analysis. Although the
predicted efficiencies are almost doubled by allowing for non-linear
behaviour in the adhesive, failure in the adhesive is still predicted to be
more probable than failure in the adherends (Table 5).
Although the adhesive fails in tension at the end of the joint, most of
the load is transferred in the overlap region. It is, therefore, instructive
to examine the effect of non-linear behaviour on the adhesive shear
stress distribution. The shear stress distributions corresponding to
maximum adhesive strains of 10% and 20% in a double-lap joint are
shown in Fig. 56(a). The shear stress distributions are only modified
significantly at high values of maximum adhesive strain, by which time
the joint is likely to have failed. The adherend (J'x stress concentrations
are reduced, but not significantly, by plastic flow in the adhesive. For a
double-scarf joint, the elastic shear stress distributions are similarly
modified by yielding of the adhesive (see Fig. 56(b)).
The elastic shear stress distribution for a double-lap joint between
0/90/90/0 crossply adherends is compared with the shear stress dis-
tribution for unidirectional adherends in Fig. 57. The higher shear
stresses at each end are caused by the lower tensile stiffness of the
crossply adherends. This also produces a higher stress concentration of
10·1 compared with 7·3 for a similar joint between unidirectional type
II CFRP adherends.
OUTER ADHEREND

7 6 5 4 3
4

CENTRE ADHEREND

~ ~
. ~ 5'6 (i.
53 54 55
DISTANCE ALONG JOINT
mm

FIG. 54. Incremental zones of plastic deformation at tension end of CFRP-CFRP double-Iap joint (from Adams and
Peppiatt, 1977a).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 81

70

60

50

_ ...

~..
40
\'!Z
...
... u
'" 30
...;!!;IL
",

...,

10

Q3
... DHESIVE M...XIM\JM PRINCIP...L S TR... ,N

FIG. 55. Joint efficiency plotted against adhesive maximum principal strain in
CFRP-CFRP double-lap joint (from Adams, 1981).

It is not uncommon for composites to be bonded to metals and there


are one or two important points to be brought out. Figure 58 shows
the adhesive shear stress distributions for double-lap joints between
aluminium and unidirectional type II CFRP adherends. If the outer
adherends are aluminium and the centre adherend is CFRP, the
highest shear stress occurs at the compression end of the joint where
the aluminium adherends are loaded. This is because the aluminium
adherends have a lower tensile stiffness than the composite adherends.
However, the adhesive stress concentrations at each end of the joint
are similar (Table 6). Therefore, as far as the adhesive is concerned,
the joint is weH conditioned. Alternatively, if the outer adherends are
unidirectional CFRP and the centre adherend is aluminium, then the
higher shear stress that occurs at the tension end of a joint with similar
adherends is increased still further by the adherend dissimilarity. The
stress concentration at the tension end of the joint is now 3·6 times the
stress concentration at the compression end, with the result that the
joint efficiency, in terms of the strength of the aluminium aHoy
adherends, is reduced from 79% to 48%.
82 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

- - - Elastic solution
2·5 ------Elastlc-plastic inc. nO.5 (10"lomax.
adhesive strain)
- - -Elastic-plastic inc. no. 8 (20"10 max.
adhezsivez strain)

11)
11)

>.
" 1-5
'-
.....
"
11)

.t! '-
cu
'" ~"
"'-'" u'"
tl .S! 1'0~---+-lr----------i-+I---

!! ~
~0'5
~.
::::----- - /Y'
F/
,,'

(a)
o 5 10 15 20 25
Distancez along ovezrlap (mm)

2.6 Adhezsive shear, 'xy, stress distributions


24 - - - Elastic solution
------ Elastic-plastic inc. nO.6 (10·/. max
2·2 adhesive strain)
-- - Elastic-plastic inc. nO.10 (20·/. max
2·0 adhesive strain)
11) 1·8
11) 11)

~ ~ 1-6
1:;1;) :
11) '- 1-4
L cu •
~ ~ 1·2 ~
~ 11) -

~ lOll---~~----------+'---4--
"
11)

> .--
'iii 2:0·8
" cu
~ eO,6
'0 cu
<l: L:
" 04

0,2 (b)

o 2 4 6 8 10 12 14 16 18 20 22 24
Distance along overlap (mm)

Fm_ 56_ Adhesive shear stress distributions in CFRP-CFRP joints: (a)


Double-lap joint; (b) scarf joint (from Adams and Peppiatt, 1977a).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 83

\J --_.~-.ct

3.5

3.0

I
I

o 5 10 15 20 25
DISTANCE ALONG OVERLAP
mm

FIG. 57. Adhesive shear stress distribution in crossply CFRP-crossply CFRP


double-lap joint.
84 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

OUTER AOHERENO

SJ CENTRE AOHERENO
- - - - -----'--li

4.0

3.5

3.0

I~
I
2.5 ALUMINIUM OUTER; I
CFRP CENT RE I
'"
iJ«
0: I
.w I
",:r I
"''''
w I
~O t'1 I
I
'" w:J
2.0
0: Q.
1 CFRP OUTER; I
« Q. 1
ALUMINIUM CENTRE ------.,'
w« 1
l:
"'w 1
Cl 1
«0:
1 I
w 1
«> l5 1
I

\ I
\ I
.. I
1 I
I
1.0
I
I

\
\
0.5 \
\
\
,,
,,
... ... ~
~

0 5 10 15 20 25
OISTANCE ALONG OVERLAP
mm

FIG. 58. Adhesive shear stress distributions in aluminium-CFRP double-Iap


joints.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 85

TABLE 6
ELASTIC STRESS CONCENTRATIONS AND JOINT EFFICIENCIES FOR ALUMINIUM-
CFRP JOINTS

Overlap Adhesive stress Joint


Joint length Outer Centre concentration efficiency
type (mm) adherend adherend (%)
End where End where
outer centre
adherend adherend
is loaded is loaded

0'9mm 1·8mm
Double lap 25·0 7·0 6·9 79
aluminium CFRP
0'9mm 1·8mm
Double lap 25·0 3·2 11·5 48
CFRP aluminium
1·5mm 0·75mm
Double lap 25·0 5·0 13-4 28
aluminium CFRP
Double 0'9mm
1·8mm
butt 25·0 aluminium 13·8* 6·9 41
CFRP
strap (straps)
0'9mm 0'9mm
Single lap 25·0 13-8 11·25
aluminium CFRP
0'9mm 1'8mm
Double searf 25·0 7·6 4·0 74
aluminium CFRP
0'9mm 0'9mm
Single searf 25·0 10·2 5·2
aluminium CFRP

* Between butt faees


N.B. Joint efficieney is based on strength of aluminium alIoy.

The joints described above do not attempt to use the high strength
of the unidirectional type 11 CFRP. If the tensile strength of the CFRP
is assumed to be 1550 MPa and the yield strength of the aluminium
alloy is assumed to be 325 MPa then, if we make full use of the higher
strength of the CFRP, the total thickness of the aluminium alloy
adherends would need to be approximately 4 times that of the compo-
site (allowing for a safety factor of 1·2 for the strength of the more
brittle CFRP). The adhesive shear stress distribution for a double-lap
joint with two 1· 5 mm thick aluminium outer adherends bonded to a
0·75 mm thick CFRP centre adherend is shown in Fig. 59. The lower
tensile stiffness of the thinner centre adherend increases the shear
stress and the stress concentration at the tension end of the joint. The
joint efficiency, based on the strength of the aluminium adherends, is
now only 20% assuming linear elastic failure of the adhesive at a
tensile stress of 65 MPa (Table 6).
86 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

El<t. SI
ALUMINIUM

\
CFRP

4.0

3.5

3.0
Matched adherends
(as shown abov/l)

j!
\
\
\
2.5 \
\

x" \
Equal thlckness adhlZrllnds
IJW
",,,,
.:1: \.---- (From Fig. 58)

'"
\:!o
..
... w
"':::; I

Ir" I

" "w
W
:l:
I

"'<!l

"
0:
W
I
~ I
I

\
\
\
\
\
\ I

0.5
, I
I
\
/
/
/
.,. "

o 5 10 15 20 25
DISTANCE ALONG OVERLAP
mm

FIG. 59. Adhesive shear stress distribution in aluminium-CFRP double-Iap


joint (matched adherends).
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 87

ALUMINIUM
f\.
~ '\J CFRP
<t
DOUBLE LAP JOINT

I ALUMINIUM CFRP
~ Si ~
DOUBLE SCARF JOINT

3.0

\
\
\
\
2.5
\

\
\
\

I~
\
\
.' 2.0 \
"",

""
\ Double-Iap Jomt
_____ IFrom Fig 58)
.~
:::'"
w
\
"'0
~ j 1.5 \
",n. I
\ DOublll scarf JOint
~~ I
M

:< I
"'wC) \ I
\
~
W
~ 1.0 ~------",,-------------------

I
I
I
I

0.5 \ I
\
,I
\
, /

"
/

------
o 5 10 15 20 25
DISTANCE ALONG OVERLAP

FIG. 60. Adhesive shear stress distribution In aluminium-CFRP double-scarf


joint.
88 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

___ d~0.15mm}
20 (\ _________ d~0.5mm CFRP-CFRP adherends

~
_._.- d~0'5mm, alumin,um·CFRP adherends
!\
.btI'I 1.5 I" '\
.J \ "
1\

tI'I~
u'l..c
:
I
\
\. J
/ \
I
~ In I \ I t
~ ~ 1>O'"",---1c-\------+-1-+-'r,--------+-++-\'r--------r--+-
~~ /
~~
lJ1 8,0'5
1/
''-e;"
~
10 15 20 25
Distance along overlap (mm)

FIG. 61. Adhesive shear stress distributions in step joints.

The effect of scarfing the adherends (Fig. 60) is to unbalance the


stress concentrations to such an extent that they are much higher at the
compression end than at the tension end of the joint. Although the
maximum shear stress is about 7% less than in a double-Iap joint, the
adhesive fillets of the scarf joint are much smaller, giving a 9% higher
stress concentration (Table 6).
Stepped lap joints are an important method of joining CFRP com-
ponents. The adhesive shear stress distributions in 4-step joints be-
tween unidirectional CFRP adherends are shown in Fig. 61. For the
small butt spacing (0·15 mm), the maximum shear stress at the ends of
the steps is less than the average applied shear stress. This is because at
least half the applied load is transferred by the adhesive between the
butt faces giving stress concentrations in these regions of approxi-
mately 10. The proportion of the load transferred by shear, rather than
by the butt faces, is increased considerably by increasing the thickness
of the glue-line between the butt faces to 0·5 mm. Increasing the butt
face glue-line thickness increases the shear stresses at the ends of the
steps but reduces the stress concentrations between the butt faces. The
predicted joint efficiency is increased from 6-10% in the case of
unidirectional type 11 CFRP adherends. We also show in Fig. 61 the
shear stress distribution in a step joint between aluminium and uni-
directional CFRP. The highest shear stress and stress concentration
ELAST1C 8EHAVIOUR UP TO JOINT EFFICIENCY QF 10.2 °/0

INCREMENT 1; JOINT EFFICIENCY 12.8 0f,.


~

~ !~ 6l
~
~
INCREMENT 2; JOINT EFFICIENCY 15.3 °/0
a::
~
Cl
Z

r'" W4 ~ fIIf@ ( ~!"TI


o"Tl
INCREMENT 3; JOINT EFFIC!ENCY 17.8 0/0
~
!"TI
i Wb."",,, 1m". " ~ '"
!"TI
'"
~?! 1?">?2>?211@ '"
Z
~
t:I
INCREMENT 4; JOINT EFFICIENCY = 20.4 °/0
::r:
!"TI
'"
~
"m ". " "''''' '" fl!l ""''''''" """"""fill" mm= wa". """ .,,, ~ ....
L"""" i o
~
FIG. 62. Incremental zones of plastic deformation in CFRP-CFRP step joint. '"
00
\0
90 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

occur at the end where the lower stiffness aluminium adherend is


loaded.
When elastic-plastic behaviour is assumed for the adhesive, yield is
predicted to occur first between the butt faces at each end of the step
joint. Figure 62 shows the spread of the yield zones as the load on the
joint is increased. At higher loads, the proportion of the load transfer-
red by the butt faces is reduced. If a maximum strain failure criterion is
assumed for the adhesive, the load carrying capacity of the joint is
increased by increasing the butt face glue-line thickness. For example,
assuming a 5% failure strain, the non-linear analysis predicts joint
efficiencies of 9 and 17% for the two glue-line thicknesses compared
with efficiencies of 6 and 10% predicted by the linear elastic analysis.
Higher strain adhesives give correspondingly higher joint efficiencies.

TUBULAR JOINTS

Adhesive bonding provides a convenient and light method of assembl-


ing structures consisting of thin-walled tubes. Typical joints in such
structures are the tubular lap joint shown in Fig. 5 and the very similar
tubular scarf joint.
There is less literature dealing with the stresses in lap joints between
thin-walled tubes than there is concerning lap joints between flat plates
although, as in the case ofaxial loading, the stress concentrations arise
by the same three mechanisms, i.e.:
(i) differential straining;
(ii) bending introduced by the non-colinearity of the overlapping
tubes;
(iii) end effects.
In torsion, there are no bending effects and only differential straining
and end effects need be considered.
The system ofaxes used here is such that z represents the longitudi-
nal direction, r the radial direction, and e the hoop direction (cf. Fig.
63 where a similar system is used for the butt joint).
Lubkin and Reissner (1956) have analysed the stresses in tubular lap
joints under a tensile axial load and give solutions for both the shear
stresses, Tzn in the adhesive layer, and the normal stresses, an across
the thickness of the adhesive layer which are due to adherend bending.
Their analytical method assumes that the adhesive can be approxi-
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 91

Spew fillet

Adhesive Adherend

(a)

Adherend

Adhesive

Spew fillet
(radius. a)

(b)

FIG. 63. Circular butt joints: (a) annular; (b) solid (frorn Adams et al., 1978b).

mated to an infinite number of tensile and shear springs, and that the
work of the stresses Tzr and Ur in the adherends can be neglected in
comparison with the work of these stresses in the adhesive. They
present their results in a tabular form for 48 joints with different
geometries and (always linearly elastic) material properties. These
results show that T zr and Ur are a maximum at the end of the adhesive
layer. However, because of the free surface at the end of the adhesive
layer, Tzr must be zero here. There should, therefore, be a high shear
stress gradient near the end of the joint, as the shear stress increases
from zero on the free surface to some maximum value in a very short
length. Because of stress equilibrium considerations, this high shear
stress gradient is associated with a normal stress gradient across the
thickness of the adhesive layer (Adams and Peppiatt (1974».
92 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Volkersen (1965) has given a closed-form solution for the shear


stresses, T r !), in tubular lap joints acted on by a torque but he assumes,
for the purposes of analysis, that the tubes are of the same diameter.
The simplification is unnecessary and a closed-form theory which
overcomes this limitation is given by Adams and Peppiatt (1977). As
there is no free surface in the hoop direction, the shear stress, T r !), is a
maximum at the ends of the joint.
Stresses in tubular lap joints under other loading conditions have
also been investigated. Terekhova and Skoryi (1973) give a closed-
form solution for the stresses in tubular lap joints under external and
internal pressures which neglects the effects of adherend bending.
Kukovyakin and Skoryi (1972) set up differential equations for the
stresses in tubular lap joints, acted on by a system ofaxisymmetric
moments and forces, and which allow for the effects of adherend
bending. However, as they give results for thick-walled tubes, they
consider bending effects to be negligible, and so the equations are
simplified to neglect bending. Alwar and Nagaraja (1976b) described a
finite-element analysis of tubular lap joints in tension, which allows for
the viscoelastic behaviour of the adhesive. However, they give only a
brief mention of the elastic case and do not consider the practically
important region of the adhesive fillet, wherein the maximum stresses
occur.
Adams and Peppiatt (1977) used the finite-element technique to
determine the stresses in tubular lap joints loaded in tension or torsion
and were able to obtain realistic results for stress concentrations by
allowing for the stress-relieving effect of the fillet. They showed that, in
the axial load case, the stress concentrations predicted from the
finite-element models with the fillet have been shown to be greater
than those predicted by the Lubkin and Reissner theory. This is
because the closed-form solution does not evaluate the true stress
concentrations, i.e. those caused by end effects. The influence of the
adhesive fillet on the stress concentrations in the torsion al case is
shown to be less significant, as the stress concentration values from the
closed-form theory are of similar size to those predicted by the
finite-element models. They also investigated the effects of partial
tapering of the adherends to form a scarf joint. It was concluded that
the reductions in stress concentration obtained with this form of joint
do not make its manufacture for this reason alone worthwhile, and in
the axial case the reductions in the stress concentration were not found
to be significant.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 93

BUIT JOINTS

Butt joints, usually axisymmetric but also with square or rectangular


adherends, are widely used as specimens for testing the response of
adhesives to shear, tensile and compressive stresses. It is therefore
important to be able to interpret the experimental results in the light of
the stress distribution in the joints. Butt joints are not usually encoun-
tered in practical load bearing situations since, although strong in
tension, compression, and shear, they are readily fractured by a modest
bending moment owing to the high peel or cleavage stresses so
produced.
The major importance of butt joints is that they provide an appar-
ently convenient means of determining the mechanical properties of
structural adhesives. Various techniques are available for analysing the
stresses and strains. For the elastic region, a simple closed-form
analysis is sufficient, although this is less easy to apply when yield
occurs. The advantage of using butt joints is that the adhesive is tested
in the thin-film form as used in most joints, thus overcoming any
possible objection to bulk specimens (some of which are difficult to
make since the adhesives can react exothermally). On the face of it, the
stress distribution is simple. However, the ogre is once again the
problem of end effects. If joints are to be loaded to failure and if the
failure stress is to mean anything, then it must be the true stress and
not a convenient but misleading approximation.
De Bruyne (1951) suggested the use of an annular butt joint (Fig.
63(a)) loaded in torsion to measure the response of an adhesive to a
pure shear stress. The annular butt joint or 'napkin ring' specimen
minimizes the variations of shear stress in the adhesive and has been
used by Foulkes et al. (1970), Bryant and Dukes (1964), Humpidge
and Taylor (1967) and McCarvill and Bell (1974), for measuring the
shear strength of adhesives. Kuenzi and Stevens (1963) and Bossler et
al. (1968) used this configuration for obtaining shear stress-strain
curves. The shear stress distribution ca1culated from simple elasticity
theory is:
2Tr
(20)

where T z 9 is the shear stress at a radius, r, caused by an applied torque,


T, while rj and r0 are the inner and outer radii of the annulus. This
relationship is quoted, for example, by Foulkes et al. (1970). When
94 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

bonding the specimen, it is usual for some of the adhesive to be


squeezed out to form a 'spew fillet' as shown in Fig. 63 and this may be
expected to modify the stress distribution given by eqn (20).
Butt joints for testing adhesives in tension are also usually designed
with a circular cross-section to facilitate manufacture and to maintain
symmetry. In this case, it would appear to make little difference
whether they are annular (Fig. 63(a)) or solid (Fig. 63(b)). In a butt
joint subjected to a tensile load, the adhesive is restrained in the radial
and circumferential directions by the adherends. In the absence of this
restraint, the adhesive would tend to contract radially with respect to
the adherends because of its much lower modulus. The presence of the
adherends has the effect of inducing radial and circumferential stresses
in the adhesive, so increasing the stiffness of the joint. The simplest
analysis makes the assumption that the radial and circumferential
strains in the adherend and the adhesive are zero, in which case the
radial and circumferential stresses are given by:

(21)

where v is Poisson's ratio of the adhesive and (Tz is the applied axial
stress. The apparent Young's modulus (defined as the applied axial
stress divided by the axial strain) is given by:
B,=(Tz= B(l-v)
(22)
Ez (1+v)(1-2v)
where B is Young's modulus of the adhesive.
Kuenzi and Stevens (1963) modified this approach to take account
of the adherend strains by assuming that the radial strain in the
adhesive is equal to the Poisson's ratio strain in the adherends, that is:

(23)

where Va and Ba are the Poisson's ratio and Young's modulus of the
adherends. The radial and circumferential stresses now become:

(24)

These simple analyses ignore the requirement that the radial stress
must be zero at the free boundary of the adhesive, which implies that
shear stresses must exist in the adhesive layer.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 95

The presence of stress concentrations has been suggested by the


analyses of cylinders compressed between rough rigid plates by Filon
(1902), Pickett (1944) and Benthem and Minderhoud (1972). Al-
though these analyses deal with much smaller aspect ratios (the ratio of
the diameter to the adhesive thickness) than would be encountered in
adhesive joints, the boundary conditions are similar to an axially
loaded butt joint. Gent and Meinecke (1970) have calculated the stress
distribution in a rubber block in tension or compression between rigid
adherends, making the assumption that the rubber is incompressible
(i.e. its Poisson's ratio is 0·5). Lindsey (1966) has analysed the butt
joint assuming rigid adherends and a uniform stress distribution across
the thickness of the adhesive. This analysis has indicated that shear
stresses are present near the free surface of the joint. Harrison and
Harrison (1972) have analysed a two-dimensional butt joint in tension
using triangular finite elements. They assumed that the adherends were
rigid and that the adhesive was in plane strain. They obtained a
maximum shear stress near the free boundary. Alwar and Nagaraja
(1976a) analysed axisymmetric and two-dimensional plane stress butt
joints in tension assuming flexible adherends, also using triangular
elements.
All the above analyses, with the exception of those by Kuenzi and
Stevens (1963) and by Alwar and Nagaraja (1976a) make the assump-
tion that the adherends are infinitely stiff compared with the adhesive.
Although the adherends will normally have elastic moduli at least an
order of magnitude greater than those of the adhesive, there will
always be some Poisson's ratio strain in the adherends when the joint
is loaded. Also, the interface does not remain plane under load since
the stresses in the adhesive are not uniform. None of the above
analyses makes any reference to the radial and circumferential stresses
in the adhesive, except for that by Kuenzi and Stevens (1963), which
makes the unrealistic assumption that the axial stress distribution in
the adhesive is uniform. As in the case of torsion specimens, the
presence of a spew fillet may affect the stress distribution, and none of
the above analyses takes this into account.
Adams et al. (1978b) studied solid and annular butt joints loaded in
torsion and tension, and examined the effects of adherend ftexibility in
the spew fillet. They used an eight-node parabolic isoparametric ele-
ment which gives a good estimation of the stresses in regions of high
stress gradient. Their results were confined to the linearly elastic
behaviour of aluminium adherends and an epoxy based adhesive which
96 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

1-2 r----r-----r----,.------,---.---,--,...--.--,------,

3 t{lf---t--
c
"~ 08 f---+--+---l-
~o

.'
- .. 0-61---+---+--+---!-

~ Q-41---+-----l---+--?b:;:;.<::.- - - I + - - r - - - - + - ----j
:;:

6 8 10 14 16 18 20
RCld lus/Adh~StVl' thlckntss

- " ' 1 6 at Z=O - - - 'rIO ot z - tl2

Fro. 64. Shear-stress distributions for solid butt joints in torsion (aspect ratio,
40) (from Adams et al., 1978b).

had a Young's modulus of 2·5 GN/m2 and a Poisson's ratio of 0·4.


Figure 64 shows their results for a solid butt joint in torsion. As
predicted earlier (eqn (20)), the stress increases linearly from zero at
the centre to a maximum at the outside. Tbe magnitude of the applied
torque was selected to give a unit shear stress at the outside of the
joint. Tbe same torque was then applied to a similar joint having two
different sizes of spew fillet. Tbe resulting T z 6 shear stress distributions
are also shown in Fig. 64. Over most of the bonded area, the shear
stress does not vary across the adhesive thickness and its general level
is reduced by the presence of the spew fillet. The larger spew fillet
causes the greater reduction in stress. However, near the outside of
the joint, the shear stress at the mid-plane (z = 0) decreases, whereas
there is a considerable increase in shear stress at the adhesive-
adherend interface (z = t/2) near the adherend corner. The presence of
the spew fillet was calculated to give a maximum shear stress (at the
outer radius) of at least 1·8 times higher than that obtained without a
fillet. The stresses in the fillet away from the adhesive layer are very
low, on average less than 20% of the maximum shear stress in the case
of a joint with no fillet, showing that the shape of fillet assumed for the
finite-element mesh is not significant.
The shear stress distributions for an annular butt joint with and
without a spew fillet are shown in Fig. 65. As with the solid butt joint,
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 97

I
~~
I
t
1·0
--i--=:t: ,"
I .\
I -<
p =--- S~QII .pe" l~rq••p."
'I \
.
I

I I
~,

I I I I 1\........
2/ I I
Adh.rtnd
I
90 95 100 105 110 1\5 1175
Radtu\ I AdhUlVf tt'lIcknns

- ....'8 ot 1=0 - - - .,." ot r = t f'l

FIG. 65. Shear stress distributions for annular butt joints in torsion (from
Adams et al., 1978b).

the torque was scaled to give a maximum shear stress of unity in the
joint with no spew fillet. As before, the spew fillet reduces the general
level of stress and yet introduces a stress concentration near the
adherend corners.
In practice, most structural adhesives exhibit considerable plastic
deformation when subjected to a shear stress, and it is quite probable
that the small volume of adhesive near the adherend corner will yield
without causing premature failure of the joint when loaded in torsion.
However, the spew fillet also makes the joint stifter in torsion and,
since this would lead to an overestimation of the shear modulus of the
adhesive, it is probably advisable to remove any spew fillet present on
test specimens, although this may be difficult on the inside surface of
'napkin ring' specimens.
In tension, simple analyses have often been made in which the
adherends were assumed to be rigid and there was no spew fillet. The
resulting stress distributions can be very misleading. Adams et al.
(1978b) showed (Fig. 66) the axial, radial and circumferential stresses
at the mid-plane of the adhesive (z = 0) and at the adhesive-adherend
interface (z = t/2). The shear stresses are plotted at z = 0, t/4 and t/2. It
should be noted that the stress distributions on the mid-plane are very
similar to those obtained using rigid adherends for the same adhesive
98 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

12.------.---,---,--~--~------~--_,--~

101__- - - ' - - 1 - - 1 - _ + - -!----+----+----+-'~<f-___lI

~08 1__-,--~---+--_+--_+--~---

- : r - t l2 - - -:r=0

FIG. 66. Stress distributions for solid butt joint in tension (no spew fillet,
aspect ratio, 20) (from Adams et al., 1978b).

Poisson's ratio. It should also be noted that in the peripheral region


there is a significant variation of stress across the adhesive thickness.
On the mid-plane of the adhesive, the direct stresses decrease to low
values at the free surface (zero in the case of the radial stress) and the
shear stress is always zero.
However, on the adhesive-adherend interface, there is a stress
concentration caused by the dissimilarity of the materials. Unlike the
case of an adhesive layer between rigid adherends, where the work on
cylinders compressed between rigid plates by Benthem and Min-
derhoud (1972) shows that there is a stress singularity at the periphery
of the adherend-adhesive interface, the stress concentration at the
interface between the adhesive and an adherend of finite stiffness is
likely to be finite. If the adherend had the same stiffness as the
adhesive (i.e. a unifonn bar), there would be no stress concentration. If
the adherend stiffness is then gradually increased, it is unlikely that the
stress concentration will become infinite until the adherend stiffness
becomes infinite. The stress concentration is difficult to detennine
because of the high stress gradients, but from the gradient of the
displacement curve, the (Tz stress concentration was calculated to be at
least 2·5.
In the central region, the axial stress is slightly higher than the
average applied stress necessary to satisfy the equilibrium condition
THE NATORE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 99

12

~ 10
w
~

"008
;;
0
"0

.[0&
g-
w
~O~ ~

~
E02
V;

00
~ 5 6
Rad ius /AdhulVf thlCklltH

FIG. 67. Stress distributions for solid butt joint in tension with and without
spew fillet (aspect ratio, 20) (Tz and (Te values at z = t/2 omitted for clarity:
- - with spew fillet; ----without spew fillet (from Adams et al., 1978b).

that the axial stress integrated over any plane must equal the applied
stress in the adherend away from the joint. If the value of O'z in the
central region is substituted into eqn (24), the values of O'r and 0'8-
obtained from the Kuenzi and Stevens analysis are equal to the values
obtained by the finite-element analysis.
In Fig. 67 is shown the effect of a spew fillet on the mid-plane stress
distributions. As the spew fillet transmits some of the axial load, the
general level of the axial stress, and therefore the other stresses, is
slightly reduced. The smaller spew fillet produces a similar but less
pronounced reduction in stress levels. The presence of the adhesive
fillet reduces the stress concentration predicted at the adherend corner,
the value of O'z stress concentration obtained from the gradients of the
displacement curves being 1·8. The size of the fillet has little effect on
the magnitude of this value. As in the torsion case, the average level of
stress in the spew away from the adhesive layer is low, being less than
10% of the average stress applied to the joint.
In practice, it is probable that some stress relaxation will occur to
reduce the high tensile stresses near the adherend corner, although
high-strength structural adhesives often exhibit only limited non-linear
behaviour in tension. Figure 68 shows the failure surface of an annular
butt joint which failed in tension. The adhesive was the epoxy-based
BSL-308A (manufactured by Ciba-Geigy (UK) Ltd) and the
100 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

FIG.68. Surface of butt joint failed in tension (from Adams et al., 1978b).

adherends are EN25 steel, shot-blasted and degreased in tri-


chloroethylene vapour prior to bonding. The adhesive was cured at
150°C for 1 hand the joint failed at an average applied tensile stress of
80 MN.m- 2 . The failure was mainly cohesive (i.e. failure wholly within
the adhesive layer), but elose to the edges the joint failure was
adhesive (i.e. at the interface), presumably because of the high axial
stress concentration. An area of the adherend surface was exposed
near the outer spew fiBet. At the inner radius of the annulus, the
failure was at the interface between the adhesive and the other
adherend. The difference between the adhesive and the cohesive
failure surfaces is quite pronounced. The presence of this stress con-
centration, and the probability of premature failure initiating in this
region, suggests that the average stress applied to a butt joint at failure
is an underestimation of the tensile strength of the adhesive. However,
this type of specimen should give a reasonably accurate stress-strain
curve up to the point at which this premature failure occurs, providing
a high aspect ratio is used to ensure that the central region of uniform
stress is large compared with the peripheral region. When interpreting
the data obtained from this type of specimen, it should also be
remembered that most of the adhesive is subjected to a triaxial stress
state which, apart from increasing the stiffness, will probably affect the
yield or failure behaviour.
Further work by Adams and Coppendale (1979) considered what
happens when the adhesive yields. They used a pressure-dependent
yield criterion (i.e. one in which the von Mises deviatoric yield stress
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 101

100

oll / \ A FA'LURE

~
oll
W
CI: / FAllURE
.. 50
oll:
.... ::E
<I
"TENSION
~

-006 -004 -002 0 ,02 004


AXIAL S TRAIN

f
1-50
f
I
I

-100

COMPRESSION
-150

-~'
,
"

"1',
,
-,, -200

,
-,.
,
.(
-250

,
,-

-300
,-

-350

FIG. 69. Adhesive stress-strain curves from BSL308A butt joint tests: -e-
specimen A; --+-- specimen B (from Adams and Coppendale, 1979).
102 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

causing yield increases linearly with hydrostatic pressure) as proposed


by Raghava et al. (1973). The yield surface becomes a parabola of
revolution. They postulated that under compressive loads, the adhesive
in a butt joint should not yield. Indeed, if yielding does begin to occur
then, since Poisson's ratio will increase, the yield zone would be very
restricted.
To investigate this phenomenon, aseries of controlled tests was
carried out on three Ciba-Geigy epoxy-based materials (a structural
adhesive BSL308A, and two casting resins, A Y103 and MY750). The
BSL308A butt joint specimen used high-strength steel adherends. It
was loaded to a nominal uniaxial stress of 340 MPa in compression
(see Fig. 69) which approached the yield stress of the steel. The joint
was then unloaded and re-loaded in tension. Yield occurred at about
20 MPa and failure at 78·8 and 80·2 MPa (two specimens). Similar
tests were carried out on other materials up to 150 MPa (HE-30WP
aluminium alloy adherends). Again, in compression the stress-strain
curves were linear while tensile failure occurred at much lower
stresses.
More will be said of the behaviour of butt joints in the chapter
devoted to testing adhesives (Chapter 3), where comparisons are made
between bulk and thin film properties.

THE USE OF JOINTS IN DESIGN


The essence of the effective use of the joints that have been described
is to design for bonding and not simply to substitute it for other me ans
of joining. Care should therefore be taken that stress concentrations be
avoided and the loads be carried over as large an area as possible. Peel
loads are the greatest enemy of the designer of bonded joints. Wherever
possible, the adhesive should be loaded in shear so that peel and
cleavage stresses are avoided.
Figure 70 shows how unavoidable peel loads can be countered by
using a toughened adhesive which will yield extensively before frac-
ture. In Fig. 70(a), although there may be a large total bond area, only
a very small proportion is loaded, so that local stresses are very high
and fracture occurs at low loads. The toughened adhesive (Fig. 70(b))
yields extensively prior to failure, thus allowing the load to be carried
over a larger area. However, by applying a simple strap as shown in
Fig. 70(c), the higher strength of even the toughened adhesive peel
joint can be increased by up to 10 times.
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 103

Tension I ~
compression~
(a)

Tension I /\
compresslon~
(b) •t
......... strap

(c)

FIG. 70. Dealing with unavoidable load in T-peel: (a) brittle adhesive; (b)
duetile (tough) adhesive; (e) strapped peel joint.

Increasing the width of lap or peel joints increases the strength pro
rata, whereas increasing the length is beneficial only for very short
overlaps. However, the reader's attention should again be drawn to
Hart-Smith's philosophy of the benefits to be gained from having a
large area of lightly-stressed material in the middle of the joint,
especially when creep and fatigue need to be taken into account.
104 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

But joint design also means being able to assemble the parts in such
a way that a good bond is reliably achieved. Entrained air pockets can
result in adhesive being pushed out during assembly or heat curing.
Designs in which the adhesive can run out of the glue-line should also
be avoided, or countered positively by using gap filling adhesives or
some means of sealing the leakage paths. Also, it is often necessary to
provide jigs and fixtures to hold and locate the parts accurately until
the adhesive has cured.
There are many successful ways of using adhesive joints. In the
preceding pages, some indication has been given of how the designer
can analyse the stresses in a joint, or at least how to understand the
mechanics and to recognize which situations may lead to high local
stress concentrations. To proceed to a successful design the following
information needs to be assembled: the geometry and materials of the
parts to be joined, the service loads, the environmental conditions and
what requirements exist for inspection of the finished joint or product.
These may range from nil to fuU non-destructive testing. With this
information, we can:
(i) design the local geometry of the joint;
(ii) select a suitable surface pre-treatment;
(iii) select a suitable adhesive;
(iv) decide how to assemble and (if necessary) to jig the joint.
It is impossible to cover all types of joint, so a selection of situations
is given which should give the designer confidence in how to proceed.

Lap Joints
Figure 5 shows the most common forms of lap joint. Some are stronger
than others, but none is simpler to make than the single lap. However,
the bevel, step and butt-strap have the advantage of presenting at least
one smooth surface to the air ftow, customer, and so on. Scarfing or
tapering is of limited benefit since Thamm (1976) has shown that the
adherends have to be tapered to a fine edge if significant benefit is to
be achieved, and this is usually impracticable. In any event, the use of
a toughened adhesive immediately dispels much of the advantage of
tapering joints, except where a smooth surface is required for other
reasons.
Where peel is encountered with laps, or even to counter the peel
loads inherent in loading even the double-Iap joint (see Fig. 11),
various techniques can be used; these are iUustrated in Fig. 71. Of
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 105

Rivet. bolt or spot weid

Increase area

Increase stiffness

Bead end (jf possible)

FIG. 71. Other techniques for combating peel.

these, the best is probably the positive constraint of the joint end by
riveting, bolting or spot welding.
Similarly, despite Thamm's advice, some designers choose to try and
reduce the stresses at the joint ends by reducing the stiffness. This is
best achieved by tapering, although single or multiple grooves or
rnilled ends may be used as shown in Fig. 72.
Arecent example of joint design in which one of the authors (R. D.
Adams) was involved illustrates the mechanics c1early. The problem
106 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

~_ _VMill<2d
IL-_ _ _ _..JoV

~-----i~ Tap<2r<2d
/z::::::: V
IV

~---1V Groov<2d

l<-_ _ _ _---"V
FIG. 72. Joint ends with reduced stiffness.

was how to bond steel to unidirectional carbon fibre reinforced plas-


tics, using a modern, rubber-toughened epoxy adhesive. A simple lap
joint, shown in Fig. 73(a), failed by transverse cracking of the compo-
site. This was due to stress concentrations at the change of section
which led to high loads being transferred aCrOSS the composite to the
steel. Tbe steel was therefore finely tapered as shown in Fig. 73(b) to
reduce the spatial load transfer rate. Tbe same type of failure was
observed at a load only slightly sm aller than before. Reversing the
taper as in Fig. 73(c) had a similar effect, only slightly increasing the
joint strength. Finally, in an attempt to reduce further the peak stresses
at the edge of the joint, a fillet of adhesive was cast during manufac-
ture, as illustrated in Fig. 73(d). With an inc1uded angle of approxi-
mately 30° in the fillet, an increase in joint strength of more than
3-fold was achieved. In this case the transverse stresses were suffi-
ciently reduced to prevent failure initiating within the composite, and
failure occurred by cracking through the adhesive layer as shown in
Fig. 73(d).
What this example illustrates is how basic mechanics can indicate the
likely mode of failure and provide an indication of how to improve the
situation. A finite-element analysis was used to prove correct the
design 'hunch' and it was able to predict not only the initial mode of
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 107

BQ
Fallur~ load
(25rrm ~ joint)
kN
- ' fWgll/llIllIlIllZlIllIIll~ 2, 233
<t.--- CFRP_ ~ _ ~ f · 15_
(al Slmpl~ lap JOint \rransv~rs~
f
rac
t
ur~
of composlt~

(bl Scarfing of st~~1


--
_"':'>-'---L _ _

(c) R~v~rs~ tap~rlng 01 st~~1

3()0 angl~ 01

=:E'llllllmmm,~IIII"
7 --=--1_._
- _ __

Adh~siv~ failur~

(d) R~v~rs~ tap~rlng cf st~~1 and adh~siv~ 11"~t

FIG.73. Bonding of steel to CFRP (Iength dimensions in mm).

failure and the load, but also how this mode of failure would alter with
a corresponding massive improvement in the joint strength.

Tubular Joints
Many of the principles which govern lap joints also apply to tubular
joints such as those shown in Fig. 74. Tbe rotational symmetry helps
load transfer, although transverse peel loads are still present (Adams
and Peppiatt, 1977). All of these designs are difficult to inspect since
the inner surface is inaccessible Figs 74(a) and (c). The sleeved tube
fitting in Fig. 7 4(b) can be assembled wet or dry. If dry assembly is
used, the adhesive is pumped into the joint through radial holes in the
sleeve. In the joints shown in Figs 74(d) and (e), an air space must be
left at X, or a vent to the outside, since air will be trapped when the
joint is assembled, forcing out the adhesive. Alternatively, the 'vent
hole' may be used to force adhesive into the joint which is assembled
108 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

(al 84V4114d tUb4

(b) SI44V4d tUb4

r:::::.'" :l~
(cl B4V4114d tub4 and shaft

(d) Parall41 shaft

FIG.74. Tubular joints for axial or torsionalloading.

dry. Fixing, or the use of spacers such as wire in the glue-line, is


necessary in this case (and often advisable in the others) so as to
maintain both alignment and a consistent bondthickness.
In all of these joints, relative rotation of the mating surfaces helps to
spread the adhesive evenly, but this must in no circumstances be
allowed once a eure has begun.

T-joints
In this category we consider joints such as those shown in Figs 75(a)
and (b) in which the members may be at 90° or at some angle 6. The
loading may be in the plane of the sheet, as N, or transverse to it, as T.
The joint design is more complicated than for lap joints or tubular
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 109

(a)

FairttGOOd

~'-r~ (c)
Good"Good

~~I~
(e)
GOOdt,GOOd

GOO,1-7'
(9

~ Fair Good
0)

~a,,_ ~_F~"
(k)

FIG.75. Possible T-joints.

joints and possibilities are shown in the remaining sketches of Fig. 75.
Most of these are self-explanatory. In general, poor results are always
obtained when the joint is stressed such that tensile transverse (peei)
stresses occur. Compressive transverse stresses do not generally cause
failure. Compressive forces in the direction of the sheet (N in Fig.
75(a» will rarely cause problems unless the sheet can buckle. In this
case, the otherwise satisfactory joint shown in Fig. 75(g) may fail in
110 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

L (al (bl

FIG. 76. Possible corner joints.


THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 111

peel at the upper end if the buckle occurs as in Fig. 75(i), while the
buckle shown in Fig. 750) would not lead to such a failure. The
possibility of buckling under compressive loads must therefore be
taken seriously and joints such as those shown in Figs 75(e) and (h)
given consideration. It may be possible so to dispose the structural
stiffnesses that the buckle will always occur as in Fig. 75(j), but this
must be regarded as a chancy procedure.
Grooving the base as in Fig. 75(k) is normally expensive and may
weaken this member significantly since, unless the groove is deep, the
joint can only be regarded as 'fair'.

Corner Joints
Where two separate sheets or plates meet at 90° or some angle f) as
shown in Fig. 76(a) and (b) similar rules apply as to the T-joints.
Figure 76(c) is generally weak, while Fig. 76(d) is weak only to one
mode of loading, but bear in mind that the compressive loading would
tend to bend the bottom plate, promoting a buckle which would lead
to peel in the upper part of the joints. The joint shown in Fig. 76(e) is
generally good, but difficult to assemble. Bending one or both of the
sheets is possible for these joints, leading to the configurations shown
in Figs 76(f)-(k) inclusive.
If this joint forms the corner of a box which takes internal pressure
then joint (i) would appear to be a strong contender while joint (h)
would not. Unfortunately joint (i) is weak for the in-plane tension but
the modification shown in Fig. 76(k) overcomes this problem.
Figure 77 shows two sequences of construction suggested and tested
by Keimel (1966). Figure 77(a,b) gives two methods of joining T section
aluminium to give a right-angled corner, which was tested by opening
th~ corner by loading in the directions indicated by the arrows. The
failure loads for the two joints stood in the ratio of 1 : 1· 20, that for the
weakest joint being 8 kN. Figure 77(c-f) shows a corner formed from
angle pieces, the four methods giving strengths standing in the ratios
1: 0·98: 1 : 1·16 where unity is equivalent to 9 kN. Standing one angle
inside the other seems not only less elegant than cutting mitres but also
less efficient as other metal must either be milled, as in the second
design or formed as in the third although it must be admitted that the
strongest corner with mitred members also involves formed corner
reinforcement.
112 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

d ...J'

FIG.77. Corner formation from T-sections ((a) and (b» and L-sections ((c)-
(f».

Butt Joints
Butt joints are strong in tension and very strong in compression.
However, they are very weak if subjected to transverse loading owing
to the large cleavage stresses. The type of joint shown earlier in Fig.
5(g) is satisfactory in most types of loading, and is really the same joint
discussed above (Fig. 74(b)) for tubular joints.

Stifteners
Where large areas of sheet metal or fibre reinforced plastics are used,
it is common to stiften these by attaching deep sections, either by
welding, riveting or bonding. As an example, the 'top-hat' design is
THE NATURE AND MAGNITUDE OF STRESSES IN ADHESIVE JOINTS 113

(a) Original design

(b) Increased bond area

(c) Reduced flange stiffness

(d) Increased sheet stiffness

(e) Load spreading plates

FIG. 78. Stiffeners and load spreading.

shown in Fig. 78, and it can be seen how the designer can vary the
strength and added stiffness by using different configurations. It is also
worth pointing out that a bonded joint is also inherently stiffer than
one made with rivets or spot welds.

Doublers
If dismantleable components are to be attached to sheet metal, rivetted
or screwed connectors have to be used. To avoid local distortions and
to help to diffuse the point loads, bonded doublers are used as is also
shown in Fig. 78.

Assembly
Attention has already been drawn to the need to jig or otherwise
locate the adherends during many bonding procedures. With large
114 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

sheet struetures, it may be possible to restrain the adherends by spot


welding through the q.dhesive in a few plaees (sometimes ealled taeking
or weld-bonding) until the adhesive eures. An inereasing use is being
made of adhesives in motor ear struetures. Although the adhesive may
be eured in a jig, the body is then subjeeted to high temperatures in
the paint oven, which may soften the adhesive so mueh that parts may
slip or fall off. Weld-bonding ean be used here also.
Weld-bonding is diseussed in greater detail in Chapter 5. Although
no formal analysis of the stress-bearing behaviour of weId bonding has
been made, eonsideration should be given to the question of the
apportionment of the load between welds and adhesive.
Chapter 3

Standard Mechanical Test Procedures

The standard test procedures as listed by ASTM, BSI and other official
bodies are essentially for testing adhesives and surface treatments rather
than joints. Unfortunately, most if not all of these standard tests consist
of joints in which the adhesive stresses are far from uniform. Let us
therefore examine the reasons why we might need to carry out any
form of test on an adhesive, other than the not unworthy cause of
sheer curiosity.
Let us suppose that the adhesive is to be used in some joint, the
geometry and loading of which is quite complicated. Ultimately, the
only real test of suitability is to build a representative sampie of these
joints for each candidate adhesive and surface treatment. This, of
course, is costly and begs the question as to whether the joint design is
itself satisfactory. The designer will wish to call on previous experience
with adhesives, surface treatments, and joint designs so as to re ach a
90% certainty of success before he builds and tests a structural pro-
totype. If structures are expensive, it will be difficult to justify more
than a very limited series of prototype tests before production ensues.
Fortunately, over the last 20-30 years, aseries of standard tests has
been developed by means of which the mechanical properties of most
adhesives can be tested. Unfortunately, the designer has to select the
appropriate testes) and to know what the results mean in terms of his
particular application.
So what are we looking for? All the theoretical methods for predict-
ing joint strength need the elastic moduli such as E (Young's) and G
(shear). In addition, those theories which allow for adhesive non-linear
behaviour will need data such as the yield stress (strain) and the
ultimate stress (strain). More sophisticated analyses, e.g. Adams et al.
115
116 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

(1978a), Adams and Coppendale (1979), in which the adhesive yield is


considered to be function of the hydrostatic component of stress, need
the ratio of compressive to tensile yield points, (J'TY/UCy. This ratio is
usually of the order of 1· 3-1·4. On a simpler plane, we may simply
wish to rank a group of adhesives or to check the shelf life of a batch
by comparing the strength of joint made therefrom with those made
from 'fresh' adhesive. But the strength of a joint also depends critically
on the treatment given to the faying surfaces prior to the application of
the adhesive. This is important not only for the short-term strength
but, even more so, for the long-term strength, especially when aggres-
sive environments may be encountered. And finally, there is a matter
of great concern to all engineers. This is the legal requirement to
ensure that the correct steps were taken to select, prepare and apply
the adhesives and adherends so that product and personalliability are
kept in mind. A condensed list of the ASTM and BSI procedures is
given in the Appendix.
Non-destructive testing is also considered to be an important aspect
of quality control and is treated in this chapter.

DESTRUcrIVE TESTING

Tests with Thin Sheet Adherends


Many uses of bonded structures have occurred in aerospace situations
where the adherends consist of thin sheets of aluminium alloy or some
other light-weight material. The most commonly used test is the
single-lap joint illustrated in Fig. 79(a). These dimensions are as
specified by ASTM D 1002-72 which also specifies the adherend ma-
terials. It is recommended that the specimens be cut from a 177 mm
(7 in) wide bonded plate since this gives the most representative
results. The outer strips should be discarded. Note that this joint is
automatically misaligned before it is placed in the testing machine.
Some laboratories bond tabs at the ends to improve alignment, as
shown in Fig. 79(b). Even so, the joint will bend as shown in Fig. 79(c),
giving rise to large transverse peel stresses in the adhesive layer. As
pointed out in the previous chapter, the adhesive shear stress is
non-uniform, owing to differential straining in the adherend and other
factors: this is illustrated schematically in Fig. 79(d) together with the
associated adhesive transverse stresses in Fig. 79(e). Even though it is
still recommended in ASTM D 1002-72 and such standards that the
STANDARD MECHANICAL TEST PROCEDURES 117

l 1
144_--=6.=..3.-=.5_---.tJ
101·6
~_ 1>7'0-25
I
..
(al Single-Iap joint

F"'-~Alignm~nt ~abs;::::::::-----"2 s=J---.F


(bl

F...-C:::::
(cl

F~ ~I---,.........L__~
~__________---,I -'F

Adheslve~
shear stress ______ I-\verage
A.
shear
stress = FIArea

U
'--------'
(dl

Transverse adheslw
stress (actlng across
the bond-line thickness)

(e)

F'Io. 79. Single-Iap joint test piece to ASTMD 1002-72 (1978) (dimensions
in nun).

results be given as the mean shear stress at failure (i.e. load divided by
bond area) it has long been reeognized that this average shear stress
bears little relationship to what is actually happening in a joint,
especially when geometrie, adhesive artd adherend non-linearities be-
come significant. Reeent work by one of the authors (R. D. Adams)
and his colleagues has shown that it is important to observe that the
adherend material specifieation be kept to. In a partieular applieation,
a eompany whieh was proposing to use a new, rubber-modified epoxy
adhesive with either soft or half-hard aluminium alloy adherends
118 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

decided to carry out lap shear tests using this material for their
adherends. The results for the single-lap tests were disappointingly low
and it was shown (Rarris and Adams, 1983) that this was due to extensive
yielding of the adherends prior to joint failure. When the tests were
repeated using 2L73 alloy or equivalent, a much higher adhesive 'shear
strength' (load/area) was obtained.
The advantages of the single-lap test are that it is simple, cheap, uses
a standard tensile testing machine, and there are a lot of data available
for comparison. Its main disadvantage is that the reported nominal
shear stress bears litde relationship to any intrinsic adhesive property.

~2.--rl--------------':TT
~~~.I______~,~~(__ F
F'2+-1~------------~

~
Adhesive
shear

H
stress

tr~~~~:~se
stress

(b)

T C T
SC
(c)

FIG. 80. Double-Iap joints: (a) to ASTMD3165-73; (b) stresses in a double-


lap joint; (e) double butt-strap joint (dimensions in mm).
STANDARD MECHANICAL TEST PROCEDURES 119

Curiously, this very disadvantage is also an advantage since no-one


really believes that the average shear stress means anything fundamen-
tal while in other tests with apparently more precisely controlled
conditions, it is COffiffion for misleadingly definite values to be quoted.
Also, the complex stress situation which pertains in the single-Iap test
makes it quite representative of many structural applications and
loading situations. This test is widely used, often abused, but remains
one of the most trusted standards.
Nevertheless, there have been many attempts to improve the single-
lap test. These include the laminated assembly (ASTM D 3165-73)
shown in Fig. 80(a). Some of the load non-linearity is removed com-
pared with the simple single-lap joint, but differential adherend strain-
ing and high transverse stresses still exist. The advantage of this type of
specimen is that it may be taken from the scrap edge of a laminated
sheet assembly so that production parameters may be more reliably
checked. This type of joint is sometimes wrongly referred to as a
double-lap joint.
Another variation of the lap test is the true double lap illustrated in
Fig. 80(b). Even in this case, as was shown in Chapter 2, we still have
differential straining of the adherends and the loads are still non-
colinear internally, even though the external forces balance. Such
joints tend to fail at T where the adhesive is in transverse tension
rather than at C where it is in compression. The double butt-strap joint
shown in Fig. 80(c) is a form of double-Iap joint and again it fails at T.
Sometimes, if the gap between the two tension members is small, most
of the load is carried directlY across the stiff, butt faces. Premature
failure ensues here but the joint is still able to carry a substantial load
in excess of this.
A recently developed method is the so-called 'thick adherend test'*,
(ASTM D 3983-81) in which it is attempted to minimize the effects of
differential straining using stiff, thick, metallic or wooden adherends.
In this form of joint, there is a considerable increase in the flexural
stiffness of the adherends. The combination of these two properties has
led to the popular belief that the adhesive is now in astate of uniform
shear and there are no significant transverse peeling loads (but see
Renton, 1976). .

* Strictly speaking,
this should be considered in aseparate section, since ipso
facto thick adherends are not thin adherends as this section is entitled.
However, similarity of action advises its inc1usion.
120 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

However, the results of this test should be interpreted with caution.


There are two possible sources of error, a non-uniform adhesive stress
distribution and adherend deformations, the magnitudes of which are
not normally quoted. In its form as specified by ASTM D-3983, the
thick adherend test is limited to the measurement of the shear modulus
of nonrigid adhesives. Specifically, the ratio of the adherend tensile
modulus (E) to adhesive shear modulus (G) should be greater than
300 : 1 which, for steel adherends, limits the adhesive shear modulus
that can be measured accurately to 700 MPa. For a particular adhesive,
curves based on the analysis of Goland and Reissner are used to
determine the geometry required to produce a sufficiently uniform
shear stress distribution along the adhesive layer. Deformation of the
glue-line is measured by extensometry attached to the adherend sur-
faces. However, using closed-form analysis, Renton (1976) has shown
that bending deformations of the adherends can introduce significant
errors, and recommends that if an adherend-surface measuring device
is to be used, it should be attached to each of the outer adherend
surfaces which are parallel to the glue-line, on the centre-line of the
overlap, so that the errors due to bending in the adherends will offset
each other. Because of the limited accuracy of positioning such exten-
sometry, Renton gives a realistic limit for the use of a surface measure-
ment device as an E to G ratio of 2000: 1. Generally, structural
adhesive properties will lie outside this range so that the test is not
really applicable.
In order to overcome the problem of errors due to adherend
deformation, the KGR-1 extensometer was developed by Krieger
(1980): this is attached to the thick adherend specimen at points only
1· 2 mm each side of the glue-line. However, even with this device, a
correction is required, significant in many cases, for the adherend
deformation between extensometer and glue-line. Again this is not
generally quoted, so that its significance is not known. By using the
KGR-1 extensometer, not only is adhesive shear modulus data pro-
duced, but complete shear stress-strain curves are generated up to
failure, based on the assumption that the glue-line deformation is a
uniform shearing action over the entire range.
In the PABST program, the thick adherend test was compared with
the napkin ring test for a nylon-epoxy adhesive FM73. The napkin ring
method gave a shear modulus of 590 MPa, whilst the thick adherend
test gave values as low as 275 MPa. Also, the thick adherend test gave
a lower yield and ftow stress and an almost doubling of the failure
STANDARD MECHANICAL TEST PROCEDURES 121

strain. This result is cmious as there will be some element of peel


present at the edge of the adhesive layer in the thiek adherend test
piece, which should promote premature failure of the glue-line, leading
to lower apparent shear strains to failure than true material values.
Although the thiek adherend test piece may be a cheaper and more
simple joint to manufacture than the napkin ring test piece, it has a
number of limitations to its applicability that appear to be not gener-
ally recognized. In partieular, the measurement of the small glue-line
deformations requires considerable experimental care and accuracy, if
reliable results are to be obtained.

Tests for Properties of Adhesives


In order to determine the mechanieal properties of adhesives whieh
can then be input to the various stress predietion techniques, these
have to be unambiguously determined. Controversy still exists as to
whether adhesive properties in the thin film form (whieh is how they
are normally used in joints) are the same as when bulk specimens are
prepared. For instance, Volkersen (1965) quoted experimental work
by Muller (1959) which showed that Young's modulus of an adhesive
decreases as the glue-line thickness increases, whereas the shear mod-
ulus is independent of this same glue-line thiekness. Similarly,
Franzblau and Rutherford (1967) obtained a higher value of Young's
modulus in the thin film form than in the bulk form. However, in both
cases, it was questionable as to whether the transverse restraint im-
posed by the adherends was not affecting the values. On the other
hand, it is possible that the process of producing a thin layer of
adhesive may cause the material to become anisotropic. Also, there
may be some 'contamination' of the purely polymerie content by the
adherends and their protective coatings. Finally, owing to the possibil-
ity of exothermic reactions on curing, many polymeric adhesives are
difficult to produce in the bulk form. Thus, tests based on adhesives in
both the thin film and bulk form have been used. We therefore give
below a critical appraisal of some of the more commonly-used tests.

Axially Loaded Butt Joint


(Sometimes called the 'poker chip' test.) This is effectively a butt joint
to whieh tensile loads are applied, as shown in Fig. 81(a). At first sight,
this appears to be a simple test in whieh the adhesive is loaded
uniformly in tension. Unfortunately, this is far from the case. If the
adhesive layer were thiek (i.e. many times its lateral dimensions) as
122 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

-(a)

'Adhcz~nd ' 'Adhczslvcz'

(c)

(d)

Fro. 81. Tensile butt joint tests: (a) 'poker chip' test (thin adhesive film); (b)
and (c) schematic for thick adhesive test; (d) gauge length and deflections in
'poker chip' test.

shown in Fig. 81(b) and then pulled axially, it would neck as shown
schematically in Fig, 81(c). This is because the same load is carried in
both the adherends and the adhesive, and because if the former are of
sayan aluminium alloy and the latter an epoxy, the ratio of the moduli,
E 1 /E2 will be 20 or more. Thus, the axial strains in the adhesive will be
20 or so times greater than those in the adherend, with a similar ratio
for the lateral (Poisson's) strains. But where the two materials join, the
lateral strain in the adhesive is resisted by the much stifIer adherends.
The conflict is resolved by generating large radial shear stresses on the
interface. Now, if we revert to the relatively thin film shown in Fig.
81(a), we find that the whole of the adhesive is dose to the interfaces,
and so it is afIected by the adherend restraints and this leads to the
complex stress discussed in Chapter 2. The ratio of the applied stress
to the strain across the adhesive is defined as the apparent or con-
STANDARD MECHANICAL TEST PROCEDURES 123

strained Young's modulus, E'. If the adhesive is completely con-


strained by the adherends and the adherend contraction is assumed to
be zero the ratio of the apparent Young's modulus to the true Young's
modulus, E, is given by Adams and Coppendale (1977) as
E' (I-v)
E (1+v)(1-2v)
where v is Poisson's ratio.
Kuenzi and Stevens (1963) analysed the axially loaded butt joint
allowing for the radial strain in the adherends but igiloring any necking
in the adhesive. They obtained the following equation for the Poisson's
ratio of the adhesive:
2G-E'
v = 2(G- E' +2GE'va /Ea )
where G is the shear modulus of the adhesive and E a and Va are the
Young's modulus and Poisson's ratio of the adherend. The Kuenzi and
Stevens analysis is most accurate when applied to butt joints with a
small glue-line thickness to diameter ratio. However, it is an advantage
in testing to use a reasonably thick glue-line to increase the adhesive
deformation and therefore the accuracy with which it can be measured.
The validity of the Kuenzi and Stevens analysis was investigated for a
circular butt joint using an axisymmetric finite-element analysis
(Adams and Coppendlaie, 1976), and it was shown that the errors are
small (about O· 5% for E) for structural adhesives.
However, if purely elastic behaviour is considered, the axially-
loaded butt joint can be used to provide useful data on the elastic
properties of structural adhesives. The loading can be static or
dynamic. In the former case, the problem is to measure the strains and
displacements across a thin glue-line. Usually, extensometers will be
attached to the adherends and a suitable correction made for the
adherend strains as shown in Fig. 81(d). Then
8 = 81 +82 +8 3
= Bl d l + B2d2 + B3d3
= Bl(d 1 + d 3) + B2d2
where B is the strain and 8 is the deflection in the sections 1, 2 and 3.
If d 1 and d 3 are of the order of 1 mm and d 2 is of the order of 0·1 mm,
then the displacements in the adhesive and the adherend are of the
same order and the test is well-conditioned.
124 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Adams and Coppendale (1976, 1977) used adynamie technique, for


the axial butt joint, using the adhesive layer as a spring in an axially
vibrating system. By comparing the frequency reduction when having
an adhesive layer instead of a solid bar, they were able to calculate E',
the effective Young's modulus. They then tested the same bars in
torsional oscillation and obtained the shear modulus G by the same
frequency change technique. From E' and G, it is possible to calculate
Poisson's ratio P, and hence to obtain E.
But what happens in a butt joint when the loads are taken beyond
the elastic limit? In compression, yield in the adhesive occurs, but,
owing to the complex way in which polymers behave under yield
(Adams and Coppendale, 1979), further yielding may be suppressed
completely. In tension, the butt joint still yields, but, owing to the
triaxial stress state, the nominal stress at which this occurs is greater
than that for uniaxial tension. Thus, the relation between the tensile
strength of butt joints and the tensile strength of bulk specimens
depends on several factors. Bulk specimens are unable to support a
greater stress than the uniaxial yield stress of the material. However, in
a butt joint, gross yielding is suppressed by the predominantly triaxial
stress state in the adhesive, although some local yielding may occur at
the stress concentrations around the edge of the joint. If the bulk
specimens of a particular adhesive fail in a brittle manner, then the
butt joints of the same material are likely to fail at an even lower stress
because of the stress concentrations. However, if the adhesive yields in
a ductile manner in uniaxial tension, the butt joints may be stronger
than the bulk specimens. This was observed by Jennings (1972) for an
epoxy resin at temperatures in excess of about 35°C; Lewis and
Ramsey (1966) reported similar results. A flexibilized epoxy tested by
Adams and Coppendale (1979) had a butt joint strength similar to its
uniaxial tensile yield stress at the testing temperature (20°C). A more
ductile adhesive could have a butt joint strength considerably higher
than its uniaxial yield stress. This explains why brazed joints are often
stronger than the bulk strength of the ductile braze material.
Thus, the complex strain distributions in axially loaded butt joints
make it difficult quantitatively to predict their stress-strain behaviour
and ultimate tensile strength from the bulk properties of the adhesive
without the use of a non-linear stress analysis and a detailed under-
standing of the response of the adhesive to the local stress concentra-
tions around the perimeter of the joint. Conversely, it would be very
difficult to use the stress-strain data obtained from axially loaded butt
STANDARD MECHANICAL TEST PROCEDURES 125

joints to predict the behaviour of an adhesive in a different type of


joint (e.g. a lap joint) in which the stress distributions are likely to be
completely different. Bulk specimens are not susceptible to the prob-
lems inherent in butt joint tests and are more suitable for providing
reliable data on the response of adhesives to various known states of
stress although care should be exercised in controlling the curing
schedule. If a particular adhesive is not suitable for making into large
bulk specimens, it should be possible to obtain uniaxial tensile stress-
strain data by testing an unsupported thin film of the adhesive.
ASTM list two specifications for the tensile testing of butt joints.
The first, ASTM D 897 -78 uses short, stubby, circular specimens of
metal (or wood) with a cross-sectional area of 1 in2 , while the second
(ASTM D 2095-72) allows for longer specimens of metal or reinforced
plastics. For the latter test, specimens (round or square) should be
prepared according to ASTM D 2094-69 which designates alignment
tolerances and fixtures to minimize misalignment problems.

Shear Tests
First, we must (with one exception) disqualify as 'shear tests' the
single-lap joint and its allies described earlier. The exception is the
thick adherend shear test. As discussed above, provided proper pre-
cautions are taken, this test can be used to determine the shear
properties of an adhesive, although the ultimate strength cannot be
attained (para. 4.2 of ASTMD 3983-81).
But favoured for many years is the napkin ring test shown in Fig. 82.
By applying equal and opposite torques T, the adhesive is stressed
purely in shear and the maximum stress, 7', will be that at the outside
radius, rm and is given by:
Tro
7'=-
J

- j}o
.-
_ho.
-'-....}Spew fill ets
Tub uar/
I
adherend
FIG. 82. Napkin ring test geometry.
126 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

where:
7l"
J =- (r 4 - r'7)
2 01

The shear modulus G is given by:

G=J(~)
where b is the glue-line thickness and TIef> is the measured gradient of
the torque-twist curve. Again, ASTM in E 229-70 describe a method
of testing and evaluating the shear modulus and shear strength of
adhesives by the napkin ring test. Wisely, they advise removing the
adhesive spew fillets (inner and outer) prior to testing. Not only does
the fillet permit load transfer to take place, thus making the calculation
of T and G less exact, but as mentioned in the previous chapter,
Adamset al. (1978b) showed that the presence of a spew fillet causes
reduced shear stresses at the outer edge on the mid-plane of the
adhesive, but increased shear stresses at the adherend-adhesive inter-
face. Using typical values for an epoxy-aluminium napkin ring, they
showed that a stress concentration existed of at least 1·8 times higher
than that obtained without a fillet. In practice, most structural adhe-
sives exhibit considerable plastic deformation in shear and it is quite
probable that the small volume of adhesive near the adherend corner
will yield without causing premature failure. In many forms of the
napkin ring test, it is impossible to clean away the fillet on the inner
radius, and this is a likely source of error in both the shear modulus
and the shear strength.
The napkin ring test was used because, if r 0 = 'i, the adhesive is
essentially at the same shear stress (since T is proportional to the
radius, r). However, a modification of the napkin ring shear test can be
used in which the adherends are solid circular bars. This has advan-
tages when an adhesive of low viscosity is used since it is difficult to fill
a napkin ring joint properly. Adams and Coppendale (1977) used the
jig shown in Fig. 83 to produce fully-filled butt joints with low viscosity
adhesives. The adherends were c1amped in accurately aligned vee-
blocks at the required distance apart and adhesive injected from below
using a hypodermic syringe. As the adhesive rises, it displaces the air
and the header pipe compensates for seepage losses or contraction
during curing. A triangular fillet is cast around the joint to ensure
complete filling, but is machined off before testing. A butt joint for
STANDARD MECHANICAL TEST PROCEDURES 127

HEADER PIPE
V- BLOCK JIG
SUPPORTING Sf'ECIMEN SlEEVE

4-- - SYRINGE

FIG. 83. Manufacture of butt joint using liquid adhesives (from Adams and
Coppendale, 1977).

torsion testing is shown schematically in Fig. 84. The overall rotation,


cf>ov, between X and Y can be measured by such as a rotary capacitance
transducer or by linear gauges positioned at some radius, provided the
rotation is sufficiently small. Then:

But:
cf> = l'Ylr

6 [., Er~'~'
I l, l2 L3 I
I I
x y

FIG. 84. Solid butt joint for torsion test.


128 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

where 1 is the length of a right circular cylinder, Y is the radius of the


cylinder, and "I is the shear strain. Also, the shear stress, 'Tb is given
by:
'Tl = G"I

where G is the shear modulus. For an applied torque T, we have that:


T 'T

J Y

where J = second polar moment of area = 'TTy4 j2. Then:

and so:

Using typical values for such a joint, we have that 11 = 13 = 3 mm,


12 = O' 3 mm (= glue-line thickness), G adherends = 26 GPa and G adhesive =
1·3 GPa. Then:

Thus, the twist in the (aluminium in this case) adherends must be


allowed for in calculating the twist in the adhesive when this is e1astic.
However, when the adhesive yields, G2« G 1 and so <P2» <Pl or <P3' But
note that if the joint is twisted at a constant rate, when the adhesive
becomes plastic most of the deflection takes place therein and the
strain rate in the adhesive will increase.
Such a test will give the relationship between rotation and torque. In
order to determine the true shear stress versus shear strain curve, we
need to use Nadai's correction (Nadai, 1931). This is satisfactory even
when the adhesive is partially plastic and partially elastic. The correc-
tion is illustrated in Fig. 85. A general point B on the non-linear part
of the torque-twist curve is shown. A tangent to the curve at this point
is drawn, which intersects the torque axis at C. The correction to the
height AB is then made by subtracting one quarter of the intercept
height De, giving the corrected height AE.
STANDARD MECHANICAL TEST PROCEDURES 129

Peel
In recent years, various forms of the 'peel' test have been used to
assess the performance of structural adhesives. These are shown
schematically in Fig. 86 and are all essentially variations of a common
theme, shown in Fig. 87. Kaelble (1959, 1960) and Crocombe and
Adams (1981a, 1982) showed that the key factor in determining
fracture is the bending moment, M, at the tip of the propagating crack
(Fig. 87) which is reacted over a very short length of the adhesive,
resulting in large local stresses, particularly in a direction across the
adhesive thickness (cleavage). This causes the adhesive to be loaded in
tension and results in the premature failure (and hence low peel
strength) of a number of adhesives which give otherwise satisfactory
performance in such as single-Iap joint tests. Ideally, it should be
possible to relate the force required to cause failure in a peel joint with
the geometry and material properties (of adhesive and adherend) in
that joint. Crocombe and Adams (1982) showed, by using a large-
displacement finite-element technique in which both the adhesive and
the adherend could yield, that it was indeed possible to predict the peel
strength. A rubber-modified epoxy adhesive, which had a high strain to
faHure, was used with high and low strength aluminium alloy
adherends. The former adherends correspond to those specified in
ASTM D 3167-76 (floating roller test) and the latter in the UK Minis-
try of Aviation Aircraft Material Specification DTD 5577 (1965).
Crocombe and Adams (1982) showed that, despite the considerable
difference in yield strength between the two alloys, the peel strengths

Torque

--
Nadai corrected
__ E__ _ .L!h~ar stress
C .... -
,. ,,- '"
AE=AB-DC
4

A
Twist
FIG.85. Nadai correction.
130 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

(cl

FIG. 86. Various forms of the peel test: (a) 1800 T-peel test for flexible-to-
flexible assembly; (b) 180 peel test for flexible-to-rigid assembly; (e) climbing
0

drum test; (d) floating roller test (dimensions in mm).

PEEL LOAD

FLEXIBLE ADHEREND PEEL ANGLE


ADHESIVE

RIGID ADHEREND

FIG. 87. Diagrammatie representation of the peel test (from Croeombe and
Adams, 1981a).
STANDARD MECHANICAL TEST PROCEDURES 131

ought to be (and were) of similar values. Thus, once again, the joint
strength has been shown to be predictable from bulk material proper-
ties. Even so, this will not result in the abolition of the peel test, but
rather provide a c1earer understanding of its mechanics. Since a
considerable amount of data have been accumulated on the various
peel tests, these will continue to be used.
Probably the most commonly used test is the T-peel (effectively cl
900 peel test) as specified by ASTM D 1876-72 (Fig. 86(a)). The
adherends may be metal, fabric or plastics. An older but still
commonly-used test (ASTM D 903-49) requires an adherend that can
be bent through 1800 (Fig. 86(b)). The climbing drum test (Fig. 86(c))
is recommended for determining the peel resistance of adhesive bonds
between a relatively flexible adherend and a rigid substrate, or the
flexible facing of a sandwich structure and its core (ASTMD 1781-76).
Finally, the floating roller test shown in Fig. 86(d) is specified by
ASTMD 3167-76 for bonds between rigid and flexible adherends.
ASTM regard this as a more severe test than the c1imbing drum
method since the angle of peel is greater.
For sandwich structures, the climbing drum method is clearly to be
preferred: the simplest standard peel test is the T-peel method, while
the floating roller technique gives more repeatable results. The 1800
test has little to recommend it beyond simplicity.
Before leaving the consideration of peel tests, we should look at
what is often referred to as the Boeing wedge test, which is now given
the designation ASTMD 3762-79. This test, shown in Fig. 88, is
activated by forcing the wedge into the bond-line of a flat-bonded
specimen of aluminium or other adherends, thus creating cleavage
(peei) stresses in the adhesive. This method is not intended to be
quantitative but it is very cheap and very sensitive in discriminating
variations in adherend surface preparation and adhesive environmental
durability. Essentially, the crack growth rate is noted and the specimen
finally fractured to observe the failure mode-i.e. cohesive, adhesive,
primer/adhesive, etc.

FIG.88. Boeing wedge test.


132 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Dynamic Tests
Since adhesively-bonded structures are often subject to time-varying
loads, it is necessary to allow for the behaviour of joints under these
conditions.
For essentially static loads, creep may be important and there are a
variety of ASTM tests for assessing this (D 1780-72, D 2293-69,
D 2294-69, D 2918-71 and D 2919-71). Creep may be especially
important where environmental attack or high temperatures cause
softening of the adhesive.
For fatigue loading, ASTMD 3166-73 specifies a specimen similar
to the single-Iap shear specimen of D 1002, but with a shorter free
length to resist buckling. Again, fatigue data on bulk specimens might
be more reliable and would certainly be more applicable to other
designs of joint. Romanko (1979) has recently given an important up-
to-date review of bonded joints under cyclic loading, and Althof
(1982) has suggested the thick adherend specimen as a suitable config-
uration for evaluating the fatigue properties of adhesives.
An important property of adhesives is the ability to withstand shock
loading. It is well-known that most, if not all, polymers have rate-
dependent mechanical properties. Usually, with increasing strain rate,
the modulus increases slightly, as does the strength, while the strain to
failure decreases. Harris and Adams (1982) recently showed that, for
single-Iap joints tested under impact or quasi-static conditions, there
was little change in the failure strength: with some adhesives it
increased and with some it decreased as shown in Fig. 89. In joints
subjected to impact loading, there are two main considerations. First,
the joint strength should not be particularly diminished. After all,
no-one wishes to travel in an adhesively-bonded motor car which falls
to pieces when struck. Indeed, the need is not only for strength
retention but also energy absorption. In motor vehicle impact situa-
tions (crash), the modern method of absorbing the impact forces and so
protecting the passengers is by progressive yielding and crumpling of
the structure. Ideally, the energy is absorbed over as long aperiod as
possible and without seriously distorting the passenger compartment.
The work by Rarris and Adams (1983) again showed that the best
energy absorption in lap joints was achieved by using a high perfor-
mance, rubber-modified epoxy adhesive which had a high strain-to-
failure, together with a low yield strength aluminium alloy. They also
showed that it was possible to predict the failure loads in lap shear
STANDARD MECHANICAL TEST PROCEDURES 133

l\ --,,
,
BAND
.~ --'-MEAN
\
......
u u
~

14 ~

:I:
\- r-' tf-
~

12 ~\ I I
I I

~ ,,-
10
[\ --
\
z 8
,\
-" \
\ ,--1--1
I I I

6 ~ I
:-'i --
\ ,'--1--.
~ l\ -

~
~
\ ~
0
~ ~
crBN ESP10S AY103 MY7S0

FIG. 89. Static and impact joint strengths for four adhesives with 2L73 ad-
herends: epoxy modified by carboxy-terminated butadiene-nitrile rubber
(CTBN); ESP 105 (Pennabond Ltd); AY103 and MY750 (Ciba-Geigy Ltd).

from the measured impact stress-strain curve of bulk specimens of the


adhesive.
The principal standard test far impact is ASTM D 950-78. The
apparatus is shown in Fig. 90. The impact velocity is 11 ft/s
(7·5 miles/h or 3·5 m/s). The energy absorbed is measured in the usual
way with pendulum-type impact machines: a correction is made for the
kinetic energy imparted to the upper block. Precisely what this value of

Pro. 90. ASTM impact test piece D 950-78 (Dimensions in mm).


134 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

energy absorbed relates to is not dear, and ASTM regard this striet1y
as a comparative test.

NON-DESTRUCTIVE TESTING

The objective of any system of non-destructively examining an adhe-


sive joint must be to obtain a direct correlation between the strength of
the joint (howsoever defined) and some mechanical, physical or chemi-
cal parameter whieh can readily be measured without causing damage
to the joint. We are therefore looking for faults, which may be defined
as anything which could adversely affect the short- or long-term
strength of a joint. There are two basie areas for examination, the
cohesive strength of the polymerie adhesive, and the adhesive strength
of the bond between the polymer and the substrate.
Adhesive strength is very difficult to measure since it is an interfacial
phenomenon involving a very thin layer of material, thin even in
comparison with bond-line dimensions! Effectively, we would need to
assess intermolecular forces and this is not really possible with existing
techniques. This aspect of quality control is usually reduced to asses-
sing the nature of the adherend surfaces prior to bonding.
The cohesive strength of the adhesive is really the only parameter
which can be estimated with any degree of confidence and it is this
which features in most non-destructive tests of bonded joints.

Nature of Defects
It is instructive at this stage to examine the types of defect which may
occur in bonded structures. First, let us consider lap joints in which the
bonded area is 1 cm2 or more. Figure 91 illustrates many of the
possible defects.
Porosity is caused by volatiles and extrained air in the adhesive. It is
therefore present in most bond-lines to some extent. Adhesive cracks
are due to problems with curing (cure and/or thermal shrinkage) or to
large applied stresses, either one-off or repeated (fatigue). Local areas
of poor eure are due to incorrect mixing of the adhesive system. Larger
areas, possibly extending through the whole bond-line, are either due
to incorrect mixing, incorrect formulation, or insufficient thermal ex-
posure. Sometimes, poor cure is self-correcting with time in that the
chemical re action continues, albeit slowly. However, if the component
is in a cold environment and is quickly subject to stress, failure will
STANDARD MECHANICAL TEST PROCEDURES 135

Poor curcz

Adhczsivcz

Poroslty

Zczro- volumcz
unbond

FIG.91. Typical defects in an adhesive bond-line.

occur. Voids are due to air becoming trapped by the pattern of laying
the adhesive or to insufficient adhesive being applied. Large voids
cannot be caused by volatiles, unless something is very wrong with the
adhesive system. Surface unbonds are an alternative form of void,
often caused when adhesive is applied to one adherend only and
unevenly. Zero-volume unbonds occur where the adhesive and
adherend are in contact, but there is no significant bond strength
between them. Such defects may be caused by poor surface prepara-
tion, failure to remove completely the manufacturer's backing film, a
loose substrate, arid so on.
As to whether any of these defects are critical depends on their
extent, position, and the nature of the applied stresses. Their presence
is more likely to be indicative of poor joint manufacture than of an
impending failure site, especially for short-term loading. Over the
long-term, these defects may allow faster ingress of water or aggressive
substances, or provide the sites for fatigue crack nucleation. It still,
therefore, remains necessary to look for these defects. Wang et al.
(1972) used epoxy-bonded aluminium alloy lap joints with a disbond in
the centre as shown in Fig. 92. The disbond was achieved by inserting a
polypropylene disc in the central region of the joint. Even though
there is a large 'defect' present, the joint strength is little changed.
The other major form of joint used with structural adhesives is the T
joint used in bonding honeycomb to the skins as shown in Fig. 93. The
skin and core are bonded by a large number of these lightly-stressed
but improbable-looking joints. It is essential that the adhesive forms a
generous fillet, Fig. 93(a), and not the apparently more economical but
136 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

8
Z
-'L

~6
+'
01
C
CI
L
1ii4
L ~Xy
'" 2
CI
.c:
11)

..!2
.-~-.
.iij
c
:?-
o

FIG. 92. Variation of tensile shear strength with bond area for single-Iap joint
specimens with adebond (polypropylene disc) inserted (after Wang et al.,
1972, by permission John Wiley & Sons Inc.).

weaker joint shown in Fig. 93(b). Some years ago, one of the authors
(R. D. Adams) tried to stiffen an aluminium alloy honeycomb core by
filling with a foamed-in-situ polyurethane foam. This was then scraped
down to the level of the honeycomb and the bond made. Unfortu-
nately, the foam prevented the formation of the fillet and the joint
failed at a low load.
With bonded honeycomb structure, the major defects consist of a

Honeycomb
(a) Good

Honeycomb
(b) Bad

FIG. 93. Good and bad bonds between honeycomb and skin.
STANDARD MECHANICAL TEST PROCEDURES 137

I /DiSbOnd-COj e damage

)
( /

(a)

\~ ,--I----'

Disbond-skin
imperfection

(b)

Disbond-Iack of
adhesive

(C)

FIG. 94. Defects with honeycomb core structure.

lack of attachment between the core and the skin. This may be due to
several causes such as locally crushed honeycomb (Fig. 94(a)), skin
defects (Fig. 94(b)) or lack of adhesive (Fig. 94(c)). In themselves, none
of these defects may prove deleterious to the short-term joint strength.
However, as for the lap joint, they may show poor preparation and
may provide sites for fatigue crack propagation.

Tests Carried Out Before Bonding


These are principally surface inspection tests based on the need for a
properly-prepared adherend surface. The adhesive properties of the
surface may be poor if there are present excessive amounts of water
vapour, hydrocarbons or other contaminants.
A simple test involves the wettability of the surface, which is a
subjective measurement of the contact angle. If the surface is clean, it
is readily wetted and a drop of water will spread over a large area.
However, if it is contaminated, the water will remain as droplets (c.f.
the waxed surface of a motor car which has been freshly polished). For
example, a useful procedure with carbon fibre reinforced plastics is to
rub the surface with a fine 'wet and dry' carborundum paper until a
138 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

thin film of water can be spread over it. A more quantifiable test
involves measuring the spread of a liquid drop of constant volume
through a transparent gauge placed over the drop.
The Fokker Contamination Tester, described by Bijlmer (1978),
uses an oscillating probe to measure the electron emission energy. This
varies greatly with the degree of surface contamination, and can even
be used to detect residues from alkaline c1eaning operations.

Post-bonding and In-service Testing


By far the majority of non-destructive testing (NDT) techniques as-
sociated with adhesive bonds take place after the joint has been made,
and can usually be carried out at the manufacturing point or during
service. Most techniques are void detectors and, although it is c1aimed
that the cohesive strength of the adhesive is being assessed, this is
unlikely to be the case.

Ultrasonics
Various techniques based on ultrasonics are used in connection with
adhesive joints. Essentially, a small piezoelectric crystal is pulsed
repetitively, causing it to radiate high-frequency sound waves at the
natural frequency of the crystal. Normally, these brief wave trains
(about 5 cycles long, at 1-20 MHz and repeated say 1000 times per
second) traverse the joint and can be detected, either by using the

o
same crystal or by using a separate receiver (Fig. 95). By interrogating

S.'d"

I
~ WaveztralnS

! ~t
o Rezcezlvezr

FrG.95. Ultrasonic through transmission.


STANDARD MECHANICAL TEST PROCEDURES 139

the change which has taken place in the wave train, it is possible to
deduce various characteristics of the structure through which it has
travelled. The probes have to be correcdy aligned and care taken that
the pulse is not reflected or diffracted by curvatures in the surface. If
the adhesive contains porosity, the waves will be scattered and less will
be detected by the receiver. If there is a void, the wave can be partially
or totally reflected, again reducing what is transmitted. Fine detail can
be revealed only by using focussed or otherwise concentrated 'beams'.
Poor quality or damaged transducers can give erroneous results. For
this reason, they should be regularly inspected.
Good coupling between the transducers and the structure is essen-
tial. This can be achieved by pressing them together with a film of
glycerine or some similar coupling agent between them. One develop-
ment is to use a rubber wheel which can be rolled over the surface of
the joint. But perhaps the most common technique in use is the
C-scan, shown in Fig. 96. Coupling is achieved by totally immersing
the structure in water. The pulse is reflected by a smooth surface (glass
plate) so as to return to the probe (used in the send/receive mode).
Sometimes, a second receiver probe is used instead of the reflector, the
two probes being linked by a yoke. This is often necessary with
honeycomb structures since the returning signal is very weak and is
swamped by the much more powerful reflected signals.
The probe is traversed automatically over the structure and the
amplitude of the signal associated with transmission through the struc-
ture is indicated on an X-Y or similar recorder. The X and Y
co-ordinates correspond to the position on the surface, while some
other parameter is used to indicate quality. One method uses different
tones to indicate the strength of the received signal (i.e. joint quality),
while another uses a modification to the pen X or Y coordinate to
indicate a change in the signal.

Probe (sender and receiver)

Water

~~;::;:;:;:;:;;:;:;:;::;:t.,'"
, , ,'. L--_-.;---Structure
\ .-./,1 ,,~~l+=~::::~~ .~ \ -~ ~I "
Reflector
,: ~i;;:=_==,=,,~.,=,~,,=_=,=,=.=,=,.=.,=,,~,,=,=,=..::;;:;j3H'.71,1 plate

FIG.96. Principles of ultrasonic C-scan.


140 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

A different principle is used in ultrasonic resonance testing. Here,


thickness vibrations are induced in the adherend/adhesive/adherend
sandwich in which the two adherends are considered to act as rigid
masses while the adhesive is an almost massless spring. If the type and
thickness of the metal sheets is constant, then the resonant frequency is
a function of the thickness and bulk modulus of the adhesive. The
well-known Fokker Bond Tester (Mark II) overcomes the difficulty of
the many variables in the ultrasonic resonance test by coupling the
joint to a system of well-defined resonance characteristics. Changes in
the frequencies of the joint-transducer system are correlated with
known defects and a calibration curve obtained. The calibration curves
can then be used to establish acceptance limits which can be (and are
extensively) used for quality control in production.
By injecting short pulses of broad band ultrasound (0·5-15 MHz)
into a bonded panel and using modern digital techniques to obtain a
frequency spectrum in 'real time' of the structural response, it is
possible to observe differences between good and bad structures. A. F.
Brown and co-workers (Lloyd and Brown, 1978) have shown that
certain phenomena, such as water absorption, can be readily detected.

Other Acoustic Methods


The coin-tap technique is one of the oldest NDT techniques and it still
has its supporters. The sound emitted by tapping over a well-bonded
section is different from that over an unbonded area, the change
indicating voids, lack of bonding, and so on. Unfortunately, it is a
subjective test, but there have been some attempts to quantify the
result by using an electromagnetic tapper, measuring the response of
the structure with a microphone or piezoelectric transducer, and
analysing the frequency spectrum. Similar devices, such as the Har-
monie Bond Tester (Botsco, 1968) use eddy-current coupling to pulse
the surface.

Acoustic Emission
Application of stress to a material will eventually lead to microscopic
fracture or slip. This is usually associated with a local release of energy
which propagates as a stress wave. The wave has a high frequency
content and is referred to as 'acoustic emission' which can be detected
either by a high-frequency microphone or by a piezoelectric trans-
ducer. Unfortunately, it is necessary to stress the joint to a high
proportion of its failing load in order to generate sufficient emissions,
STANDARD MECHANICAL TEST PROCEDURES 141

and there is doubt whether this is practieal in most cases. It is therefore


not recommended for general use although there are occasions where
it may be effective.

Thermal Methods
By heating one surface of a bonded sandwich structure and observing
the temperature rise of the opposite face, areas of debond, whieh resist
the transfer of heat, show as cool areas. Alternatively, if the heated
face is scanned, debonds will show as hot areas. Temperature sensing is
normally done with a scanning infra-red camera (e.g. AGA Thermovi-
sion). More recently, heat pulses or moving heat sources have been
used (Vavilov and Taylor, 1982). Temperature sensitive paints or
liquid crystals, and thermoluminescent coatings are also used.
An alternative is to cause the structure to vibrate at one of its
resonant frequencies such that defective locations may show by frie-
tional heating, thus leading to a local rise in temperature (Pye and
Adams, 1981).

Radiography
Conventional X-ray techniques are of little use on metal-to-metal
bonded joints since the polymerie adhesive is much less dense than the
adherends. Metallic fillers can be used to enhance the contrast and
show tapering or voids. However, the density of fibre reinforced
plastics adherends is of a similar order to the adhesive and so X-rays
can be used, by choosing a suitable energy and flux. For honeycomb-
cored panels, X-rays are used for checking the position of the core and
whether it has been locally crushed or otherwise damaged.
A new technique is to use a thermal neutron source, since neutrons
are absorbed or scattered by hydrogenous materials such as hydrocar-
bons. Unfortunately such sources are typically nuclear reactors and
accelerators, and so are neither cheap nor readily available!

Optical Holography
By using holographie interferometry, it is possible to measure surface
displacements to less than 0·5 /-Lm. Load is applied to a structure by
vibration, vacuum cup, pressure or heat. Defects show as local pertur-
bations in the holographie interferogram. The technique has found
application with sandwieh structures but not lap or similar joints.
142 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Recent Reviews of NDT Methods


Two excellent reviews have been published recent1y, and readers are
referred to them for a fuller account and references of non-destructive
testing methods of adhesive joints. They are by Schliekelmann (1979)
and Segal and Rose (1980).
Chapter 4

The General Properties of Polymerie Adhesives

There is a lot of evidence that the strength of a properly made


adhesive joint is directly related to the strength of the adhesive with
which it is made. Not only does the failure pattern indicate that this is
so (see Chapter 2) but there are also thermodynamic arguments which
indicate that very special conditions are required for failure to take
place exactly along the interface joining adhesive and substrate. The
properties of joints also depend on the modulus of the adhesive and
since in the case of metal or carbon fibre adherends their modulus will
be much higher than that of the adhesive, any displacement under load
willlargely be due to strain in the adhesive. All this might be regarded
as commonplace if the adhesive could be regarded as behaving like a
metal but with a rather low modulus and strength. Unfortunately the
properties of adhesives arise from very different atomic and molecular
structure. Their strength and elasticity arise from molecular rather
than, as with metals, atomic interactions and both are sensitive to
temperature over a range that would leave the structural met als
virtually unaffected. Only fracture propagation in them can be re-
garded as exactly analogous to the process occurring in brittle metals
and even here toughening mechanisms are very different.
The scale of these differences is indicated in Table 7. In this table
Young's modulus is quoted, though for polymers the tensile stress-
strain relation is usually far from linear and the strain over which
linearity can be assumed is frequently small compared with the strain
at break.
Briefty, the modulus of metals arises in the forces required to
separate atom from atom whereas that of polymers arises mainly from
the forces required to move molecules with respect to each other and
143
144 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 7
COMPARATIVE STIFFNESS AND STRENGTE OF STRUCTURAL MATERIALS*

Specijic
Material Relative Young's Specijic Shear shear Tensile Specijic
density modulus modulus modulus modulus strength strength
(GN.m- 2 (GN.m- 2 ) (GN.m- 2 ) (GN.m- 2 ) (MN.m- 2 ) (MN.m- 2 )

Mild steel 7·5 210 26·7 80 10·7 400 53-3


Brass 8·3 100 12·0 40 4·8 300 36·1
Aluminium 2·6 70 26·9 26 10·0 550 212
(max for
alloys)
Wood (spruce 0·7 14 20 100 143
along the grain)
Epoxy adhesive 1·2 4 3·3 1·4 1·2 50 42

* The figures, taken from various reference books, are illustrative only.

parts of molecules with respect to other parts. In the case of the


rubbery polymers the force arises to a large extent from the entropy
contribution made by a configurational change rather than a change of
internal energy. This results in a relatively low modulus which in-
creases with increasing temperature, unlike moduli which arise from
interatomic or intermolecular forces.
The tensile strength of a freshly drawn quartz fibre approaches its
theoretical strength which is allied to the work of cohesion; that
required to separate the groups of atoms from each other. Metals,
under ideal conditions, would yield and flow and hence fail in shear as
planes of atoms slide over each other. In fact, conditions being always
less than ideal, the true cohesive strength of the metal is far from being
realized and brittle fracture often occurs before the specimen yields.
Yielding, with one plane of metal atoms shearing with respect to
another involves strong interatomic forces. Polymers similarly fail
under stress either by a ductile or a brittle process. In the former only
intermolecular forces (far less strong than interatomic forces) are
involved and in the latter a mixture of intermolecular forces with a
very small proportion of interatomic forces (as some of the long chain
molecules are broken between one carbon atom and another). Which
process occurs depends partlyon the stress state as weIl as the rate at
which it is applied, partlyon the temperature, but mainly on the nature
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 145

of the polymer. The environment also may cause a change from a


brittle to a ductile mechanism of fracture.

POLYMER STRUCTURES

The structure of a polymer is formed by taking a small group of atoms


and using numbers of such groups to build a large molecule. The
building process may join up groups into a very long chain-like
molecule or it may so join them that a random process is at work and
a large three-dimensional, interconnected structure results. There is a
very wide range of polymers which is used in adhesives and there are
various ways of classifying them. Table 8 gives the more important
adhesives likely to be used in engineering practice and classifies them
according to the type of structure on which the polymer is based in its
final condition, i.e. after it is 'cured' if curing is essential to its
performance.

Unsaturation
Many of the chemicals referred to in Table 8 depend for their
reactivity on unsaturation or the occurrence of a double bond between
two carbon atoms instead of a single bond. An unsaturated centre in a
molecule usually allows the molecule to unite by simple addition with
identical molecules or molecules different in other respects but similar
in containing unsaturation. The polymerization of vinylacetate, methyl
methacrylate, chloroprene, cyanoacrylate, diacrylate and styrene are
all of this type. A second method of building larger moleeules from
smaller ones is by condensation. In this process two molecules unite
with the elimination of a small, simple molecule, usually water. Thus
the formation of the polyester occurs by the difunctional acid and
difunctional alcohol reacting to form a long chain through the elimina-
tion of a molecule of water at each point of union. The reaction of the
various phenols and amines with formaldehyde are also of this type.
The polyester res ins familiar as the matrix in glass reinforced plastics
involve both type of polymerization. The polyester is formed by a
condensation re action but because some of the acid used possesses
unsaturation, it is able to take part in direct addition with another
unsaturated body. The long chain polyester is then dissolved in
styrene, which contains unsaturation and its polymerization takes in, as
....
~
0\

TABLE 8
THE CHEMICAL STRUCTURE OF SOME POLYMERS USED AS ADHESIVES

Monomer or groups 0/ atoms J Polymer structure (be/ore curing) Final structure


etc., curing agent if needed

Lang regular chain


~
Simple monomer Unchanged
e.g. vinyl acetate Polyvinyl acetate (PVA)
(CH2=CH.OCOCH3l -CHz-CH- ]
[
;p
bCOCH3 n
~
r
Methyl cyanoacrylate Poly cyanoacrylate ~
CH2=C(CN).COOCH3 CHz-T(CN)- ]
[ @
COOCH3 n r:/l

~
Simple monomer Lang regular chain Three dimensional, cross~linked by metal oxide OI sulphur. o'-<
Metal oxide or sulphur
e.g. Chloroprene Polychloroprene There will be one cross-link between any two chains every, ~
(CHz==CCl-CH=CH2) [-CH z-CCl=CH-CH2-]n say, 100 carbon atoms. Z
~
More camp/ex monomer with functionality separated ~ Three dimensional structure with each monomer unit ~ Unchanged
to bOlh ends 0/ a Zarger moleeule joined to two others at each end
~
tTl
tTl
e.g. Diacrylate Anaerobic adhesive
Epoxy resin CHZ-CH-[Ar-CH-CH2] -Ar-CH-CH2 Cross-linked by -reacting terminal epoxy groups with
~
Q
e.g. Diglycidyl ether of bisphenol A '""'0/ I "./ polyamines cr epoxy groups and hydroxy groups with
OH n 0 acid anhydrides. There are other cross-linking agents.
Ar represents the complex group derived from bisphenol
A. The monomer itself occurs in admixture with poly-
mer. n being small. Sometimes the monomer. which
can be same ether diglycidylether is reacted directly
with curative.
Prepolymer and small moleeule Three dimensional, cross-linked structure
e.g. Unsaturated polyester + styrene monomer Polyester resin
Saturated\ polyester + dllsocyanate or polyisocy- Polyurethane
anate
Rubber polymer of high M.Wt. dissolved in acrylic Modified or toughened acrylic
...,
monomer or mixture of monomers. Accelerator @
applied separately to one adherend.
filz
m
Condensation of small moleeules to resinous adhe- ~ Reactive stage liquid or powder Fully reacted inert mass
sives ~
t""'
e.g. Phenol + formaldehyde } Stage 'A' phenolic resin Stage 'C' resin '"0
Resorcinol + formaldehyde :;.:I
Urea + formaldehyde
o
ri'l
Mixed polymers Three dimensional, cross-linked structures
Epoxy-nylon } Epoxy cross-linked by reaction with othcr polymer somc-
~
m
(Il
Epoxy-polysulphide times in thc presence of other curatives.
Epoxy-phenolic o'Tl
Phenolic-nitrile Thc 'nitrile' is an acrylonitrile-butadiene copolymcr or '"0
nitrile rubber. It reacts chemically with thc phenol for- ot""'
maldehyde rcsin which is present in its 'B' stage.
Phenolic-polyvinyl- } As with phenol-nitriles, a chemical reaction may occur hut
formal is less likely. The PF component converts to a 'c' stage
~~
acetal resin. (")
butyral :.-o
@
(Il

~
(Il

......
~
-.J
148 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

it were, the unsaturation of the polyester and binds the whole into one
gigantic molecule.

MIXED ADHESIVES

The dassification, mixed adhesives, covers the dass of commercially


available materials prepared from two major components both of
which are adhesives in their own right and could function without the
other component. In many cases however, there is chemical reaction
between the two, one component frequently acting to cross-link the
other. The very commonly used phenolic-polyvinyl formal or butyral is
probably a simple two-phase mixture of a phenol formaldehyde resin
and a vinyl polymer. On the other hand the phenol formaldehyde resin
in the nitrile-phenolic composition undoubtedly reacts chemically to
form cross-links in the nitrile rubber although, again, a two phased
structure exists. The importance of one discrete phase uniformly
distributed as microscopic or even sub-microscopic particles in the
other, continuous phase relates to fracture behaviour. Normally, the
volume fraction of the dispersed phase must be less than that of the
continuous phase. As the volume percentage of a dispersed phase is
increased, instability occurs during either the preparation or curing
process and when more than 50% is added, phase inversion occurs and
the formerly dispersed phase becomes the continuous phase. Most
examples involve a somewhat brittle continuous phase and a more
rubbery dispersed phase. In such structures the energy absorbing
rubbery phase acts as a crack-stopper in the propagation of fracture
and hence toughens the material.
The most recent development in this field of polymer science is that
of interpenetrating polymer networks (IPN) in which the precursors for
two types of polymer are mixed and polymerize without phase separa-
tion so that their network structures interpenetrate each other. This
has not yet reached commercial exploitation although the Russians
have experimented with interpenetrating polyester and polyurethane
networks for repairing oil pipelines (Lipatov et al., 1979).

PROPERTIES AND TEMPERATURE

The loss in strength, lowered modulus and enhancement of creep


during service at the very high temperatures to which gas turbine
THE GENERAL PROPERTIES OF POLYMERIe ADHESIVES 149

blades are subjected, places a technological limit on the efficiency and


life of the turbine. At the other end of the temperature scale mild steel
is prone to brittle fracture in the Arctic. Within these extremes
however, engineers are but little concerned with any major change in
properties of metals as a result of temperature change. There are, by
contrast, several important temperatures which characterize the
polymers used as adhesives and through which change is marked.
These are the decomposition temperature, the melting temperature,
the crystal melting temperature and the glass transition temperature
although only regular linear polymers will show all four. Because the
molecules of a polymer are not of identical weight-there is a distribu-
tion of weights either of the individual molecules or, if there is a
network structure, of the weights between network junctions-these
characteristic temperatures will exhibit a range rather than the sharp
melting point of a crystal of a simple organic compound.
All four temperatures exist within a moderate range of temperatures
over which the properties of the common metals vary only in degree
and not in kind.

The Glass Transition Temperature


The most important temperature is the glass transztwn temperature,
usually written Tg and which is exhibited by all polymers. With changes
of state from crystal to liquid there is a discontinuity in any plot of a
fundamental quantity such as volume, free energy, or entropy against
temperature. On melting or freezing the volume for instance, changes
abruptly. The glass transition is not a first order transition in the sense
in which melting iso There is no abrupt change in any quantity but
there is a change in the derivatives of the fundamental quantities with
respect to temperature. Although the exact nature of the glass transi-
tion is still a matter for argument among thermodynamicists, it suffices
here to refer to it as a second order transition because of the change in
the derivatives. There is a marked change in the mechanical properties
exhibited above and below the glass transition and in the neighbour-
hood of the transition, the mechanical energy loss, i.e. the imaginary
part of the complex modulus (vide infra), shows a maximum. Above
the glass transition the material, if it is not crystalline, will be rubbery.
If the polymer is linear and is very regular, it will crystallize above its
glass transition temperature and hence the latter will be masked and
the mechanical properties will reftect those of the crystalline material
rather than those expected of a polymer above its Tg • An amorphous
150 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

polymer above its Tg extends because the long chain molecules which
are more or less coiled, uncoil. This change in form need not involve
internal energy, as would be involved if the distance between atoms in
the chain were altered. Below the Tg , the molecular motion which
permits this change in coiling is frozen and extension then involves
intern al energy, the material becomes glass-like and relatively inexten-
sible. The glass transition temperature is greatly inftuenced by struc-
ture. Bulky molecular groups such as constitute aromatic molecules or
groups with high intermolecular attraction, as is associated with an
electronic structure giving permanent dipole moments, lead to high Tg •
The hydrocarbon straight chain rubbers, devoid of polar groups, show
low Tg . The lowest Tg is associated not with carbon chains but with the
repeated silicon-oxygen-silicon structure of the polysiloxanes or
silicone rubbers. These remain rubbery down to very low tempera-
tures. The glass transition temperature of polydimethyl siloxane is in
the neighbourhood of -12Soe but as considerable crystallization exists
below -sooe, it therefore loses its rubbery properties and 'brittle
points' higher than -120o e are recorded.
The glass transition temperature, depending as it does on molecular
motion, is dependent on the rate of testing when the measurement is
one involving mechanical deformation. The effective glass transition
for a polymer undergoing cyclical deformation is therefore frequency
dependent. It is also possible to shift Tg downwards by the use of
suitable plasticizers. The results of a rate-dependent test on a polymer
at two different temperatures at one rate can be duplicated by per-
forming the test at one temperature with two different rates. Thus,
there exists a time-temperature equivalence which is of particular
importance in dealing with cyclical deformation. Structural adhesives
are used in rigid structures and need themselves to be of as high a
modulus as possible. This means using them below their glass transi-
tion and since, in general, they will be used at air temperature, the Tg
must be elevated above this by the introduction of polar groups or
other structural features. On the other hand, adhesives that are to be
applied as aqueous emulsions must form films when the water evapo-
rates. Polyvinylacetate, as an uncompounded polymer, has Tg equal to
28°C. An emulsion of this material would dry to a powder-not a
coherent film. A plasticizer must therefore be present to reduce the
glass transition to below room temperature and if the plasticizer is so
chosen that it is slowly lost into the atmosphere after the joint is
formed, the Tg will slowly rise and the properties of the joint improve.
THE GENERAL PROPERTIES OF POLYMERIe ADHESIVES 151

Decomposition Temperature
Many polymers simply revert to the monomers from whence they
derive if the temperature is sufficiently raised. Others, particularly in
the presence of air, char and give off gases. In the former case,
obviously the adhesive and therefore the joint strength, vanish. In the
latter case, the carbonaceous char remaining may enable a joint to
sustain a load for a sufficient period of time to serve its purpose.
Obviously, this load will be less than the initial strength of the joint.
Some military projectile applications depend on char strength remain-
ing, perhaps for seconds, to complete the mission. Table 9 gives some
typical decomposition temperatures. As is recorded later, continuous
exposure to temperatures much less than the decomposition tempera-
ture causes deterioration and shortens joint life. The table implies a
fairly rapid rise to the recorded temperature.

TABLE 9
DECOMPOSITION TEMPERATURES

Adhesive polymers Decomposition Adhesive Decomposition


temperature polymer temperature
caC) (Oe)

PolymethyJ methacrylate 180-190 Silicone 135-240


Epoxies 150 Polymides >500
U/F 200 Polysulphones 500
Nitrile rubber 300
Polychloroprene 310
P/D 300

Melting Temperature
Amorphous polymers of low molecular weight melt, though it is
difficult to state at what temperature above their glass transition they
really are behaving as liquids. If such materials are cross-linked, that is
the essentially linear chains are fastened together at points more or less
widely spaced along the chains, the behaviour exhibited is that of a
rubber and melting is inhibited, alteration to a liquid on heating
indicating some decomposition. Crystalline polymers will show a defin-
ite melting temperature. Cross-linking of these essentially linear chains
will prevent crystallization, and hence rubbery behaviour will be shown
both below and above the crystalline melting point. Some crystalline
152 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

polymers on melting show, over a limited temperature range, some


rubbery behaviour before becoming truly liquid. Both the softening
point of amorphous polymers and the melting temperature of crystal-
line ones can be depressed by plasticizers or by admixture with bodies
of low molecular weight but of sirnilar chemical structure. An example
of this is the use of waxes to lower the melting point of polyethylene in
the production of certain hot melt adhesives (HMA). Hot melt adhe-
sives are rarely the concern of the design engineer though they may
become so in the future. He should, however, be aware of them as
some products, whilst not structurally load bearing, are assembled with
them. The edge veneering of block-board panels and doors is an
exarnple. Additionally, newer polymers such as the polysulphones may
be used in HMA formulations in engineering applications, as their
tensile creep modulus at room temperature is approxirnately constant
above 2 GN.m- 2 over aperiod of 1 year under a load of 28 MN.m- 2
(Anon. 1973). Even when the (crystalline) melting point can be pre-
cisely defined, the conditions for its determination make it more
convenient to identify a softening point by some arbitrary test which
has a meaning within its applicational frarnework. This point is vari-
ously known as a softening point, a heat distortion, or defiection
temperature. The heat defiection temperatures quoted (Anon. 1974)
for three polysulphones obtained by loading a bar at 1·8 MN.m- 2 were
174, 274 and 203°C for Union Carbide's 'UdeI', 3M's 'Astrel' and
ICI's polyethersulphone respectively.
In view of these very profound changes occurring for all the poly-
mers of interest as structural adhesives over the extreme range of
-125°C (Tg of dirnethyl siloxane) to 500°C (upper limit for short-term
use of polybenzimidazole) and for most over the range, say, -15°C-
150°C it will be obvious that substantial changes in the mechanical
properties of modulus, strength and creep are to be expected with
quite moderate changes in temperature.

Tbe Deformation of Adhesive Polymers by Stress


Structural adhesives are used at temperatures below or at worst near to
their glass transition temperatures. The stress properties of interest are
therefore those of the material in its glassy or crystalline condition.
The characteristic of long chain amorphous polymers above their glass
transition, namely their S-shaped stress-strain relation and very high
extensibility, sometimes more than 100%, will not be discussed.
Nevertheless, in terms of metal behaviour, the extension from which
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 153

recovery is virtuaHy complete, can be relatively high, say, 5% although


there may be considerable delay in reaching complete recovery. Exten-
sion at break of glassy polymers can be as high as 100% but this is very
sensitive to the extension rate.
Measurements made on adhesive joints show a rather more limited
range. Shen and Rutherford (1972) working with a nylon-epoxy struc-
tural adhesive (FM 1000, Cyanamid Corp.) found that stress across the
bond, i.e. the bond being stressed in tension, gave linear elastie
response up to an extension of 0·3% and they report a yielding at
0·7%. These figures are deduced from the stresses and modulus given
in the paper. The film thiekness was 0·2 mm and the area between the
adherends approx. 1·33 cm2 . The authors did not check recoverability
from the 'yielded' strain so that they could have mistaken a delayed
elastic response for an irrecoverable yield.
Because adhesives are used in very thin films and shearing is in the
plane of the film, the shear strain can be large. Shanahan (1974)
measured shear strains of up to 50% whilst the adherends of a
lap-shear joint were creeping under load. This corresponds to a very
substantial shear angle of 26·5°. The adhesive was a structural adhesive
of the polyvinylformal/phenolic resin type (Redux 775, Ciba-Geigy,
Ltd) and had a shear modulus of about 1·0 GN.m- 2 .
It is necessary therefore to consider the deformation of polymers as
affected by temperature and rate of loading as weH as of the stress
imposed. Unlike the elasticity of the rubbery state, for which a
reasonable theoretical treatment exists dependent on fundamental
physical constants and molecular concepts, the glassy state can only be
treated by a phenomenological approach.

Viscoelasticity
The response of a metal wire to an imposed stress is an instantaneous
extension proportional to that stress. On removal of stress the length
instantaneously recovers its initial value. The extent of instantaneous
linear response by a polymerie material above its glass transition
temperature is a smaH fraction of its total response to stress but there
is a time-dependent component which is much more important. The
models used to represent phenomenologieally this time-dependent
component are based on arrangements of springs and dashpots and are
associated with the names of Maxwell and Voigt.
Figure 97 shows the two models and also the slightly more realistie
model obtained by one method of combining them (Fig. 97(c)). For the
154 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

FIG.97. Maxwell and Voigt models of deformation properties.

Maxwell model Fig. 97(a), the strain under stress, S, varies with time, t,
as described by eqn (25):
de 1 1 dS
-=-S+-- (25)
dt 'Y/rn Ern dt

where e = strain, S = stress, 'Y/rn and Ern are constants respectively


identified with the viscosity (of the liquid in the dashpot) and a
modulus Ern (of the spring).
If the strain is held constant (dE/dt = 0) then the stress will decay to
zero and can be obtained by integrating the expression:
1 1 dS
-S+--=O
'Y/rn Ern dt
i.e.
S = So exp {- (Em/'Y/~t} or So exp (-t/Trn)
The ratio 'Y/rn/Em has the dimensions of time and is usually written Tm,
the relaxation time of the process.
The Voigt model, Fig. 97(b) yields a differential eqn (26):
de
S ='Y/v dt +Eve (26)

If this model is subjected to stress which is then removed, the strain


decays at a rate given by eqn (27):
S
e = Ev {l-exp (-thv )} (27)

In this expression Tv is the retardation time.


THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 155

As the two models are representing different physical situations the


springs and dashpots cover different functions and hence the relaxation
time is not the same as the retardation time. In order to reproduce the
characteristic peak in the viscous part of the dynamic modulus (see
Figs 98-102) the model must contain at least three elements, two
springs and a dashpot. With the more simple two element models the
loss modulus is only modelled on one side of its peak value. Real
materials are best modelled by expressions which contain a distribution
of relaxation times though from the viewpoint of the variation of
mechanical properties with temperature, this is adequately expressed
by a single figure representing the mean of a logarithmic distribution of
relaxation times.
If the time during which stress is applied is small compared with the
relaxation or retardation times, the response is determined by the
spring rather than the dashpot. It is readily appreciated that for a
polymer at a temperature below its glass transition the relaxation time
is almost infinitely long and hence its response is substantially elastic
though not linear. As the temperature is raised and Tg approached, the
modulus starts to fall and, in an experiment in which the stress is
applied cyclically, it can be shown that a viscous component becomes
increasingly important.
Although the tensile testing of adhesively bonded joints at room

'E
~ 108 1--------'-----------'---;--;---1
:;
"0
~
::l~
. '"

FIG. 98. Real and complex parts of the dynamic modulus, G' and G", of
Redux 775 (Ciba-Geigy) (after Shanahan, 1974).
156 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

102
\
-
[\
-\
\ ,
!
1 o

I~~I~\ _-. ............


I

,
,,

1~
I
I

.
I
)'J
I

"
j
I

I
I
, I

100 - - .-- -
.- - - ,
-25 o 25 50 75
~.
100 12
<

T<lmp"ratur" «'C)

FIG. 99. Real and imaginary parts of the dynamic modulus, G', and A (the
ratio of the logarithm of the amplitude of torsional pendulum swings) of Redux
775 (after Matting and Draugelates, 1968).

temperature and many uses of adhesive bonding only involve the


adhesive in its wholly glassy state, weIl below T g , there are stages in
production and in the use of joints in which viscoelastic properties play
an important part.
During the manufacture of bonded components with some adhesives
it is necessary to apply them to the faying surfaces which are combined
when the adhesive is substantially free of solvent, partially
polymerized, or in some other intennediate state depending on the
adhesive itself and on the process used. At this stage of combination,
the adhesive is required to make an instantaneous bond which need
not be of high strength but is required for placement purposes. The
ability of the adhesive to make this bond arises from the property of
tack which is a viscoelastic property. Pressure-sensitive adhesives used
in tapes remain permanently tacky but structurally used adhesives lose
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 157

tack as the setting or curing proceeds in building the final and higher
strength of the bond.
A second and possibly more important aspect of viscoelasticity arises
when polymers are cyc1ically deformed. Even at temperatures below
Tg there is still some viscous component which leads to energy loss
during the cyc1ing. Advantageously, this can lead to the damping of
free vibration but disadvantageously, the energy, converted into heat
raises the temperature.
The adhesive will then be operating at some temperature above the
ambient depending on heat losses by conduction away from the joint.

-
5
\

1
-
~'T'I
I
I ,

,
1,5

\\
~10
N I \
,
'E '.
E I
\
\
0,5
"" , I
I ,
\

1,0

\. '\
,

I
I
I
I
I

-- <
I
I
I
\ 0·5
I
I
/
I
5
, "'-
,,
. ,, ~-
2 o
-25 o 25 50 75 100 125 150
T"mp"ratur" (oC)

Fro. 100. Real and imaginary parts of the dynamic modulus, G', and A (the
ratio of the logarithm of the amplitude of torsional pendulum swings) of
Araldite 106 (Ciba-Geigy) cured at elevated temperature (after Matting and
Draugelates, 1968).
158 STRUCTURAL ADHESNE JOINTS IN ENGINEERING

r--

r
I
102
~
5 \
I

,
10 I

,
, \' , I ,
,: ,
0

\
, I

I
I
,,
l -

.' <
~
/
,I
I
I
, i"-... ,, -,
'-' "-.. , ,~

,
I

1\\
I
I

,
- -- -
5
~
3 ~O
-25 o 25 50 75 100 125 150
T<2mp<lraturtZ ("C)

FrG. 101. Real and imaginary parts of the dynamic modulus, G', and A (the
ratio of the logarithm of the amplitude of torsional pendulum swings) of
Araldite 106 (Ciba-Geigy) cured at room temperature (after Matting and
Draugelates,1968).

In theory, the rise in temperature may affect the modulus and the
endurance of the joint but, in practice, the relatively large mass of metal
and small volume of adhesive with large interfacial area gives optimum
conditions for the removal of heat. In the case of metal-to-metal joints,
the temperature rise on dynamic deformation is negligible but this cannot
be assumed for other substrates such as carbon or glass fibre compo-
sites for which data do not seem to have been published. If a material
is deformed sinusoidally in apparatus capable of recording both stress
and strain, it will be found that there is a phase difference between
them; the strain lagging behind the stress. Their ratio gives the
complex modulus G* which can be separated, on the analogy of AC
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 159

(/)
::J
::J

"8 1dr---+-+--t------t--1
L:
c

FIG. 102. Real and irnaginary parts of the dynamic rnoduli at 110 Hz: A
polyethylene; B ethylene-acrylic acid copolyrner, 8%; C ethylene-acrylic acid
copolyrner, 15% (after Wargotz, 1969; data frorn Octocka and Kwei).

theory, into real and imaginary parts as in eqn (28):


G*=G'+iG" (28)
where G' is the real part, sometimes called the storage modulus, G"
the imaginary or viscous part and i is the square root of -1. G' is, in
fact, the ratio of the stress to the strain which is in phase with it and is
a measure of the recoverable energy. G"/G' gives the tangent of the
loss angle, i.e. the phase angle separating the maximum strain from
maximum stress. G" is also obtainable from the logarithmic decrement
shown by a torsional pendulum in which the polymer serves as the
elastic element. Figures 98-102 give dynamic modulus data for a
number of adhesive polymers an obtained. with a torsional pendulum
but whereas Fig. 98 (Shanahan, 1974) plots G" as wen as G', Figs
99-101 plot the logarithm of the ratios of the amplitudes of successive
swings of the pendulum (Matting and Draugelates, 1968). Figures 98
and 99 refer to Redux 775 (Ciba-Geigy). Figure 102 refers to an
ethylene-ethylacetate copolymer (Wargotz, 1969) used as a hot melt
adhesive (HMA); it is not used as a structural adhesive but is included
because of the high modulus shown at low temperatures. Additionally,
the dynamic modulus was recorded at 110 Hz and not at the natural
frequency of a pendulum system.
160 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Tbe dynamie properties of the polymer used as the adhesive are one
of the prineipal determinants for the fatigue life of a joint subjeeted to
oseillating stress, the other determinants being temperature, joint
design and the nature of the adhesive; particularly its glass transition
temperature. Matting and Draugelates (1968) state that elevated
operating temperatures as well as low frequeney vibrations reduee the
vibration al stability of adhesive joints. Tbe loss function only beeomes
important at high temperatures and low frequencies and it is this,
apparently, whieh disturbs the stability of the system.
The Modulus 01 an Adhesive
The equilibrium stress-strain eurves of polymerie materials depart
eonsiderably from linearity in tension but not in shear. Although there
ean be eonsiderable differenees in quoted moduli depending on
whether a tangential or seeant modulus is quoted and to what elonga-
tion these refer, there is mueh less ambiguity in quoting the shear
modulus.
Table 10 is derived, with change of units, from Kuenzi and Stevens
(1963) and gives inter alia modulus figures of adhesives measured in
joints formed by bonding aluminium washers or tubes and subjeeting
them to tensile or torsional stress. Table 11 gives further figures
derived from various sourees.
TABLE 10
MODULUS PROPERTIES OF ADHESIVES (AFrER KUENZI AND S'IEVENS, 1963*)
Adhesive Young's Shear Poisson's Work to
modulus modulus ratio proportionalty
(GN.m- 2 ) (GN.m- 2 ) limit (kJ.m- 3 )

Redux K-6 3-45 1·27 0·36 517


FM 47 2·24 0·81 0·385 241
Epon422 2·72 1·10 0·29 241
Epon V111 3·50 1·24 0·41 379
FM 1000 1·24 0·44 0·41 828
3·79
(Shen and
Rutherford, 1972)
Metlbond 408 0·96 0·34 0·41 241
Scotweld AF-6 0·13 0·04 0·49
0·07 0·02 0·49 10350
Metlbond 4021 0·11 0·04 0·47
0·44 0·01 0·50 35880
Metlbond MN 3C 0·04 0·01 0·50 15180
* Units converted.
~
tI1

TABLE 11 ~
MODULUS PROPERTIES OF ADHESIVES tI1

Adhesive Young's Shear Source ~


>tl
modulus modulus :;0
(GN.m- 2 ) (GN.m- 2 )
~
Araldite A Y18/HZl18 (epoxy/diamine hardener) 2·4-3·1 ~
tI1
SBD Certite 19-19 (filled polyester) 9 Shields (1976a) Vi

Araldite 775 (polyvinylformal/PF res in) 3·3


} ~
Polyvinyl formal 2·9 1·23 Whitney and Andrews (1967) >tl
Epoxy-polysulphide }
o
0.10}
Polyvinyl acetate Wood adhesives 0·35 Krueger (1962)
Phenol-resorcinol-formaldehyde 0·93 ~
Polyvinyl acetate, ceramic tile adhesive 1·7 Counsell (1979) c:
n
Polychloroprene contact adhesives 0·035 Bates (1971) E;
S
~
Vi

......
0\
......
162 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Poisson's Ratio
The two fundamental constants used in elasticity theory are the Lame
constants A and /-L, and the commonly used material parameters can be
expressed in terms of these two constants. However, it is convenient to
introduce into engineering practice a third materials constant, Pois-
son's ratio, v, which strictly is valid for simple stress fields where there
are no shear components and only a single main tensile stress. The
second Lame constant is then equivalent to the shear modulus G
which then becomes related to the Young's modulus by the expression:
E
G=---
2(1 + v)
For polymers, Poisson's ratio varies between the extremes of O· 5 and
0·33 the former value being appropriate weIl above the glass transition
temperature and the latter weIl below it. In practice, this means that
hydrocarbon rubbers exhibit a Poisson's ratio of 0·49 and that resinous
adhesives such as the epoxies show a value of O· 37 or larger which
increases when the adhesive yields.

Strength Properties 01 Adhesive Polymers


The appearance of the adhesive surface after breaking a joint can vary
considerably with the temperature, rate of strain as weIl as with the
nature of the material. Figure 103(a) shows a failure surface from a
joint with a polyvinylformal/PF (Redux 775) broken at very low tem-
peratures (-196°C) and contrasts with Fig. 103(b) broken at room
temperature. The type of failure shown at (a) is almost wholly within
the adhesive film and seems a typical brittle fracture the surfaces of
which have been damaged by grinding together by movement relative
to each other after the break occurred. Figure 103(b) shows evidence
of plastic distortion of an apparently ductile material a lot of which has
been pulled or peeled from the substrate metal. Superficially, the
polymer has in one case yielded before failure and in the other has
failed without yielding.

Yielding Stresses 01 Polymers


The standard criteria for yielding developed in elasticity theory for
metals, namely the Tresca, von Mises or Mohr-Coulomb, do not apply
quantitatively to polymerie materials because these criteria ignore the
effect of the hydrostatic component of the stress tensor. This is
important in determining the yield behaviour of polymers, not neces-
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 163

FIG. 103. Typical failure surfaces from torsional test pieces with Redux 775
(Ciba-Geigy): (a) broken at -196°C to illustrate brittle failure; (b) broken at
room temperature to illustrate ductile failure.

sarily through dilatation but because of a change in the state of the


material caused by stress (Bowden, 1973). This change in state is
probably one in which the glass transition temperature is altered by the
stress.
The joint shown in Fig. 103(b) failed in the neighbourhood of
70 MN.m- 2 (data from Foulkes et al., 1970) whilst the yield stress in
uniaxial tension of polyvinyl formal at room temperature is given by
Whitney and Andrews (1967) as 78 MN.m- 2 . The uncertainty in the
stress failure of torsional test pieces arises from differences in calcula-
tion of stress from failure torque. Either plastic or elastic failure must
be assumed to make the calculation and which is assumed can only be
decided from an examination of the appearance of the failed test piece.
It may, however, be taken that most of the metal-metal adhesives,
and certainly those with reasonable peel resistance, will yield before
failure whilst structural wood adhesives will fail in brittle fracture
without yielding.
164 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

In most situations of testing materials to examine their tensile or


shear properties, a test piece is subjected to continuously increasing
stress. If the material is ductile at the rate of stress applieation, it yields
and the slope of the stress-strain curve shows a change of direction. In
testing polymerie materials the c1amps holding the test piece, whieh is
usually waisted, separate at a fixed rate and hence the curve turns over
with the recorded load levelling off or even falling. With constant rate
of loading, the strain accelerates. In both types of test the yielding
occurs when the true stress reaches a value appropriate to the temper-
ature and rate-of-Ioading or straining. If however, a test piece is
loaded to some lower stress and remains loaded, yielding can occur
after a lapse of time. This is likely to be of importance in the structural
applications of adhesives where stresses are high and a limited life is
allowed for in the design. Indeed, delay times before the onset of creep
have been observed in structural adhesives with lap-shear joints
(Shanahan, 1974. Allen and Shanahan, 1975, 1976).
Delayed yielding in glassy polymers was discussed by Matz et al.
(1972) following earlier work on textile fibres by Coleman and Knox
(1957) and Robertson (1963). The mechanism postulated by all these
workers and used by Allen and Shanahan in work on the creep of
adhesive joints is one based on a rate process for the breaking and
reformation of chemical bonds between molecules. These may depend
on secondary or van der Waals's forces or primary valency bonds with
an equilibrium of the rates in both directions being established appro-
priate to the temperature. The addition of stress biases the rate in the
breakage direction. It has been shown that a delay time for the onset
of deformation can be quantitatively expressed by a function of a type
weH known to physieal chemists involved in reaction-rate theory and
dependent on the theories advanced by Eyring (1936) to explain
chemieal reactions. Matz et al. (1972) showed that for polycarbonate,
polymethylmethacrylate and polysulphone, log (delay time) is linearly
dependent on stress above a certain limiting stress, below which the
delay time approaches infinity. Allen and Shanahan observed similar
relations with lap-shear joints made with a polyvinylformal/PF adhe-
sive (Redux 775) and a modified epoxy novolac (BSL 906).

Failure Modes After Yielding


Because of the attrition which occurs if the joint members slip past
each other during failure the appearance of a joint which has broken
after yielding of the adhesive may not be very different from one that
THE GENERAL PROPERTIES OF POLYMERIe ADHESlVES 165

has failed by purely brittle fracture. Failure in direct tension affords a


better chance of examining the nature of the fracture than the more
usuallap-shear test piece although with the relative absence of a shear
component the requirement for yielding will be different.
Figure 103(b) does show evidence of the failure mode and micro-
scopic examination of the substrate after failure also showed that
apparently clean substrate was, in fact, covered with a very thin layer
of adhesive. Despite the appearance to the naked eye, failure was
wholly within the adhesive layer.

Creep
Failure following crazing is not the inevitable consequence of yielding
under stress. Another possible response is aperiod of continuous
deformation. Under conditions of shear, the only situation that has
been studied for adhesive in detail, continuous creep is preceded by a
time delay before the continuous deformation commences. Most atten-
tion has been paid to the linear polymers such as polyethylene,
plasticized as well as unplasticized polyvinyl chloride and polymethyl-
methacrylate none of which are used even as semi-structural adhesives.
It is, in fact, the absence of cross-linking which allows for extensive,
continuous deformation and, at the same time is one of the major
reasons for neglecting them. The only cross-linked material for which
creep has been extensively studied is natural rubber and, being above
its transition temperature it, too, is not of interest in the present
context.
The strain behaviour of the linear thermoplastic polymers under
continuous load is governed by time, stress and temperature. If tem-
perature is ignored and discussion limited to one fixed temperature
there are left two variables the effects of which are only separable at
very low strains (Turner, 1973). This will readily be appreciated from
the use of an equation of the form:
e = a sinh bS + ct" sinh dS (29)

where e = strain, S = stress, t = time and a, b, c, d are fitted constants


characteristic of the material being investigated.
The form of this equation is one of the few suggested relations
between the variables which has some basis in theory and depends,
identically with the time delay before yielding already discussed, on
Eyring's theory of rate processes. When strain is plotted against log
(time) equations of this type give fairly Hat curves at low stresses and
166 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

curves which show a moderate rise followed by a steeper upturn when


the stresses are higher. In fact, the upturn after aperiod of creep is
frequently sharper than predicted and leads to stress rupture. The
period of more rapidly increasing strain may be due to a further
yielding with the formation of voids (Benharn and McCammond,
1971).
The creep referred to in the previous paragraph is a relatively long
term phenomenon and is the deformation which is wholly irrecovera-
ble and non-equilibrium. The term is, however, sometimes used for the
delayed elastic response which all high polymers show but which, if the
temperature is in the neighbourhood of the glass transition tempera-
ture, may proceed so slowly as to occupy many minutes. The shape of
this 'creep' curve when strain is plotted against log (time) is linear for
any given stress but with the slope of the line increasing with increasing
stress. It passes smoothly into true creep as departure from linearity
occurs.
There appears to be no information on the uniaxial creep of poly-
mers used as structural adhesives such as is available referring to the
creep of adhesive joints in lap-shear or torsion. The latter is reserved
for Chapter 7 where the few data that are available are given. An
apparent exception is the careful study of a nylon-epoxy adhesive
(FM 1000) by Shen and Rutherford (1972). These authors achievedl
something approaching uniaxial stress using a cylindrical butt joint in
direct tension. However, what they called 'creep' was simply the
delayed elastic response. They likened the adhesive to a metal in its
behaviour and used the classical but inappropriate concept of separat-
ing the creep behaviour into three stages as was done many years aga
for metals.

Failure without Yieldiug-Brittle Fraeture


Highly cross-linked, resinous polymers fail under stress without the
stress-strain curve showing any point of inflexion or other indication of
a· yielded condition. Although the uniaxial stress-strain curve may be
far from linear, the stress never falls whilst the test piece is intact, i.e.
the break occurs at the maximum load. Stress-strain curves of resins in
shear are linear and proceed linearly to the breaking stress. Only when
flexiblizing agents have been added does departure, from linearity
occur. The surfaces of the broken test piece show typically conchoidal
facets with some fragmentation arising from multiple cracks.
The theory of brittle fracture arising from crack formation and
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 167

growth applies for polymers as for metals but with greater emphasis on
the development of a plastic zone around the tip of the growing crack.
Most of the work on this subject has been carried out on test pieces
which have been subjected to cleavage stress. A crack at the face edge
is opened by a tensile force and it is customary in published work to
refer to this as the opening mode or, more briefly, with a capital,
Roman I. The propagation of a crack by shearing parallel to the plane
of the crack is mode 11.
Mode III, like mode II propagates the crack by shearing but in
torsion around an axis instead of across a plane.
The energy required to extend the crack to produce two units of new
surface is known as the fracture energy, Ge, the subscript indicating
that it is the critical energy, i.e., the minimum required to continue the
extension of the crack. If the crack is being propagated in the opening
mode the fracture energy is written G1e •
Consideration of the stress intensity at the tip of the crack instead of
the energy released as the crack grows, provides an alternative though
related approach to the strength of materials and, through the adhe-
sive, to the strength of adhesive joints. The stress intensity factor K is
employed when defining the stress fie1d around the crack. Fracture is
propagated when K exceeds some critical intensity K e which is known
as the fracture toughness and refers to a given material. The relation
between these parameters, as it applies in adhesive joints, has been
discussed by Kinloch and Shaw (1981). They quote:
Ki = EG for plane-stress (30)
and
Ki = EGj(1- v 2 ) for plane-strain (31)
where G is the fracture energy, E is Young's modulus and v IS
Poisson's ratio.
When other modes are involved, as they most probably are in a thin
adhesive layer in a joint, then:
(32)

The stress intensity factor within the adhesive can be approximated


from the fracture energy of the joint by means of these equations.
Approximation is, however, involved because, unlike the case of
pure1y homogeneous materials, the crack in an adhesive joint is
propagating near the discontinuities of two interfaces. As will have
168 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

been realized from the discussions in Chapters 2 and 3, joint failure,


whatever the nominal distribution of stress, may involve both K1 and
Ku, the values for which, in the conditions obtaining in a joint, are not
strictly associated with the tensile and in-plane shear modes of failure.
Moreover, the stress field other than very dose to the crack tip, may be
very different from that expected in homogeneous material. Although
the stress intensity factor is more familiar to engineers than fracture
energy, the latter is preferred by many using fracture mechanics in the
context of adhesive joints.
For an ideally brittle material the fracture energy may be equated to
the surface free energy of the newly produced surfaces but for real
materials, work is also done in deforming material in a zone around
the tip of the growing crack. Tbis work of deformation usually exceeds
the surface energy so greatly that the surface term can be negligible
compared with it. Moreover, the formation of the new surfaces in-
volves breaking as weH as slipping of the long chain molecules and the
energy required for this is greater than the surface free energy of the
stable surface normally in contact with air.
It might be thought that G Ic would be characteristic for a given
material and so to a certain extent it is, particularly if the rate of crack
growth is extrapolated back to zero. If measurements are made with a
test piece as in Fig. 104(a), the load at which the crack propagates is a
function of the crack length and the compliance of the system but the
variation of these two quantities lead to the fracture energy being
independent of the crack length. If the test piece is formed into the
form shown in Fig. 104(b) (Mostovoy and Ripling, 1969; Patrick et al.,
1969) the rate of change of the compliance of the system with crack
length is independent of the distance the crack has traveHed into the
test piece. Tbe fracture energy is then only dependent on the rate at
which it is propagated. For G Ic to be strictly a characteristic of the
material, the figure obtained by experiment should be the same limit-
ing figure whether it is obtained by calculation from the force required
to commence crack growth or as the limit below which a moving crack
ceases to move but this is not always found and two values, somewhat
apart, seem characteristic of some materials in place of a single value.
Rather more important than the question of the actual values to be
ascribed to GIc for various materials in various circumstances is the
observed connection of G1c with long term behaviour indicated by
other tests and the effect of moisture or immersion in water on GIc-
Kinloch et al. (1976) have shown a simple linear relation between GIc
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 169

\f--+---11- ( a l -----,~
C~ntr~ crack
for a short
distanc~

(b)

FIG. 104. Test pieces for the determination of fracture energy.

and log (time to failure) for aluminium lap-shear joints with epoxy
res ins held under continuous load and, moreover, have also shown that
if the joint is immersed in water the linear relation is simply displaced
to a line parallel to the original one at about 50 J.m- 2 lower. In the
latter case crack propagation proceeded along the interface rather than
in the bulk of the glue-line but this did not change the relation. The
drop is about 30% at the higher energies and proportionately higher as
the energy decreased.
Although properly made joints fail by cohesive fracture in the
adhesive, some adhesive on some substrates, show this sensitivity to
water and the sensitivity is obviously at the interface for, if generaliza-
tions can be made from so few data, the adhesive resin itself does not
fall in strength on immersion in water by an amount sufficient to
account for the loss in strength of the joint.
Published data on fracture energies have been limited to research
studies aimed at explaining given phenomena rather than at providing
comparative data and, as already mentioned, much of it is concerned
with the efIect of the presence of moisture. However, a selection of this
data is given in Table 12.
170 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 12
FRACTURE ENERGIES FOR EPOXIDE RESINS BASED ON BIS-PHENOL A (WITH V ARI-
OUS CURING AGENTS AND RUBBER TOUGHENING)

Curing agent and toughening agent if present Fracture Reference


energy
(J.m- 2 )

Amine eure;
Diethylene triamine 130 Phillips et al. (1978)
Hexamethylene diamine 575
Tetraethylene pentamine 70 Mostovoy and Ripling (1966)
Anhydride eure;
Hexahydrophthalic anhydride 116, 136 Baseom et al. (1976)
Methyl nadic anhydride 154 Diggwa (1974)
Toughened with CTBN,* cured with piperidine.
Proportion of CTBN 0% 121,t 116
4·5% 1050 -
10% 2725,3500 Bascom et al. (1976)
20% 3590 -
30% - 2200

* CTBN is earboxyl terminated butadiene-acrylonitrile rubber.


t First column refers to the bulk resin and the second to performance in a constant compliance
cantilever joint.

Traditionally, susceptibility to brittle fracture has been assessed by


some form of impact testing in which the test piece is notched and the
fracture energy expressed per unit of notch length rather than in terms
of area exposed by the fracture. However, the impact strengths so
recorded may have little relevance to joint strength because of the
relatively thin but extensive area of the adhesive layer and the very
considerable modification of this property above all others by the use
of blends or mixtures of polymers. For example, a phenol-
formaldehyde resin is in bulk form a brittle material. An Izod impact
strength for an unfilled cast resin is quoted (Roff and Scott, 1971) as
between 13 and 27 J.m- i . Although this can be modified by fillers in
moulded resins, the best that can be achieved is about 50 J.m- I .
Although no impact figures are available for polybutyral formal-P/F
mixtures in which the phenol-formaldehyde is the continuous phase,
typical metal-to-metal peel strengths quoted by Barth (1977) of 3'5-
4·5 kN.m- 1 suggest considerable resistance to crack growth. These
figures refer to an adhesive containing 62·5% of the phenolic compo-
nent and the remainder, polyvinyl butyral, the joint being tested under
a most severe condition for brittle failure, namely a low temperature of
-70°C. Nitrile-phenolic adhesives are similarly noted for their tough
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 171

behaviour although, as with the polyvinyl butyral-P/F, the nitrile


content is quite substantial being at least 30% of the total polymer
content.

Crazing
In some cases the fracture of a polymer is preceded by crazing whieh is
often accompanied by stress whitening. This phenomenon undoubtedly
occurs within some adhesive joints although the evidence, from the
nature of things, is not so visible as when stressing bulk plastics. The
appearance of broken joints can suggest that a craze or some form of
cavitation was present prior to rupture.
Crazing in polymerie materials has attracted attention principally in
connection with the presence of solvents or solvent vapours when the
stress required to produce crazing is very greatly reduced and leads to
premature failure of the material. The appearance of crazing in sheets
of clear polymeric material is characteristically of a thin zone of
material whieh, reflecting the transmitted light from internal surfaces
gives a bright, lace-like effect. If a fine particle filler has been incorpo-
rated, then failure occurs at the surfaces of the powder and stress
whitening becomes visible.
Crazing involves localized but not general yielding and, of course,
only occurs in polymers which show stress softening. The stress must
contain a negative hydrostatie component but most authorities state
that local stress-raising flaws must also be present. These determine
that a multitude of cracks arise instead of a single large crack but
equally they modify the stress field. This modification could be ex-
tremely localized and, in line with fracture mechanics theories for
polymers, the plastic zone at the tip of each growing crack, because of
the stress intensification, may enhance considerably the hydrostatie
component and cause the production of a void in the path of the
growing crack. The void will not, of course, be spherical but its
envelope will contain within its shape the parameters of the deviatoric
stress field. The combination of the overall shear stress with a cleavage
stress at the ends of a lap-shear joint provides the conditions necessary
for cavitation leading to general weakening of the polymer layer and
hence to failure.
Unlike a true crack in the material, a craze is capable of sustaining a
stress across it. The craze may grow and then break down to give a
crack which initiates failure but, as a result of the presence of the
craze, the breakdown to a crack may occur at a stress lower than would
172 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

cause yielding by shear. The result is brittle fracture in normally ductile


material. A detailed discussion of the mechanism is given by Kinloch
(1980).

Coef6cient of Thermal Expansion


Most of the structural adhesives need curing at elevated temperatures
whilst they are used at lower temperatures. As the coefficient of
thermal expansion of organic resins is greater than that of metals,
shrinkage stresses are sometimes added to the load transmitted across
the interface.
The figure usually quoted is the coefficient of linear expansion, the
increase of unit length per oe. For mild steel this is about 11 x 10-6 , fm
aluminium about 25 x 10-6 • Those for polymers below their glass
transition may be illustrated by figures of about 50 x 10-6 for a cast
epoxy, among the lowest of the polymerie materials. The thermal
expansion is reduced by the incorporation of mineral fillers such as
finely divided silica or alumina and, as many structural adhesives
contain some filling powders to assist flow or gap-filling properties,
actually adhesive resins may show somewhat lower figures. A Ciba-
Geigy wall chart quotes figures for individual adhesives of the Araldite
range from 40-100 x 10-6 . Other figures, which, because they refer to
unfilled polymers, should be regarded as maxima rather than typical
for adhesive formulations, are in Table 13.
Above the glass transition temperature the coefficient of thermal
expansion is much higher and that of rubbery materials is over 200 x
10-6 , the silicone rubbers having particularly high values of 250-400 x
10-6 .
The problem of shrinkage stresses is complicated by the higheI
coefficient above the glass transition being coupled with a lack 01

TABLE 13
COEFFICIENTS OF LINEAR THERMAL EXPANSION, a, OF
POLYMERS

Polymer
Polyvinyl acetate 85
Polyvinyl formal 80
Polyvinyl butyral 150
Cast, unfilled phenol-formaldehyde 80
Moulded, unfilled phenol-formaldehyde 45
THE GENERAL PROPERTIES OF POLYMERIC ADHESIVES 173

knowledge of the conditions at which adhesion starts to be effective.


Moreover, although at some still elevated temperature during cooling
after curing effective adhesion exists at the interface, the adhesive itself
may still exhibit stress relaxation, mitigating the onset of shrinkage
stress until a still lower temperature. Possibly a greater importance
attaches to the coefficient of linear expansion when an adhesive is
likely to sustain intermittently very low temperatures although nor-
mally maintained at those of the temperate atmosphere.

Resistance to Deterioration
The corrosion of metals in adverse atmospheres has a corresponding
phenomenon in polymeric materials whereby they deteriorate in the
presence of the oxygen and/or moisture of the atmosphere. In some
cases heat must also be present to assist these agencies. The oxidation
of many polymers is catalysed by light but this is not usually a problem
with adhesives because the joint is rarely exposed to light. However,
the point should be borne in mind and designs should be avoided which
expose free edges and, in particular, fillets of adhesive to sunlight. The
chemical changes which accompany slow atmospheric oxidation or the
more rapid changes which occur when used in air at elevated tempera-
tures vary with the nature of the polymer. The phenol-formaldehyde
resins slowly change their chemical constitution with quinone forma-
tion and increased brittleness but quite high temperatures are required
for this. Resorcinol-formaldehyde resins, used principally in the pres-
ent context as wood adhesives, are more stable than the wood itself to
atmospheric deterioration. Polyvinyl acetate and polychloroprene both
lose acid as a consequence of oxidation. Polyurethanes and epoxies are
more likely to deteriorate by hydrolysis than by oxidation and the
former polymers are CJluite unsuitable for use in hot, wet climates.
The effect of hydrolysis is a loss of molecular weight and hence of
strength and modulus but moist conditions, particularly in the presence
of stress, are likely to cause deterioration of the joint strength quite
separately to the loss in the bulk properties of the adhesive polymer.
Relevant to this is a warning against the use of adhesives in situations
where live steam is present or is likely.
Adhesives are usually formulated with additives to protect or to
counteract the effects of the degrading agent to which the polymer is
most susceptible. In the case of materials normally used at atmospheric
temperatures, if susceptible to oxidation, an antioxidant is added. The
degradation of polyurethanes by hydrolysis can be mitigated by the use
174 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

of carbodi-imides and by chelating metal ions which catalyse the


oxidation (Hole and Abbott, 1973). High temperature adhesives are
more difficult to stabilize though the polymers themselves are intrinsi-
cally more stable. The problem is the volatility of the USUal antioxid-
ants whieh quench free radieal chain reactions. With polyimides and
polybenzimidazoles, operating at temperatures above 250°C and up to
500°C, arsenious oxide is used, its action being said to be to remove
ferrous ions which diffuse from the surface of stainless steel into the
adhesive and which would otherwise act as oxidation catalysts
(Black and Blomquist, 1962). Its presence is not so necessary when
bonding titanium but it is an unpleasant additive and has militated
against the development of high temperature bonding of stainless steel.
However, the chemistry of the means of protecting these special
duty adhesives is not the concern of the engineer except to the extent
that the presence of hazardous materials must be allowed for in
considering the applicability of adhesive bonding in the given cir-
cumstances. More properly is he concerned with how joints incorporat-
ing such materials behave. Indeed, the whole question of the effect of
adverse or severe environments on all polymerie adhesives is best
dealt with in the context of the performance of joints rather than that
of the polymer. This approach is followed in Chapter 7.
Chapter 5

Factors In8uencing the Choice 01 Adhesive

The choice of an adhesive for a given situation is by no me ans the


proverbial Hobson's choice. There is usually, though not invariably, a
real choice to be made as adhesives are versatile and the overlaps
between their usages frequent. Multiple choice is common for all but
the most demanding of applications. It should not be possible to write
a computer pro gram into which all the relevant particulars were fed
and from which only a single answer would be extracted. If the
program were written to produce unique answers to all applications of
adhesives, then the programmer would have built into the pro gram at a
number of stages, subjective judgments masquerading as objective
facts.
This book is concerned with engineering applications and therefore
with adhesives employed to transmit stresses of appreciable mag-
nitude; pressure sensitive adhesives and a wide range of thermoplastic
adhesives are therefore not considered. Some thermoplastic materials
are, however, used in engineering assembly operations where signific-
ant temperature rise is unlikely and some in applications important to
engineering but where stress is virtually absent or very low. These
materials are called non-structural engineering adhesives in Table 14,
which lists the adhesives which are considered in this chapter as
available for choice.
Some adhesives such as the epoxies are very tolerant of differences
of substrates. Other adhesives with equally good stress-bearing proper-
ties are only usable if a primer is first applied to the substrate surface.
It could therefore be stated that such adhesives are unsuitable for such
substrates and although this is an obviously true statement it adversely
prejudges the issue since the appropriate combination of primer and
175
176 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 14
LIST OF ADHESIVE TYPES

Structural adhesives
Polyvinyl formal-phenol formaldehyde
Polyvinyl butyral-phenol formaldehyde
Nitrile-phenolic
Epoxy-polyamine
Epoxy-anhydride
Epoxy-nylon
Epoxy-phenolic
Epoxy-polyamide
Epoxy-polyurethane
Toughened acrylic (sometimes referred to as 2nd generation acrylic)
Polyimide
Polybenzimidazole
Polyquinoxaline
Polyester plus isocyanate
Phenol formaldehyde (abbreviated elsewhere in this book to P/F)
Resorcinol formaldehyde (abbreviated to R/F)
Urea formaldehyde (abbreviated to U/F)
Melamine formaldehyde (abbreviated to M/F)

Semi-structural adhesives
Polyurethane
Polyvinyl acetate
Hot melt adhesives such as:
Polyester, polyamide, polycarbonate, nylons (which are a particular type
of polyamide).

Non-structural adhesives
Polychloroprene
Di-acrylic esters
Cyanoacrylate
Acrylic esters as monomers.

adhesive will give a perfect1y satisfactory joint. The compatibility


required is that of substrate and adhesive system, the word 'system'
subsuming not only the use, where necessary of a primer, but also
whether application is by liquid, solution or the melting of asolid.
Most structural adhesives will be used with metals, timber, concrete
or fibre-reinforced plastics, and it is to these substrates that attention is
here confined. An important application in the use of adhesives is the
joining of two different substrates where other methods of fastening
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 177

are applicable to one but not both substrates. Adhesive systems which
cope with mixed substrates are important and the problem of making
such joints may be solved by the use of a primer on one substrate to
enable an adhesive to be used which is fully compatible with the other,
or there may be available adhesives compatible with both.
So much can be achieved in altering the mechanical, thermal or
surface properties of polymers by the addition of relatively minor
quantities of other polymeric or non-polymeric materials that fre-
quently the problem of choice becomes that of choosing from a range
of materials based essentiallyon the same adhesive but formulated
with different additives or cross-linked by a different temperature and
time sequence. However, with all these caveats stated it is still neces-
sary to start the process of selection from a consideration of adhesive-
substrate compatibility and to proceed from that decision to others
involving the use for which the structure is intended, the method
whereby it is to be made and the environment in which it must
function for a greater or lesser period of time.

INTERACTION WITH SUBSTRATE

Theories of adhesion postulate that there is no unique interaction


between an adhesive and a substrate. Depending on the nature of the
substrate, the attraction between them can be due to the van der
Waals's forces which, electronic in nature, are ubiquitous and are
regarded as a form of physical adhesion. By contrast, chemical interac-
tion can exist when functionally active groups of atoms of the adhesive
form much stronger chemical bonds with the substrate. Of course,
these bonds are similady electronic in nature but they depend on the
sharing or transfer of electrons instead of upon the electro-magnetic
field created by these electrons. There are but two types of adhesive
bond, one physical and one chemical. When a long chain polymer is
above its glass transition temperature (Tg), its molecules are mobile in
the sense that they are capable of self diffusion so that when one plane
surface of them is brought into dose contact with a second surface, the
molecules diffuse both ways across the interface between them thus
obliterating it. Such diffusion would be expected to occur when both
sides of the interface are composed of the same molecules. Difficulties
do exist when different sorts of molecules are involved and, once again,
the term compatibility is invoked. Compatible polymers, both above
178 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

their respective Tg do diffuse, albeit to a limited extent, into each other


whereas incompatible polymers do not. This is the essence of the
diffusion theory of adhesion and is important in explaining the adhe-
sion of rubbery adhesives to themselves and to each other. To the
extent to whieh high temperatures are involved in making the adhesive
bond with, for example, a fibre reinforced plastie material, so some
degree of diffusion of polymerie adhesive into the polymerie substrate
might become a factor in making the bond. A common situation in
which a bond formed by diffusion is important is that whieh occurs in
using contact adhesives; i.e., adhesives whieh form an immediately
adequate bond simply by contacting the surfaces umder pressure. The
adhesive, usually dissolved in a volatile organie solvent, is coated on
both the surfaces to be joined; the solvent is allowed to evaporate thus
establishing the bonding of films of adhesive to both substrates sepa-
rately. The two substrates are then brought together and a diffusion
bond between the two layers of adhesive completes the formation of
the joint.
There are two other explanations whieh are relevant to some situa-
tions though neither is important in connection with structural adhe-
sives. The first is the electrostatie theory whieh, as its title indieates,
involves adhesive and substrate as the plates of an electric condenser,
work being necessary to separate the plates. The second is the mechan-
ical theory in whieh mechanieal interlocking of adhesive and substrate
involves fracture of parts of one or other before separation can be
achieved. The former theory has some part to play in explanations of
the adhesion of pressure sensitive tapes to substrates and the adhesion
of sputtered metal in the surface decoration of plastics. The latter is
important in the adhesion of polymers to textiles and papers and, to a
much lesser extent, to wood.
The word 'ubiquitous' has been deliberately used in connection with
van der Waals's forces. If mechanieal or electrostatic adhesion contri-
butes to the strength of an adhesive bond it cannot be the sole
contributor, for van der Waals's forces are always present and even
contribute to the efficiency of any mechanieal component.
Although simple theoretical treatment of the adhesion of one solid
phase in intimate contact with another solid rigid phase suggests that
physieal van der Waals's forces are alone more than sufficient to
account for bond strength, most technologists believe that chemical
interaction across the interface is present to a greater or lesser extent
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 179

in all cases where a good bond strength with the adhesive, resistant to
displacement by water, is obtained. It must be admitted that, as with
many beliefs, conviction is based on circumstantial rather than compel-
ling evidence. Nevertheless, in choosing an adhesive, the presence of
chemically reactive groups which can conceivably react with receptor
groups on the substrate surface provides, in general, asound guide.
Adhesives lacking such groups may require a more reactive priming
layer before consistent, adequate adhesion can be guaranteed. Trans-
lating these generalities into specific instances is less easy and less
certain. All except the noble met als are covered with an oxide coating
and most of these oxides are hydrated to a greater or lesser extent.
Surface preparations are aimed at the preparation of a thin, reproduci-
ble coherent oxide or basic oxide layer. These are discussed in Chapter
6 and the foIlowing discussions and indications assurne that a clean
prepared surface is being used. Hydrated oxide layers aIl seem to be
capable of forming good bonds with epoxide resins. The chemical bond
responsible for this adhesive bond may be due to hydrogen bonding, of
a type intermediate between a strongly polar van der Waals's interac-
tion and a truly chemical linkage. Such interactions are frequently
invoked because good adhesive bonds are often easily formed with the
types of comp')und which are known, in other circumstances, to form
hydrogen bonds. Epoxy resins, as cross-linking occurs, transform their
terminal epoxy groups into hydroxyl groups pendant to the larger
molecular structure which cross-linking builds. Hydroxyl groups form
hydrogen bonded structures very easily if other suitably receptive
atomic groups are available. However, some hydrogen bonds between
adhesive and substrate are susceptible to attack by water as weIl as
other hydroxyl-containing reagents and hence the relative strength of
the different hydrogen bonds will determine the resistance of the joint
to attack by water. Epoxides can be used as adhesives for wood but the
ability of wood to absorb and store quantities of moisture, a proportion
of which will naturally migrate to the interface with an epoxy adhe-
sive, leads to disruption of any hydrogen bonding at the interface. By
contrast, the methylol groups of phenol-formaldehyde and resorcinol-
formaldehyde resins have been shown (Allan and Neogi, 1971) to react
chemicaIly with guaiacyl groups of the lignin molecules present in the
wood. It is not necessary for the engineer to be concerned in detail
with these chemical niceties but they are mentioned to show that there
is a rational basis for the adhesive preferences stated in Table 15.
180 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 15
ADHESIVE CHOICE INDICATED BY SUBSTRATE TO BE JOINED

<Il
::: ~
.~
~ e: <Il
:::
C::: r....
r.... r....
e
Q,
"c: ~ e: <Il
..:::Q, -~
..s
<Il
::: I.)
e:0 i: ~ ~
..:::I.) 1!tJlJ-
.~ ~ ....
0
.... ~ I .5~
;::l e:
::: ;:! 0 ~
..cE
.~
;>,
;:! c- e: ;>,
'E
r.... r....
z
0
~ ~~ G ~ ::5 ~ ~ ~~
Aluminium x x x x xx xx xx
Brass x x x x xx xx xx
Bronze x x x xx xx xx
Cadmium x x x xx
Chromium x x x xx
Copper x x x x xx xx xx
Ferrous
alloys x x x x xx xx xx
Nickel x x x x xx xx xx
Titanium x x x x xx xx xx
Tungsten x xx
Zinc x x x x x xx xx xx

Polyester xx xx
Epoxide x xx xx

Wood x x x xx xx x x x
Concrete x x xx xx x
Glass x x x x x x x

x Can be used
x x Preferred in structural applications

STRUCTURAL ADHESIVES FOR METALS

Metals possess high energy surfaces and hence almost any polymer will
adhere to them. The variations between metals noted in Table 15 arise
partly from ignorance, e.g., bonding tungsten with polychloroprene is
not recorded and may result in corrosion though this is unlikely and its
general behaviour towards adhesives would be expected to replicate
that of chromium. Other differences between met als in the table arise
from the chemistry of the processes whereby liquid adhesives become
cross-linked rigid solids. Phenols or resorcinol-formaldehyde resins or
amino-formaldehyde resins (i.e., recorded as P/F, R/F and V/F resins
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 181

in Table 15) do not make good adhesives for metals although they are
used for wire coating as insulators. The difficulty lies in the need to
apply the resin dissolved in solvent, to remove the solvent without
cross-linking the resin and then to combine the two (metallic surfaces).
Following this stage, cross-linking with the elimination of water must
occur but the water is unable to escape and causes porosity. These
difficulties do not occur in wire coating nor when a PjF resin is used as
a primer. A number of very efl'ective mixed adhesives incorporate PjF
resins but, although these still eliminate water on cross-linking, the
amount eliminated is necessarily smaller and it is dissolved in the
second component of the adhesive because this is always rubbery in
nature in that it has a much lower glass transition temperature than the
P jF resin. Nevertheless, such mixed adhesives must always be cured
under pressure to prevent porosity and hence weak joints.
With these preliminary explanations, the way is cleared for a discus-
sion of the selection of structural adhesives for aluminium and the
ferrous metals, leaving other metals to be treated by way of deviations
from the main steam structural metals.
The adhesives from which choice is to be made are the epoxies in all
their various modifications and combination with other polymers, the
polyvinyl acetals or the nitrile rubbers, both the latter in conjunction
with phenol formaldehyde resins. Shields (1976a) lists 56 compositions
specifically designated as structural adhesives for metal-to-metal bond-
ing from six primary manufacturers operating in the UK or importing
from the USA. There are many more materials on the market for
which the resins have been bought in from one of these primary
manufacturers, formulated before selling under a brand name. Epoxy
resins are easily available, if expensive, and with the multiplicity of
curing agents the possibilities of useful new combinations are numer-
ous. Moreover, unlike other adhesive compositions, it is not uncom-
mon for the user to purchase an epoxide resin from one source and to
mix it with a curing agent or hardener purchased from another.
Mixtures of other polymers with epoxies and the phenol-formaldehyde
mixtures with polyvinyl acetals or nitriles are the province of the
adhesive supplier and are not usually made by the actual user of the
material.
Historically, the first structural adhesive successfully used was used
in the aircraft industry for bonding aluminium to wood (Garnish,
1977). It was a polyvinyl formal composition with a phenol formal-
dehyde resin.
TABLE 16
STRUGrURAL METAL-TO-METAL ADHESIVES (FOR USE AT NORMAL TEMPERATURES)

Adhesive type Curing conditions Primer Lap-shear strength Peel strength Product Notes
needed? at room to which
High pressure Temperature RT(MN.m- 2 ) c.80°C temperature data refers
needed (OC) (kN.m- 1 )

Polyvinyl Yes c. 150 No 44 14-28 5 Redux 775 Performance at


formal-P/F (Ciba-Geigy) higher temperatures
Polyvinyl depends on amount
butyral-P /F Yes c. 150 No 12·9 of P/F but the
Nitrile phenolic Yes ;;.150 Yes 40 c. 12 9·5 larger the P /F ratio
c.1MN.m- 2 the lower the RT
strength and the
lower the peel
strength
Epoxy-polyamine No RT and up No 15-40 0·7
to 140
Epoxy-anhydride No 2h at No 27 33 Poor
150
Epoxy-nylon 0·3 170 No 58 20 FM 1000
MN.m-2 (Cyanamid)
Epoxy-nitrile No No 25 3·9 See Garnish (1977),
p. 88; also Bolger
(1973), p. 47
Epoxy-phenolic Yes No Torsional Torsional 1·6 Hidux-1196C
46 39 (Ciba-Geigy
30
Epoxy-novolac No Yes 28 2·6 DowX2638·3 See Dannenberg and
May (1969), p. 48
Epoxy-poly- No No 31 10 3·2 See Panek (1977),
sulphide p. 379
Epoxy-poly- No No No information
urethane available
Toughened epoxy
1 part No 120 No >33 3·4 HysoI
2 part No RT No 33 6·8 (Dexter)
Modified epoxy No 170 No 55 9·0 BSL 308
(Bonded
Structures
Ltd)
184 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

The differences and similarities of the three main types of structural


adhesive are shown in Table 16 with some indication of the variations
obtainable with the epoxies by blending with other polymers. It will be
appreciated from Table 16 that the epoxy-nylon adhesives can show an
outstanding combination of high lap-shear strength and peel resistance
but this is at the expense of increased susceptibility to moisture and, as
is discussed in Chapter 7, trials in hot, wet climates show poor
durability, though loss of strength by outdoor exposure in England is
not particularly serious (Cotter, 1977). Unfortunately, it is not possible
to show in a table the effects of the various curing agents on epoxy
resins nor to discuss the effects of various changes which lead manufac-
turers to describe their products as 'modified' epoxies because the
modifications are never detailed. The following statements are relevant
to the various curing agents which can be used.
Room temperature curing epoxies are usually based on the
polyamine-amide compounds marketed as 'Versamids'. Although com-
positions containing them will give a satisfactory cure at room temper-
ature, there is always advantage to be obtained in giving a post-cure of
a few hours at, say, 70°C. Amine-cured epoxies, including those cured
with the Vers amids, are very tolerant of variations in the ratio of curing
agent to resin and many formulators build in an additional factor of
safety in this respect using as the curing agent a polyamine which has
already been reacted with an epoxy of low molecular weight but with
excess polyamine, thus developing a two-part formulation in which
equal volumes of some other simple ratio such as 2: 1, need be used.
This eliminates the large relative error associated with weighing small
quantities to mix with large quantities.
Adhesives are not used at stresses near the ultimate strength as
measured in the lap-shear joint, though this is considerably less than
that shown in torsion al shear where the stress is completely free from a
cleavage component. Working stresses are, at most, about 20% of the
lap-shear ultimate tensile strength and for this reason there will be but
little differentiation between the structural adhesives of Table 16.
Differentiation in terms of strength will depend upon the need for
resistance to peeling or cleavage and resistance to shock loading. There
are, however, considerable differences in these properties within the
categories listed in Table 16. Thus Dannenberg and May (1969) give
qualitative assessments of differences obtained with different
polyamines and tertiary amine curing agents as weIl as various anhyd-
rides. Moreover, since the manufacturer may use mixed curatives and
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 185

does not necessarily disclose the composition, it is necessary to use a


check-list of requirements against the manufacturer's stated properties
and then to compare candidates for use against price. Epoxy adhesives
are expensive when compared with polyvinylformal-phenol formal-
dehyde mixtures and even against nitrile phenolics. The higher lap-
shear strength that they, in general, give may not justify the extra cost.
The items which might reasonably be included on a check list
comprise the following:

Check-list for Structural Metal Adhesives Used at Temperatures up


to 70°C

Strength Requirements

Shear strength. Is this a limiting feature of the design so that


adhesives of the highest shear strength must be used? If so, choose a
suitably modified epoxy and consider whether peel strength is also
required.

Cleavage and tension. If these stresses are high then look for an
adhesive with high peeling strength. Be prepared to sacrifice some
shear strength to obtain it.

Impact strength. Look for high peel strength if impact resistance is


not separately given in the manufacturer's data sheets. Two-
component adhesives that separate into two phases will, in general,
give the highest impact resistance. These are frequently described as
'toughened'. Some relatively flexible materials show low lap-shear
strength and only moderate peel but are, nevertheless exceedingly
impact resistant, e.g., epoxy-polysulphide mixtures.

Deformation
For most designs, deformation caused by shear within the adhesive can
be ignored, the deformation of the joint being solely that of the
adherends. However, for some precisely dimensioned designs, the
modulus may need to be known and the possibility considered that the
shear deformation of the adhesive may progressively increase, i.e., it
may creep.
186 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Manufacturing Conditions

Primer. This will involve a separate stage in manufacture. Some


adhesives will only function on correctly primed surfaces. This is an
issue quite separate from the question of whether a siloxane primer is
desirable for special environmental conditions (see Chapter 6).

Bonding pressure. Choice is restricted if pressure cannot be applied


to the joint.

Bonding temperature. Room temperature curing adhesives never


give as high a lap-shear strength as is obtained if curing is at elevated
temperature. The best results demand the correct temperature
specified for the adhesive. If the temperature is determined by other
factors, then the adhesive must be chosen to fit the temperature.

Pot life. Must be checked in relation to process, facilities availabIe


and competence of statt

Form oi adhesive. Solid, liquid or paste, if solid in film form is it


supported on a carrier? Pastes sometimes contain balotini to ensure
correct spacing of the adherends and thus void a 'starved' joint.
Primers are always low viscosity solutions from which solvent must be
allowed to evaporate; they may need aseparate cure or partial cure.
The bare items of the above list can now be fiJIed out with more detail
without the danger of obscuring an overall view of those items about
which decisions must be made. This now follows using substantially the
same headings as were taken in the list.

Strength Requirements

Peel resistance. The straight epoxies cured by either poly amines


(amides) or anhydride are poor in peel. Polyvinyl formal-phenol for-
maldehyde (P/F) and butyral-P/F are moderate but the peel varies with
the P/F content. As the P/F content increases, room temperature
lap-shear and peel strength decrease but these properties improve at
elevated temperatures. The epoxy-polyamides (epoxy-nylon) such as
FM 1000 (Cyanamid Corp.) give extremely good peel resistance
under dry conditions. Hockney (1970) quotes 511 N for an adhesive of
this type with a strip 25·4 mm wide, peeled at 90°. The peeling
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 187

member was of 30 gauge (0·340 mm) aluminium alloy to BS 2L73. lt


was bonded to a backing member of 12 gauge (3·556mm) clad
aluminium alloy also to BS 2L73. The same author (1973) quotes
241 N for a typical nitrile-phenolic under the same conditions of test, a
figure which would also be regarded as good. Moderate peel strength is
shown by epoxy-nitrile, epoxy-sulphide and 1-part toughened epoxies.
Using ASTM D1876 T-peel test, Lees (1980,1982) quotes a figure of
630 N for a width of 1 cm for a heat-cured 1 part toughened epoxy.

Impact strength. This tends to be high when peel strength is high.


Epoxy-polysulphides have been regarded as giving the highest impact
strengths but modern toughened epoxies are at least equally good and
show higher shear strengths.

Deformation Characteristics

Modulus. Unmodified epoxy resins have the highest modulus values


together with epoxy-phenolics and epoxy-novolacs.

Creep. The gradual extension of a lap-shear joint can be observed


under conditions in which the adhesive is near or above its Tg , the
stress is relatively high and its shear component much greater than the
cleavage (hydrostatic) component. What little data is available refers to
polyvinyl-P/F and to a modified epoxy of undisclosed composition. If
the absence of creep is important in design then an adhesive with Tg
weH above the important working temperature must be chosen. In the
absence of other data, the heat distortion temperature can be used as a
criterion.

Manufacturing Conditions
The conditions to be considered are those relating to the form in which
the adhesive is supplied and the procedure it is necessary to adopt to
obtain the bond, excluding the question of surface preparation which is
dealt with in Chapter 6. Also dealt with in Chapter 6 is the general
question of the need for and use of primers; but with certain adhesive
systems, the primer is an integral part of the adhesive which cannot
function in its absence; these must be briefly mentioned now.

The use of primers. Primers are an essential component of some


nitrile-phenolic and epoxy-novolac adhesives. They consist of solutions
188 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

of phenol-formaldehyde resins (at an 'A' stage in the condensation


process) in an organic solvent, usually industrial methylated spirit
(IMS). The solution is applied by brushing to the metal surface
followed by a light stoving. Other nitrile-phenolic adhesives are formu-
lated with a greater proportion of resin than of rubber. These do not
need primed surfaces but will show lower peel and impact resistance.

Bonding pressure. The phenolic-containing adhesives require high


pressure to be maintained over the bonding area throughout the curing
process. The requirement does not apply to the epoxy-novolac adhe-
sives because the cross-linking process in these materials only involves
the epoxide group. Epoxide adhesives do not need pressure, although
slight press ure is always advantageous, particularly if the temperature
is to be raised. It then serves to ensure maintenance of true surface
contact which could be disturbed by thermal expansion or alterations
in the viscosity of the adhesive or by the shrinkage, say about 3%,
which accompanies the chemical curing. Some types of locating jig
involving slight press ure is all that is needed for this.

Bonding temperature. The phenolic-containing adhesives all require


temperatures of 150°C or over for curing. Similarly high temperatures
are required for all single part and far anhydride-cured epoxies. The
curing temperatures required for two-part polyamine-cured epoxies
vary from room temperature to 150°C depending on the curing agent
and the degree of cure required. The bond strength of an epoxy resin
cured only at room temperature can never be as high as that obtaina-
ble by curing at elevated temperature. Nevertheless, it may be ade-
quate since safe design never requires stresses near the ultimate
performance of the material. It is sometimes convenient to start the
curing of a polyamine-epoxy by an exposure of 1-2 h at an elevated
temperature, say 60-80°C and sometimes it is convenient to finish a
room temperature cure by exposure to higher temperatures. Amine-
curing agents vary considerably in their rate of cure and this can
indeed be influenced by the presence of catalysts. Moreover, as the
morphology of the cured material is complex and the chemical reac-
tions leading to it varied, so the structure finally obtained depends on
the thermal path by which it was reached. Properties can therefore be
infiuenced by the curing cyde and the speeding-up of a low tempera-
ture cure by drastically raising the temperature may give less desirable
properties than the use of a high temperature curing agent even though
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 189

the latter may require a longer time for eure. Some figures showing the
effeet of temperature with a typical epoxy resin eured with Vers amid
115 are given by Rayner (1965). The figures are for lap-shear tests:
Room temperature 2~ days 12MN.m-z
Room temperature 15 days 16 MN.m- z
95°C eure 30 min 39 MN.m- z
145°C eure 30 min 41 MN.m- z
The inerease in prolonging the room temperature eure after 2~ days is
small and neither is there any advantage in using temperatures above
95°C.
Table 17 makes similar points with respeet to two amine-type euring
agents, one a poly amine and the other eontaining one primary and one
tertiary amino group. These groups function during eure in very
different ways and lead to different struetures. The seeond eompound
does not give a satisfaetory eure at room temperature and is obviously
better employed at 95°C or higher. The gain in strength over triethyl-
enetetramine (TETA) if employed at the higher temperature would
not be worth its greater eost were it not that its performance at
elevated temperatures is superior.

TABLE 17
LOW TEMPERATURE VS IDGH TEMPERATURE eURE OF AMINE eURING AGENTS
(AFTERRAYNER, 1965 BUTUNITSTRANSFORMEDToMN.m- 2)

eure temperature

Room
temperature

Triethylenetetramine (TETA)
Time of eure 3 days 30min 30min
Bond strength 8·0 21·9 23·6
Time of eure 15 days
Bond strength 11·6

N.N.diethylaminopropylamine
Time of eure 16h
Bond strength 4·8 22·3 27·8

Time of eure after 15 days


Bond strength 5·8
190 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

'Post-curing' is the usual expression for aperiod of heating which


continues the curing reactions at elevated temperature after the resin is
sufficiently cross-linked and rigid to withstand removal from the hyd-
raulic press or positioning jigs. It is conveniently done in an oven if
above lOOoe or in a warm room if below 60°C. A post-eure after the
removal of jigs can usefully serve a stress-relieving function.

Pot life. All two-part adhesives start reacting when the parts are
mixed and therefore have a limited pot life. The greater the speed of
curing, the shorter will be the pot life. One-part adhesives have a
longer, but not necessarily an infinitely long, pot life. It is frequently
necessary to store such materials in refrigerators if they are to be kept
for weeks or months before use and, during use, unused material is
best returned to cold storage. One-part, anhydride cured epoxies
possess a good pot life but dicyandiamide (dicy), which is asolid at
room temperature and remains suspended without reaction in liquid
epoxy resins, may be regarded as completely stable below 40°C.

Form of adhesive. The form in which the adhesive is presented to


the surfaces to be loaded can usually be suited to the convenience of
the process to be used in fabrication. The first major division is
between adhesives which are supplied as such and those which are
solids or tacky polymers coated on to a thin carrier which is incorpo-
rated into the joint. Adhesives carried on a supporting flexible material
are, naturally, one part materials, aIthough they may be used in
conjunction with a surface that is separately primed.

The Advantage 01 Supported Fihnic Adhesives ('Tapes')


Supported adhesives are being increasingly used in the aircraft industry
and, in the US especially, are usually referred to as tapes. The word is
a misnomer. Tape adhesives are not necessarily or even usually in
narrow strips of material and are available in rolls up to 1 m wide. The
advantages may be itemized briefly as:
(a) Uniform thickness of glue-line with prevention of loss from
joint edge at high temperature under pressure, i.e. avoidance of
starved joints;
(b) Ease of placement into joint when shaped joints are involved;
(c) Ease of completely clean working;
(d) Freedom from solvent fumes;
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 191

(e) Possibly more uniform joint strength;


(f) Absence of need for apparatus either for mlX1ng two part
adhesives or spreading the adhesive when mixed.
Supported films of nitrile phenolic adhesives are formulated with
excess phenolic resin which gives a more brittle polymer with resin on
its surface. Tbis surface excess obviates the need for priming the metal
substrate with nitrile-phenolic before bonding. Tbe brittleness of the
adhesive is offset by the presence of fabric as the carrier. Solvent based
nitrile phenolics contain an excess of nitrile rubber. Materials which
are used as the support comprise: cotton; nylon; polyester; glass; tissue
paper. Tbe support may be either woven or non-woven.

Unsupported Films
Unsupported films are also available for polyvinyl formal-P/F and
some modified epoxy alone or mixed with other polymers.

Liquids and Pastes


Tbe original polyvinyl formal-P/F structural adhesive was formulated
as an alcoholic solution of a resole-stage P/F resin which was applied
to the metal surface and on to which a suitable proportion of the
powdered polyvinyl formal was sprinkled as if from a pepper pot.
After allowing the evaporation of the solvent, the joint was closed and
subjected to heat and pressure. If this appears to be a crude procedure,
it must be admitted that it certainly is, but it produced bonds of the
expected strength provided that the metal surface had been properly
prepared. Moreover, no preliminary primer was necessary as the
method of application ensured complete wetting of the substrate by the
solution followed by drying of the phenol-formaldehyde resin in con-
tact with the metal. It also allowed the manufacturer to modify the
PVF-P/F ratio to meet the conflicting requirements of a material which
would show good shear strength at higher temperatures whilst main-
taining room temperature peel. Tbis procedure is still adopted with an
epoxy-phenolic applied in solution and PV formal or butyral shaken on
as powder.
Nitrile-phenolics are formulated as two-part adhesives dissolved in
ketonic, chlorinated or, most frequently, in blended solvents. Some
one-part materials are available which do not contain vulcanizing
ingredients for the nitrile rubber but rely on a novolac resin with the
addition of hexamine to introduce some cross-linking into the rubber.
192 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Where vulcanizing ingredients are present, the pot life of one of the
two parts can be somewhat limited.
Epoxy adhesives can be liquid without the addition of a solvent and
this great advantage is further extended by employing as reactive
viscosity modifiers, chemical compounds of low molecular weight but
containing epoxy groups. Before curing, these compounds act to re-
duce the viscosity, but after curing they are locked into the network
and hence cannot be lost by evaporation or leaching to cause shrink-
age. The bulk production of epoxy resins is based on glycidyl ethers
derived from bis-phenol A. The simplest member of this series is a
crystalline solid but, by building materials of higher molecular weight,
the crystallinity is disrupted and more-or-Iess viscous liquids are ob-
tai.IJed. Liquids of much lower viscosity results when the epoxide is
based on the glycidyl ethers derived from glycerol, polypropylene
glycol or aliphatic cyclic compounds (bis-phenol A is an aromatic-
based compound). Those derived from glycerol and polypropylene
glycol give, when cross-linked, flexible resins unsllited to structural
application. The aliphatic compounds are, however, rigid.
The disadvantage of liquids, particularly those of low viscosity, is
that with even slight pressure between the adherends the liquid can
drain from the joint leaving insufficient to form a continuous film. This
condition is often called a 'starved joint'. A starved joint is necessarily
a weak joint. All liquids decrease in viscosity when heated so that the
situation is aggravated by the rise in temperature necessary for curing
until the eure has progressed sufficiently to raise the viscosity and gel
the adhesive. This state of affairs can be avoided by the use of a paste
instead of a liquid. A paste can be defined as a concentrated dispersion
of a finely divided solid in a liquid. In a suspension, the particles of the
solid are sufficientIy diluted by the liquid not to be in contact with each
other. In a paste the particles, or rather partieles with an adsorbed film
of liquid, are packed together each always in contact with several
neighbours. A suspension necessarily has a greater viscosity that a pure
liquid but a paste not only possesses a much greater viscosity, it has in
addition other properties. It shows elasticity and rigidity and will flow
only when the shearing stress is greater than a certain value charac-
teristic of the system and concentration of the powder. By compound-
ing the resin and its curing agent with a suitable powder, it is possible
to ensure the maintenance of an adequate film even under the light
pressure and at elevated temperature of curing. In general, the pres-
ence of such filling powders decreases the lap-shear strength at room
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 193

temperature in proportion to the amount present but may considerably


enhance the strength at moderately elevated temperatures under
100°C.
Epoxy resins give an excellent bond with a properly prepared
aluminium surface. This has led to the use of aluminium powder as a
filler for epoxide resins, giving an example of a filler which enhances
tensile shear strength both at room and at elevated temperatures.

Inftuence of Metal of Adherend


Much of the data reported on adhesive strength has been obtained
from the lap-shear test piece. In the single-lap shear there is adefinite
bending moment giving rise to a cleavage stress; there is also a smaller
cleavage stress associated with the ends of the outer adherends in a
double-Iap shear test piece even though the bending moment is very
much smaller. In both of these types of test, therefore, the rigidity of
the adherends influences the stress pattern and hence the recorded
strength. This fact accounts for most of the reported strength differ-
ences of the same adhesive when used with different substrate metals.
Foulkes et al. (1970), using a rigid napkin ring test piece in torsional
shear, showed identical breaking torques on stainless steel, titanium
and aluminium recorded from -60°C to + 150°C with Redux 775
(Ciba-Geigy; a polyvinyl formal-P/F adhesive). This would, of course,
be expected if fracture occurred within the adhesive layer under
identical stress distributions. Whilst reported differences are, therefore,
mainly an artifact of the test procedure, there will be an effect if
surface preparation, temperature, or other conditions lead to crack
initiation at the metal-adhesive interface.

High Temperature Metal-Metal Adhesion


Figure 105 (Wake, 1982) illustrates the temperature limitations of
epoxy resins based on bis-phenol A. The two comppnent polyamine
cure shows relatively poor performance but these figures refer to
torsional shear and the two one-component epoxies here exhibit
strengths which would cause gross distortion of lap-shear test pieces of
the usual thickness. None is usable above 120°C. By contrast, Fig.
106 (Wake, 1982) shows the very different behaviour of an epoxy-
phenolic adhesive supported on a glass cloth carrier. There are two
features to be reckoned with when considering the use of adhesives at
elevated temperatures. There is the temperature at which they can be
used continuously and at which the joint will retain adequate strength
194 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

80,----------------------------------,

70" ~:'
60- '.

~ ~
N '''.

50- • Aroldlte 111

;
:;40 \
.. -\
\
~ 30- Aroldlte .~ -: ..

.
f" AYI05/HY953F. \-
20- • \
.~
10~ .~ .~
e_. - - 0 _0
a_ ~-Ä
..\ ....
"-

~ I J ~.~p.~..
o 20 40 60 80 100 120 140 160

FIG. 105. Strength-temperature profiles of Araldite (Ciba-Geigy) epoxy adhe-


sives on aluminium (Wake, 1982).

for a time measured in days or months, and the temperature at which


the joint retains its integrity either because it is only exposed momen-
tarily to such temperatures or because the time during which it is to
function is measured in minutes or hours only. In both cases the presence
or absence of air may greatly influence the durability of the bond.

;---..~
40- • .~ •

.,.......
N

... .. ..
'E .~

~ 30 .-".!
o
~ ~
<1> ~.
-'"
<J) 20- ~.

\
oc
o
'iii
~ 10-

I •
20 100 200 300 380
oe
FIG. 106. Strength-temperature profile of an epoxy-phenolic adhesive on
stainless steel (Hidux 1197A, Ciba-Geigy) (Wake, 1982).
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 195

TABLE 18
PERCENTAGE RETENTION OF STRENGTII OF JOINTS EXPOSED TO HEAT FOR 1h

Adhesive Temperature (oC)


100 150 200 250 300 350 400 450
Nitrile-phenolic 41 30 21 4 3
Epoxy-phenolic
Hidux 1197A 84 63 59 49 42
various 88 74 60 50 42*
Polybenzimidazole,
lmidite 850 96 93 90 85 79 73 65 56
Polyimides
FM34 99 94 87 79 66 49
Skyguard 700 87 77 68 57 42 24
Isomid 87 78 70 57 32
Polyquinoxaline 99 97 95 93 90 70 50 43t

* Data from Buck and Hockney (1969) indicate that whilst film supported on
glass cloth shows this strength retention it is not equalied by two-part paste
adhesive.
t From data which gives 40% retention after 10 min exposure to 537°C.

Table 18 gives data appropriate to a substrate of stainless steel,


which, although derived from actual measurements on specific com-
mercial products that are not always identified, may be taken as
indicating the behaviour to be expected from adhesives of the same
chemical type. The data on which the table is based were obtained by
Buck and Hockney (1969), Shields (1967), Maisey and Wake (1970)
and Hergenrother and Levine (1970). From data and comments given
by Aponyi (1977), it appears that as the polybenzthiazoles and poly-
benzoxazoles become available, their temperature performance will be
very similar to that of the polyquinoxalines.
A number of adhesives undergo further cross-linking on prolonged
heating and this gives an increase in strength before the inevitable
decline sets in. Thus, nitrile-phenolics increase in strength for the first
3000 h of heating at 150°C. Some varying time after this, probably
dependent on the ease of access of oxygen, the adhesive becomes
brittle. It is therefore meaningless to quote a figure for the time taken
for the strength to decay to 50% of its initial value. Suffice to state that
a nitrile-phenolic adhesive will perform satisfactorily for 3000 h at
200°C. Most of the epoxy-phenolics show a decline in the strength of
lap-shear joints on prolonged exposure to high temperatures but one
196 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 19
TIME OF EXPOSURE TO ELEVATED TEMPERATURE FOR 50% LOSS IN JOINT
STRENG'IH

Buck and Adhesive Time at Time at Time at


Hockney 150°C 200°C 300°C
code (h) (h) (h)

D Epoxy-phenolic 3300 >3000 Becomes brittle


(2-part paste)
E (glass mat 2000 «200
supported)
F (glass cloth 2800 c.300
supported)
G (2-part paste) >3000 >3000 Becomes brittle
Polybenzimidazole
(Imidite 850) 680*
Polyimides
FM 34 >20000 2000
Westinghouse 18 10 6 105 1 500 (extrapolated)
Westinghouse 140 3000 1000 200-300

* At 260°C

of those examined by Buck and Hockney showed an increase of 67%


at 200 h followed by a decline. It is possible that the curing schedule
of this material, which was carried on glass cloth, may be understated
in the manufacturer's instructions. Table 19 gives the number of hours
during which joints of stainless steel sustained at least 50% of the
initial strength at the stated temperature. The times must be regarded
as estimates and refer to a stainless steel substrate. Similar behaviour
occurs on titanium alloys.
It should be appreciated that the adhesives listed in Table 19 are
formulated by their manufacturers to meet a number of requirements
of which retention of strength at elevated temperatures is but one. It
follows that the adhesives showing the greatest strength retention at a
given temperature are not to be regarded as the 'best' in all cir-
cumstances.

STRUCTURAL ADHESIVES FOR WOOD

A wo oden chair is an engineering structure in that it must be designed


and assembled so that the necessary joints are capable of standing the
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 197

stresses imposed during use. The whole must preferably remain rigid
without movement of members in their joints. The stresses irnposed
can be quite high, particularly if the chair is misused by tilting it back
onto two legs. Of course there are numerous examples of glued tirnber
structures, some involving structural plywood which are truly engineer-
ing structures, e.g. the passenger reception hall at Southampton docks
is a major load-bearing tirnber structure known to the authors.
A number of adhesives superior to the traditional carpenters' glue
are available for furniture, including polyvinyl acetate emulsion, urea-
formaldehyde (V/F) and polychloroprene formulations. Any of these
adhesives would be entirely satisfactory for use with furniture but they
would not be used in outdoor engineering structures. Partly, this is
because of susceptibility to moisture but also these materials are not
intended for conditions where they are exposed to sunlight, rain,
temperature cycling and the other effects that make the more rigorous
atmospheric weathering. Deterioration with such exposure would
occur in months rather than in years. However, a heat-cured
resorcinol-formaldehyde (R/F) adhesive would outlast the tirnber.
Wood adhesives are classified into groups according to their resistance
to deterioration by natural and artificial agencies. For structural en-
gineering use, most applications will require an adhesive taken from
the WBP group, i.e. weather and boil proof. The accepted groups and
the adhesives covered by them are given in Table 20.
TABLE 20
DURABTLITY GROUPS OF WOOD ADHESlVES

Group Type o[ adhesive Indication o[ struc-


covered tural application

WBP (weather Resorcinol- Joints fully exposed to


and boil proof) formaldehyde and weather, sunlight or
phenol-formaldehyde marine situations
(R/F and P/F)
BP (boil resistant) Melamine- Limited life if fully ex-
formaldehyde and posed. Interior use in
melamine modified high humidity atmos-
urea-formaldehyde pheres as in laundries,
(M/F and M-U/F) dyehouses or baths
MR (moisture resis- Urea-formaldehyde of
tant) appropriate U/F ratio
INT (interior use Urea-formaldehyde; ca- Not discussed in the pre-
only) sein; polyvinyl ace- sent context
tate; animal glues
198 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

1t is obvious that only three types of adhesive are on ofter from


Table 20 although eight adhesives are given in Table 15 as being of use
with wood. The reason for the discrepancy lies in the conditions of use.
Table 20 is to be interpreted as referring to joints made wholly of
wood. The adhesives listed in Table 15 but not in Table 20 are those
which will adhere to wood and can be used in assembling mixed joints,
but used as wood-to-wood adhesives they either do not give optimum
properties or durability or are noticeably more expensive for no gain in
properties. The choice will depend on the member to be seeured to the
wood. Depending on the stresses and environmental conditions, adhe-
sion to wood may become the weak feature of the joint. Thus, owing
to the moisture susceptibility of the wood, a joint with a metal and
seeured by an epoxy adhesive would be expected to have a more
limited life than a metal-epoxy-metal joint or wood joined to itself by
one of the preferred adhesives.
A check list for wood adhesives involves the nature of the wood
insofar as this influences the manufacturing conditions.

Check-list for the Use of Structural Wood Adhesives

Condition of Timber
Moisture content. The usual recommendation is that wood shouldbe
dried to 12-15% moisture but, with resorcinol- or phenol-formal-
dehyde adhesives, this requirement can be relaxed somewhat and
wood of 20% moisture content will show only the reduction in joint
strength expected from the associated reduction in the strength of the
timber with the increase of moisture content (Laidiawand Paxton,
1974). Some melamine modified urea-formaldehyde adhesives are also
satisfactory under these conditions.

Presence of preservatives. Creosote oil, provided it is not present to


excess so that it appears on the surface of freshly machined or sanded
wood before the adhesive is applied, is not deleterious to RjF or PjF
adhesives. Timber preservatives based on pentachlorophenol or tin,
even if up to 1% wax has been added, do not influence these materials.
Fire retardants based on ammonium salts may interfere with the
cross-linking RjF and PjF adhesives such that recommended periods of
eure may have to be extended. M-U jF adhesives are adversely aftected
if wax is added with the preservative.
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 199

Density. Although with structural adhesives failure is normally a


failure of the timber itself, there is evidence that although the strength
of wood increases with increasing density, maximum adhesive strength
is reached with a wood of a density 650 up to 950 kg.m- 3 (0,65-
0·95 g.cm- 3 ) (Douglas and Pettifor, quoted by Gray, 1967). Above this
range, joint strength falls because the adhesive no longer wets the
wood adequately.

Nature of the timber. Some manufacturers of two-part P/F adhesives


recommend different hardeners for woods of different botanical origin.
There is also evidence that some hardwoods, such as teak and afror-
mosia, influence the setting time and final strength of R/F and other
condensation-type adlhesives (Chugg and Gray, 1965). In general, a
surface freshly prepared by fine sawing, sanding or planing is the ideal
surface for making the joint. If sawn or sanded, it must be weIl brushed
with a soft brush to remove dust. Roughing the surface by vigorous
brushing or the use of rasps or wire brushes is most undesirable and
leads to weak joints. The surface of teak must be covered with
adhesive immediately after its preparation.
Whilst it is undesirable to use old timber, it is sometimes unavoida-
ble; also, repairs to old timber may have to be undertaken. Under
these circumstances it is important to remove the old surface and not
merely to sand it. Sawing ensures that a sufficient depth is removed to
give a new surface, in spite of a reluctance to use saws on old timber.
Also it must be remembered that new adhesive does not stick to old
adhesive.

Strength Requirements

Shear strength. May be taken as exceeding that of the wood assum-


ing edge grain bonding.

Cleavage. Failure should occur in the wood.

Impact strength. The wood itself should absorb energy on impact


and the joint so designed that, as with the wood itself, impact does not
cause a shear stress along the grain. Impact should be borne across the
grain and across the glue-line. If impact along the joint is to be
absorbed, then consideration should be given to the use of an epoxy-
polysulphide adhesive.
200 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Deformation
The deformation of the joint will normally be that of the adherends,
but torsional shear of the gusset plate joints in light wooden trusses
may lead to an appreeiable inerement in the total defleetion. A phenol-
resorcinol-formaldehyde adhesive joining Douglas fir adherends
showed a modulus of 0·93 GN.m- 2 (Krueger, 1962).

Manufacturing Conditions

Bonding pressures. All joints require pressure to hold the joint


during eure but some require relatively high pressure (e.g.,
15 MN.m- 2) due to the elimination of water by chemical reaction. This
may be inconvenient if an assembly is to be eured on site.
Requiring high pressure: Rot-euring, one-part P/F adhesives.
Requiring moderate pressure: P/F one-part and M-U/F
one-part adhesives.
Requiring holding pressures
only: R/F and acid catalysed P/F adhesives.
Bonding time/temperature. The temperature required for curing in a
reasonable time and the relation between temperature and time of
eure varies with the formulations used.
Temperatures in excess of 105°C: One-part P/F, one-part
M-UjF and hot-setting R/F
adhesives.
Temperatures in the range 20-60°C: R/F and acid catalysed P/F
adhesives.
The times of eure vary from one day at room temperature to about
2 min at 140°C.

Pot life. One-part adhesives have indefinite working life at room


temperature, whereas the two-part adhesives have a life limited from
1-8 h at room temperature once the parts have been mixed. This pot
Iife is greatly influeneed by the pR of the formulation. In the ease of an
R/F adhesive, the most desirable state of the material with a pR just
on the alkaline side of neutral is that which gives the shortest pot life.

Form 0/ adhesive. PjF and M-UjF are usually supplied as films on a


tissue carrier, though liquids can also be used. R/F resins are usually
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 201

liquids. Two-part adhesives have the main component as a liquid and


the curing agent or catalyst as a powder. The liquids are usually 'A'
stage resins, but are sometimes dissolved in alcohol which must be
evaporated before the joint is closed.
Durability of Joint
The most durable joint is given with a hot -cured R/F adhesive, but this
is also the most expensive. Although these adhesives have not yet been
in use for fifty years, they are expected to survive exposure to weather
and stress for such periods. Almost as good are the P/F-R/F mixed
types of adhesive. P/F adhesives, whether hot or cold cured are
expected to survive more than 25 years in conditions of full exposure
whereas, in such circumstances, an M-U/F adhesive would have a
much shorter life. Even so, this would be expected to reach 10 years if
the manufacturer offered the material as suitable for high hazard
locations. In this connection, attention is drawn to the wrong use of a
one-part U/F adhesive, not fortified with melamine, in the construction
of the roof of a swimming pool (Mayo, 1975). This structure collapsed
only 12 years after completion owing to, among other causes, either a
tension failure at a scarf joint or failure between a plywood web and
the bottom chord of a roof beam.

STRUCTURAL ADHESIVES FOR MIXED CONSTRUCTIONS

Metal-Wood Structures
The first recorded use of a modern structural adhesive was in aircraft
construction for joining aluminium alloy and wood (Garnish, 1977).
The adhesive used was a polyvinyl formal/phenol formaldehyde mix-
ture. This adhesive is still used although in this limited application other
materials, notably epoxies, have gained ground at its expense, there is
little evidence which enables a rational choice to be made in favour of
either. However, manufacturing circumstances can make it convenient
to prime wood immediately after preparation and to carry out the
bonding against metal at a later stage, perhaps after some period of
storage so that the wood is held awaiting the preparation of the metal.
In these circumstances it is on balance better to prime the wood with a
phenol formaldehyde resin and then to bond the metal to this surface
with an epoxy resin. An alternative is to prime the metal with a
polyvinyl formal-P/F adhesive such as Redux 775 (Bonded Structures
Ltd) and then to bond this to the wood with a phenol formaldehyde
202 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

resin. The possibility of high stresses at the ends of a mixed metal-wood


joint owing to moisture changes in the timber suggests that greater dura-
bility would be obtained with an adhesive giving a higher peel strength
than a straight epoxy and, in particular, suggests the use of epoxy-
polysulphide adhesives and reinforces the use of the original polyvinyl
formal-P/F combination.
There is as yet no published information on the use of the semi-
structural, toughened acrylic formulations in mixed metal-wood struc-
tures. Depending on the stress level required, these adhesives may be
perfectly satisfactory in situations where exposure to weathering is not
required.

Metal-reinforced Plastics Structures


The reinforced plastics involved are glass reinforced polyesters and
epoxides, together with the structurally more important carbon fibre
reinforced epoxides and p-xylene resins.
Bearing in mind that a polyester-based GRP is unlikely to be used
for the highest structural duty, the semi-structural toughened acrylics
would make an easily used, non-demanding, two-part adhesives which
should suffice most applications.
As the stresses which arise in metal-wood bonds from the absorp-
tion of moisture by the wood weaken the joint, so do stresses from the
mismatch of the coefficients of linear expansion cause trouble with
metal-CFRP joints. These are particularly important when a high
temperature has been used for curing but, even with the curing at more
moderate temperatures, the expansion mismatch can become serious at
sub-zero temperatures, as for example is experienced in high ftying
aircraft. These effects have been studied by Stone (1980) for
toughened epoxy and a curable acrylate adhesive with
CFRP/Aluminium and CFRP/Titanium. Whereas with metal-metal
adhesion the double-lap shear joints he used showed the usual increase
in strength with a lowering of the temperature from normal to -55°C,
the mixed joints showed a fall which, in the case of the acrylate
adhesive, would be unacceptable for design purposes. The mismatch,
and consequent loss of strength, is less for titanium than for
aluminium.

CHOICE OF ADHESIVES FOR SEMI-STRUCTURAL USE

Apart from structures bearing major loads, there are many instances
where a fastening or joint transmits or sustains a moderate load that is
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 203

within the capability of a large number of adhesives. If the application


is a small one, the use of an expensive adhesive may still be trivial and
for convenience tolerated, but in a long production run the cost of
materials becomes an important factor. Although the structural adhe-
sives reviewed in previous sections are possible in many of the situa-
tions covered by the requirements set out below, they are omitted from
consideration because they are generally more expensive and their
high strength and special properties are not required. This is especially
true of the epoxies which can be used in a wider variety of situations
than any other adhesive and hence are frequently used on grounds of
convenience when a less well-known but cheaper material would be
just as satisfactory. Exceptions to this comment on cost are the
inclusion of references to anaerobic and ionically polymerized acrylics
which are more expensive than epoxies volume for volume but which,
because of their special properties, are used in circumstances where
their economy of use brings an overall gain. A check list follows based
on the adhesives given in Table 14 but excluding the structural
adhesives.

Check-list for Adhesives for Semi-structural Use

Surface to be Stuck

Metallic. Metals have high energy surfaces because of the oxide


layer they carry. There is therefore very litde difficulty in getting
almost any adhesive to stick to them provided they are clean and free
from loose or incoherent oxide. The exceptions which give rise to
difficulty are copper, zinc and cadmium because of the nature of their
oxides and re action with components of some adhesives, and gold,
platinum and other noble metals because they have no surface oxide.
Copper forms an incoherent black oxide. A newly etched copper
surface forms a good bond with most adhesives but, owing to the
diffusion of oxygen from the edges of the joint, the life of the bond
can be quite short at only moderately elevated temperatures. Other,
equally deleterious corrosion products may be formed with polychloro-
prene adhesives.
Copper surfaces may be stuck with urea formaldehyde resins or, if
these are too brittle to suit the application, they may be used as a
priming coat above which some other adhesive is applied. Epoxy resins
vary in their behaviour towards copper owing to differences in the
curing agent employed. Anhydride curing agents are known to be
204 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

undesirable but some amine-amide compounds are also suspect. If


copper is involved with an epoxy adhesive the manufacturer of the
adhesive should be consulted. Priming with a siloxane coupling agent
(see Chapter 6) might be advantageous. The copper alloys have, in
general surfaces in which the alloying metal provides the oxide coating.
Brass, for example, has a coating of zinc oxide.
Zinc surfaces may be involved either by the use of galvanized steel
or of zinc die castings. Again, anhydride-cured epoxies should be
avoided as also should adhesives which contain tackifiers based on
rosin derivatives. These rosin materials tend to form soaps with zinc
and to form them at the interface between metal and adhesive thus
creating a weak boundary layer. Polychloroprene adhesives are recom-
mended for use with zinc as the tackifier used is normally a tert-butyl
phenolic resin which is perfectly safe, although some products may
contain other, less satisfactory, tackifier.
Small die castings are conveniently bonded with polymethylmeth-
acrylate dissolved in its monomer together with a peroxide initiator.
Metal naphthenates are to be avoided as accelerators and hence the
so-called second generation or toughened acrylics are not recom-
mended for use with zinc. Other acrylics may be used but are slow
curing after evaporation of solvent. Polyester adhesives cured with
styrene monomer frequently contain some free phthalic acid as an
impurity and this militates against their use. Polyurethane adhesives
cured with iso cyanate should behave quite satisfactorily.
Cadmium-similar considerations apply to cadmium as apply to
zinc. Cadmium plating is sometimes chromated but this provides no
protection to interactions of the type described.
Gold and other noble metals-pure gold is very ductile and will not
be used in a stressed state. Attachment to other surfaces IS conve-
niently efIected by a polycyano-acrylate.

Non-metallic. Wood-this material has already been discussed in


connection with structural adhesives. Where both loads and conditions
are less demanding, the available adhesives multiply. Polychloroprene
and polyvinyl acetate, the latter as an aqueous emulsion, are widely
used, the former accommodating a wider range of acceptable substrates
when joining wood to some other material. Casein glues can be
formulated to give water resistance and are still used for the manufac-
ture of plywood, web-strengthened Bat doors, and beams for interior
use. Natural rubber latex forms with casein a valuable adhesive with
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 205

good peel and impact resistance, especially when formulated with


formaldehyde or chromates to cross-link the casein.
Plastics materials-compared with metals and wood, all plastics
materials have relatively low energy surfaces and some, the hydrocar-
bon polymers such as polyethylene or polypropylene together with the
fluorinated polymers, have surfaces of very low free surface energy and
are difficult to bond. Mostly the surfaces can be treated to render them
capable of being bonded by epoxy, polyurethane or polychloroprene
adhesives. Other, thermoplastic plastics such as polyvinylchloride
(PVC), polysulphones or polycarbonates are best solvent welded. In this
process, a solvent for the material is allowed to soften the surfaces to
be bonded after which they are brought together and held firmly in
contact until diffusion of the polymers has obliterated the interface and
the solvent has diffused away from the site of bonding. Strength
continues to improve for several days while the solvent is eventually
lost to the atmosphere. For bonding the plastics named above, the
solvent used is methylene dichloride; for PVC this advantageously
contains 5-10% chlorinated rubber or chlorinated PVc. The process is
discussed in detail by Titow (1978).
Glass-the use of polyvinyl butyral (PVB) in the manufacture of
laminated glass for car windscreens suggests its use as an adhesive but,
in general, attaching glass to other substrates, unless a large area is
involved in which the gap-filling properties of PVB would be neces-
sary, recourse should be had to epoxy, anaerobic acrylic adhesives or a
cyanoacrylate. The last named is a thermoplastic, whilst the other
materials are thermosetting. For some production purposes, the
an aerobic adhesive formulated to cure by ultraviolet radiation provides
a convenient method of obtaining a bond of high strength which,
having been formed near room temperature, is more free from internal
stress than one formed at high temperature. Butt joints with a steel
cylinder bonded to a glass plate showed a tensile strength in direct
tension of 14 MN.m- 2 (Shields, 1977). Ringes have been bonded to
the glass doors of domestic cooking ovens by UV activated adhesives.
Joints involving glass and made with epoxy, cyanoacrylate or anaerobic
adhesives benefit in a moist environment both in strength and durabil-
ity if the glass is treated with an appropriate siloxane (see Chapter 6).

Method of Application of Adhesive


A number of structural adhesives are supplied as solid films, frequently
carried on a tissue or cloth to maintain coherence. Some hot melt
N
o
0\

SCHEME 1
CI'l
Liquid adhesive ;d
c
L Solution in organie solvent
I
r
I ~
Remains thermoplastic v>
on removal of solvent Cross-links on removal of solvent
Other (one-part solution) (two-part solution) 51
CI'l

liquids ~
(i) Polychloroprene contact (i) Polychloroprene or poly- '-
adhesive urethane plus isocyanate o
(H) Fully reacted polyurethane (one-part solution) ~
(ii) Nitrile-phenolic (requires
high temperature and press ure
Z
tn
to complete cure) Z
f-------_------Aqueous emulsion 8
z
(i) Polyvinyl acetate gJ
~
Other (ii) Sulphonated polyurethane z
Iiquids (iii) Latex modified poly- Cl
chloroprenes

I) 100% liquid adhesive at room temperatures

I
1- ._-.-. -·-----1
Viseous liquid or paste Freely flowing liquid
~________~I__~ I
~
I I Remaining Crosslinking q
One-part TwoTpart thermoplastie D" 1 o;>J
I I Iacry ates C/)

Heat cured RT eure Heat eure Cyano-


epoxies acrylates ~
(i) Epoxies (i) Epoxies c:
(ii) Polymer (ii) Polyester ~
containing plus styrene oz
acrylic Cl
monomer
(iii) Polyester ~
tn
plus isocyanate g
(iv) Polyester o
plus styrene n
tn
plus accelerator o"tl
>tJ
~
'tn<!

N
Cl
--.l
208 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

adhesives that might be used in semi-structural applications are availa-


ble as filmic adhesives and the polyvinyl butyral used for laminating
glass is supplied in sheet form. All other adhesives are applied to the
substrates as liquids and the type of liquid is, to some extent, dictated
by the polymer of the adhesive. If there are restraints on the type of
liquid which can be or is convenient to use, Scheme 1 will relate these
to the adhesives possibilities.

Rate of Setting or Curing


Many adhesive types can, by alteration of the precise chemistry of the
basic polymer or by changing the formulation by adding catalysts or
accelerators, show a very wide variation in the time required to achieve
load-bearing strength. Additionally, where curing depends on a chemi-
cal reaction rather than mere loss of solvent or crystallization, it can be
greatly hastened by an increase of temperature.
Thus the quoted data for a commercial acrylic depending on
polymerization of a monomer or mixture of monomers gives 2 days at
15°C or 1 h at 80°C, both with 2% catalyst; however, an increase of
the catalyst to 5% reduces the eure time at 80°C to 25 min. R/F
adhesives may, in the presence of a catalyst, require several hours at
room temperature while curing in a few minutes at the moderately
elevated temperature of 80°C. By working at, say, 105°C this becomes
a fast setting adhesive with a eure time of about aminute. Scheme 2
can therefore provide only a rough classification.

Gap Filling
In general, with a structural adhesive, the thinner the glue-line, the
stronger the joint. This may be accepted as applying to all adhesives
which are, when loaded in use, operating near or below their glass
transition temperature. Rubbery adhesives, that is those operating
above their glass transition, are much less sensitive to glue-line thick-
ness, though, of course, the deformation of the joint under load will be
influenced by it. It is not always possible to ensure a uniformly thin
glue-line, particularly with substrates such as timber, glass or concrete,
and adhesives which perform weIl under these conditions are referred
to as having good gap-filling properties. To a large extent this arises
from the ability of the adhesive to accept a high volume proportion of
finely divided powder without excessive loss of strength. Gap-filling
properties are, therefore, partly in the hands of the formulator who is
SCHEME 2
Rate
I
I I ----1
Immediate Fast Siow
(At RT or elevated temperature,
I I ~I __- L_ _ _ _ _ _I
I wh ich ever is the recommended
No prior Prior RT On procedure) (or post-cure.needed.)
application application and Anaerobics heating AtRT
Cyano- drying Toughened RjF PjF and RjF
acrylate Contact acrylics Solvent bon ding of
(substrate adhesives Cyanoacrylates certain plastics materials
dependent) e.g. Nitrile phenolics
Some Polychloroprene Acrylics with polymer
anaerobics Polyurethane dissolved in monomer but
Hot melts Nitrile without accelerator
rubber Casein and casein-latex
Polyvinyl acetate
210 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 21
GAP-FILLING PROPERTIES OF ADHESIVES

Substrate Not gap-jilling Gap-jilling


(as usually (or easily formulated
formulated) with this property)

Metals Anaerobic acrylic Epoxies


cyanoacrylate PVC plastisols with diallyl
phthalate
Wood RlF, M-UF (particularly if P/F, P/F-RlF (see Vick,
radio-frequency heating 1973)
is used) Acid catalysed phenolics
P/F (hot setting) PVA
General Polyesters with polymeriz- Most hot melt formulations
ing styrene monomer Nitrile-phenolic

able to provide many adhesives in alternative guises. This is partieu-


larly so for the epoxide and phenol-formaldehyde adhesives. The
former not only show a low shrinkage on curing but they can be
supplied with silica powder incorporated to a paste consistency with
very little loss in strength. The latter can be compounded with wood
flour into which the resin diffuses to some extent and with which it
undoubtedly bonds firmly. The toughened adhesives made by incor-
porating or precipitating a second polymerie phase tend to be better at
gap-filling than single phase systems.
With these general observations, some indications of possible
choices for certain substrates are given in Table 21 together with lists
of adhesives which are normally regarded as falling definitely into one
or other of the two categories.

High and Low Temperature Tolerance


The temperature properties of metal-metal adhesive, insofar as they
involve considerably elevated temperatures, have been discussed
elsewhere. For metals, therefore, it is necessary to extend the discus-
sion to sub-zero temperatures and to consider the upper working limit
for the more common adhesives. The low shrinkage which epoxy res ins
show in the cross-linking process is advantageous in reducing internal
stress in the joint and hence allows the possibility of sustaining greater
stress due to differential shrinkage of adhesive and adherend when the
temperature is lowerec!. Epoxies are at a further advantage in that
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 211

their coefficient of thermal expansion, although several times those of


metals, is among the lowest of the values for polymers and it can be
further modified by filiers. In general, the shear strength of adhesive
bonds, as distinct from joints, increases with decreasing temperature,
but the joint strength may decrease if it is a single-lap shear joint
whose strength is measured. Thus, Merriman (1965) reports that the
peel strength of epoxy-phenolics remain fairly constant up to 121°e
but that the shear strength increases slightly from -73° to 65·5°C. This
increase in shear is an artifact of the dimensions of the test piece. A
number of examples exist in the trade literature which show an
increase in shear with decreasing temperature, rising to a maximum
and then falling again at still lower temperatures. Working with very
rigid adherends forming a torsional shear (napkin ring) test piece,
Foulkes et al. (1970) examined the behaviour of a vinyl formal-P/F
adhesive over a wide range of temperature and showed a smooth
transition from a high, low-temperature value, falling through the glass
transition temperature to a low value tailing to zero as the temperature
became too high for the adhesive. Similar behaviour is summarized by
Wake (1982) br a number of structural adhesives. As failure at low
temperatures is invariably cohesive within the adhesive resin, this
behaviour accords with what is known about cohesive strength. How-
ever, as the temperature decreases, susceptibility to sudden brittle
fracture increases and the toughness of many polymers at temperatures
immediately below the glass transition temperature is lost. Shields
(1976b), in his very useful tables of the properties of adhesives, gives
in the form of bar graphs the continuous heat resistance temperature
range as starting from ab out - 300 e for most adhesives with -500 e for
the epoxies including the epoxy-polysulphide adhesive which is usually
regarded as the most impact resistant of the epoxy combinations.
These figures should be taken as conservative and may, again, reflect
test procedures since epoxies have been successfully used at cryogenic
temperatures. Sims (1972) gives a bibliography of such uses. Obvi-
ously, materials which are rubbery at normal temperatures and pass
through the glass transition at temperatures below ooe are used in
cryogenic applications. In particular, the polyurethanes are frequently
preferred but joint strength at room temperature does not suffice to
bring them into the dass of adhesives considered here.
The upper end of the temperature scale for adhesives exposed more
or less continuously to elevated temperature is certainly below lOOoe
unless special high-temperature adhesives are used.
212 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Phenolic and resorcinol adhesives used with wooden structures are


usually only exposed to extremes of temperature by virtue of outdoor
usage. There appears to be no information on failures in arctic or
antarctic conditions, plywood structures being used extensivelyon the
antarctic continent. At the other end of the scale, Knight (1959)
suggests that in England, although temperatures of 65°C do occur on
wooden structures exposed to direct sunlight, the frequency of occur-
rence is small and the duration short. He was more concerned with
deterioration of the adhesive than a temporary change in load bearing
capacity, but the latter may be assumed negligible and about the
former he states that sunlight has, in England, no effect at all over a
period of 10 years. An earlier report from the US Forest Products
Laboratory on Temperature and moisture content in wood aircraft wings
(quoted by Soden and Wake, 1951) gives a temperature of 102°C
recorded on the dark upper surface of a grounded aircraft at Tueson,
Arizona, but there is no suggestion that modern adhesives for wood
cannot withstand such conditions. Durability under tropical conditions
is another matter involving, as it does, the timber probably more than
the adhesive.

Water Sensitivity
Organic polymers all absorb moisture and those which under equilib-
rium conditions with 100% relative humidity (RH) or actual immer-
sion in water absorb less than, say, 1% may be regarded as insensitive.
Insofar as in a weH-made structural joint failure is in the adhesive
rather than at the interface, and the absorption of water williower the
cohesive strength of the adhesive, it would be expected that strength
would decrease but not greatly so, with increased uptake of water. The
facts are rather different to this simple picture. Where the moisture
uptake exceeds a few percent, failure often transfers to the interface,
particularly when the joint is under stress, and a catastrophic drop in
joint strength is found. All polymers absorb greater quantities of water
when above their glass transition temperature and hence rubbery
materials tend to show greater water absorption than rigid resins such as
the thermosetting adhesives. When an adhesive is formulated as an
aqueous emulsion or latex, the stability of this formulation depends on
the presence of emulsifiers, thickeners and freeze-thaw stabilizers, all
of which are hydrophilie chemieals adding greatly to the water absorp-
tion of the adhesive. Alliatex-formulated adhesives must therefore be
regarded as water sensitive.
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 213

The key position in structural metal-metal adhesives is occupied by


epoxy-nylon adhesives. These materials have been used over many
years with undoubted success and safety in aircraft flying in tropical
locations; however, under test conditions involving high humidity, they
have performed very badly. The inference to be drawn seems to be
that the conditions inside an aircraft are not as severe as those outside in
spite of the amount of condensed moisture commonly found between
inner and outer skin. Butt and Cotter (1976) found an equilibrium
water absorption of 14% for an epoxy-nylon exposed to 97% RH and
that the real part of the dynamic tensile modulus of a strip of the
adhesive decreased from 2·27 to 0·337 GN.m- 2 , a decrease to 14% of
the original value. This low value was maintained during the 85 days of
the experiment. On drying out, the initial value of the modulus was
substantially regained but loss in joint strength proved irreversible by
drying as it was an interfacial phenomenon.
The polyvinyl butyral-P/F adhesives show a higher water absorption
than the polyvinyl formal-P/F combinations, but the behaviour of the
latter under hot, wet tropical conditions, although not so catastrophic
as the epoxy-nylon adhesive mentioned above, was such that it must be
regarded as water sensitive (Cotter, 1977).
Among the high temperature metal-metal adhesives, polyimides
are reported by Shields (1976c) as showing water absorption in the
range 1· 5-3% and would therefore be regarded as suspect. He
elsewhere (1976d) gives them 'excellent' for resistance to hot and cold

TABLE 22
WATER SENSITIVITY OF ADHESIVES

Insensitive Somewhat sensitive Very sensitive

R/F, PIF, epoxy Polyvinyl formal-P/F PolYJlrethanes


and epoxy com- Polyvinyl butyral-P/F All aqueous-based adhe-
binations unless Polyesters with or with- sives
separately spe- out added isocyanates U/F
cified M-U/F Epoxy-nylon
Nitrile-phenolics Polyimide Polyvinyl-acetate
Toughened acrylics Ethylene-vinyl acetate Cyanoacrylate
copolymers
Polychloroprene
Toughened acrylics
when applied to oily
surfaces
214 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

water although behaviour in moist environments has not been re-


ported. As a guide and summary of the above discussion, Table 22
allocates adhesives types into the categories they would occupy in the
absence of specific formulation designed to minimize the sensitivity to
moisture.

Pot Life
This subject has already been discussed in connection with structural
metal and wood adhesives. It is intimately connected with whether or
not the adhesive is supplied as a two-part material mixed before use or
whether any chemical reactivity is latent within the formulation and is
released by rise in temperature. Another release mechanism used with
adhesives, though not with structural materials, is the moisture curing
property of acetoxy slloxanes. Moisture may also limit the life of the
isocyanate part of a two-part polyurethane or an isocyanate cured
polyester which is, of course, a form of polyurethane. Hot melt adhesives
held in reservoirs at elevated temperatures are prone to oxidation and,
although actual time limits are difficult to quote, continually recharging
the reservoir without emptying and cleaning is a bad habit even when
the reservoir is held under ablanket of an inert gas.

Gil Tolerance
To distinguish adhesives Which are resistant to oils from those which
are not, it is necessary to separate aromatic containing oils from the
purely paraffinic, with the naphthenic olls somewhere in between.
The cross-linking adhesives need to be considered separately from the
thermoplastic materials because cross-linking of itself imposes resis-
tance to swelling by any fluid with Which the polymer shows some
tendency to interact. A degree of oil absorption which in a cross-linked
adhesive might reduce strength but be tolerated would, in a thermo-
plastic material, cause joint failure by virtue of creep. Table 23 gives
guidance.
Table 23 assumes the adhesive to be in situ in a properly made joint
before being exposed to the oil. A special case is, however, worth
recording. This concerns tolerance of olls and greases on a metal
substrate before application of the adhesive. Under these conditions,
most adhesives will fail to make a satisfactory joint even though the
quantity of oil or grease present is small in relation to that of adhesive.
Some adhesives, notably polyvinyl chloride formulated as a plastisol
with a proportion of a cross-linking plasticizer, diallyl phthalate, are
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 215

TABLE 23
OlL TOLERANCE OF ADHESIVES

Adhesive Resistant Slightly Not


type affected resistant

Paraffinic Thermo- Polyurethane Ethylene-vinyl ace-


oils plastic Polyvinyl ace- tate HMA*
tate Cyanoacrylates
Polychloro-
prene
Polyester and
polyamide
HMA*
Cross- Epoxies Toughened
linked Diacrylics acrylics
Polyimide Polyvinyl
Polyure- formal-P/F
thane Nitrile phe-
P/F; R/F; nolic
M/F
Aroma- Thermo- As above together
tic oils plastic with HMA * gener-
ally
Polyvinyl acetate
Polychloroprene
Cross- Polyimide Polyurethane Polyvinyl formal-P/F
linked Epoxy poly- Diacrylics Toughened acrylic
sulphide Some epoxy Nitrile phenolic
R/F; P/F; combina- Some epoxy com-
M/F tions binations

* HMA is the abbreviation for hot melt adhesive.

capable of bonding to a greasy metal surface. The PVC plastisol is used


semi-structurally in low stress applications for bonding stiffeners to
sheet metal. Toughened acrylics and anaerobic (diacrylate) adhesives
are similarly tolerant though to a lesser degree. Subsequent exposure
to a greater quantity of aromatic oil is deleterious with alt these
adhesives.

Method of Use
Contact adhesives, hot melt, cyanoacrylate and anaerobic adhesives may
be used for forming instant joints which require no further restraint
after placement has been made. Other materials require jigging or
216 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

clamping or the use of a press. The latter is essential for polyvinyl-


formal-P/F adhesives, for heat-cured P/F and R/F adhesives, for COffi-
posite adhesives involving formaldehyde condensation products and
for high temperature metal adhesives such as the polyimides, poly-
benzimidazoles and quinoxalines.
A variant method of use involving permanent clamping of the
adherend surfaces is that of weld-bonding. Weld-bonding is the term
used to describe the fastening of two metal surfaces by adhesive
bonding followed by resistance spot welding through the adhesive film.

Weld-bonding. The design of joints in which more than one method


of fastening is introduced has not been properly studied and is not
immediately attractive. Wood joints which are nailed as weIl as glued
shows a lack of faith in the gluing procedure and, unless the nails are
merely very light tacking nails to hold the adherends steady until the
glue is set, the presence of the glue is superfluous. The load will be
borne by the nails as these provide the stifter connection between the
adherends. The combination of spot welding with adhesive bonding
was therefore received with some scepticism when it was first intro-
duced and was regarded as a method of holding the adherends, thus
enabling lower temperature cures to be achieved in ovens without
the expensive provision of jigs. Indeed this is one of the advantages of
the method. However, there are advantages for designers insofar as
stress concentrations can be reduced and fatigue properties improved if
the welds are correctly placed. Kizer and Grosko (1972) report briefly
results from finite-element analysis of lap-shear joints with aluminium
adherends 1·27 mm thick with 25 nup. overlap and a glue-line thickness
of 0·013 mm. The shear modulus of the adhesive was 689 MN.m- 2 . A
single spot weId, about 6· 3 mm dia. was placed on the centre-line of
the overlap in the centre, 3·8 mm or 6·35 mm from the centre.
According to their caIculations which assumed elastic response only by
the adhesive, the weId carried 2·1%, 4·1% or 8·4% of the load
respectively. Obviously, moving the position of the weId from the
centre of the overlap where the shear stress in a simple adhesive joint
is at its minimum, results in a higher proportion of the load being
taken by the weId. This results in some increase to the static strength
though it may be detrimental to strength decay caused by fatigue.
Most of the claims for weld-bonding derive from improved
economics rather than improved properties. Fields (1973) compares
costs against those for hand riveting showing that whereas adhesive
FACTORS INFLUENCING THE CHOICE OF ADHESIVE 217

bonding is 60-80% more expensive, weld-bonding costs only 20-25%


of the riveting cost. Metal surface preparation techniques must be
chosen as recommended for spot welding. Not all adhesives are suita-
ble and epoxies are generally favoured; in general, separate primers
are to be avoided. However, where the strength of the joint is
determined by the spot welding rather than by the adhesive bonding, a
PVC plastisol is recommended by Eichhorn et al. (1979) who discuss in
detail the efIect of specimen geometry on the strength of spot
welded/adhesively bonded joints. This idea is attractive because of the
tolerance of the plastisol towards oil and grease. In general, the main
purpose of the combination is that both methods of fastening should
bear a substantial load with the adhesive carrying the lion's share. In
these circumstances epoxies are essential as also is adequate surface
preparation and c1eanliness.

Need for Primer


The general question of priming coats is dealt with in the next chapter.
Its inc1usion here is therefore in the nature of areminder that any
check list should inc1ude adecision about the need for a primer when
choosing an adhesive for a given adherend. In particular, some nitrile-
phenolic and some epoxy-novolac adhesives require a phenol-
formaldehyde priming film on metal adherends if they are to develop
their full bonding power. Others have a sufficient proportion of P/F in
their formulation to achieve good adhesion although peel strength is
likely to be lower in such formulations. In some acrylic compositions
the necessary catalyst is applied as a primer to one of the two faying
surfaces.
Chapter 6

Surface Preparation

One of the few dis advantages of adhesive bonding as a method of


fastening is that the surfaces need to be clean and, whatever their
chemical nature, must be coherent, in the sense that they must not be
powdery or friable. All the metals in common engineering practice
possess an oxide surface, but not all oxides are coherent; some tend to
be of low density, flaking or powdering whilst continuing to grow at the
interface with the metal atoms. The purpose of surface preparation is
to remove gross contamination together with powdery or loosely
adhering oxide, and then, if necessary, to etch away the existing oxide
and to replace it with a thinner, harder and completely coherent oxide
layer. There are corresponding requirements relevant to wood and
other substrates. To the surfaces thus prepared, the adhesive can
frequently be applied directly, but sometimes the prior application of a
primer is desirable. The primer can fulfill any combination of the
functions listed below:

(i) A coating of primer applied immediately to a freshly prepared


surface serves to protect it until the bonding operation is
carried out.
(ii) It wets the surface more easily and thoroughly than the adhe-
sive itself. This may be achieved by using as the primer some
of the adhesive but dissolved in a solvent to give a solution of
much lower viscosity. It may be a solution of a different
polymer which wets the substrate and, when dry, is easily
wetted by the adhesive.
(iii) It may serve to block the pores of a porous surface thus
preventing capillary suction of adhesive away from the glue-
218
SURFACE PREPARATION 219

line. This is likely to occur with aqueous-based emulsion


adhesives even when thickened.
(iv) It can act as the vehicle for corrosion inhibitors, restricting
them to the surface where they are needed.
(v) The primer may be a coupling agent capable of forrning
chemical bonds with the adherend surface on the one hand and
the adhesive on the other. This does not necessarily add to
bond strength when dry but it prevents displacement of adhe-
sive by water.
(vi) The adsorption of the primer to the substrate may be so strang
that instead of being mere1y a physical adsorption it partakes
of the nature of a chemical bond even though the union
between the primer substance and the oxide does not corres-
pond to a chemical which could be isolated as such or is
known. Such adsorption is referred to as chemisorption to
distinguish it from the reversible physical adsorption. The
primer may itself have a surface of low surface free energy but
will be stable with a low energy adhesive whereas the adhesive
applied directly to the oxide surface would be displaced by
water. Chlorinated rubber seems to be chemisorbed to some
oxide SUrfaces and this may account for its efficiency as a
primer.
Preparation before the application of primer and/or adhesive differs
for different substrates and may be summarized as folIows:
Metals
Brushing to remove loose oxide, dust and dirt. Degreasing by solvent
wash, dip or wipe, followed by either: (a) shot, sand or grit blasting,
wet 0r dry; or (b) chemical etching; and/or (c) washing in detergent
solution followed by clean, possibly de-ionized water. (A suitable
detergent solution is 5 ml Teepol with 25 g trisodium orthophosphate
in 1 litre water.)
If dry shot, sand or grit blasting is used, then a further solvent
degreasing is advisable followed by (c) or directly by drying in warm,
clean air. If chemical etching is used, water washing, then detergent
followed by further clean water is necessary before drying.

Wood
A good, freshly planed surface should be ideal for adhesive bonding
except that machine planed surfaces are not always flat and may
220 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

require sanding to remove undulations produced by rotary planing


machines. Fine sanding or sawing must be vigorously brushed or air
blasted to remove dust. End-grain gluing, if necessary, requires special
care. The sawn surface must be sanded and examined to ensure freedom
from partially attached fibres from broked lumen walls. Old timber is
best resawn as sanding may not remove sufficient of the old, oxidized
and impacted surface.

Concrete
Bush hammering of surface followed by wire brushing and immediate
application of adhesive.

Glass or Carbon-fibre Reinforced Plastics


The surface as presented may need to be made more conforming with
the mating surface by sanding but if the 'as moulded' surface is
presented for bonding it should be solvent wiped to remove excess
moulding lubricant, grit blasted or sanded and then vapour degreased.
A technique frequently adopted, and to be commended, is the use of
peel plies. This is adapted from the manufacture of multilayer lami-
nates built from resin impregnated glass or carbon fibre plies ('pre-
pregs' or 'tapes') in which the final layer, instead of being glass or
carbon fibre, is a woven or knitted nylon impregnated with a rnuch
lower proportion of a resin which must be compatible with the resin of
the composite but not necessarily or even desirably the same. The
composite is press or mould cured or partially cured. When, and only
when it is to be bonded to its mating surface, the peel ply is stripped
off from the composite. Obviously, the success of this procedure is
dependent on the absolutely clean removal of the peel ply without
plucking of fibres frorn the composite matrix. In practice, some re-
sidual adhesive from the peel ply is tolerable.

SHOT, SAND OR GRIT BLASTING

For ferrous metals there is no reason why shot blasting should not be
used but, for other metals, grit blasting with alumina or carborundum
is preferable. For small scale work, although a surface of a different
type is produced, rubbing with emery, carborundurn or garnet paper is
acceptable. All these rnethods remove metal and can, in fact, rernove a
few thou from the surface. For accurately machined parts, therefore,
SURFACE PREPARATION 221

none of these methods are suitable but wet blasting with a fine alumina
which gives a polishing-cleaning action may be operated within the
required tolerances. Proprietary machines are available for wet blast-
ing which are used for cleaning precision moulds for rubber and
plastics mouldings and which may be trusted to remove the minimum
of metal.
Shields (1976a) in a valued and detailed survey of surface preparation
technique, counsels against the use of glass or metal beads in blasting
techniques as leading to peening of the surface. He also refers to the
need to match the grit mesh size to the metal being treated.

SOLVENT DEGREASING OR WIPING

There are four considerations which govern the choice of a solvent:


toxicity, flammability, efficiency for degreasing and cost. There is also
the factor of convenience in that it is desirable to have only one or
perhaps two different solvents available for degreasing, whereas the
optimum of efficiency, if this were the sole criterion, might require a
number to be available.
The solvent most commonly used for degreasing metals is tri-
chloroethylene known on the shop floor as 'trike'. The threshold limit
value (1LV) for this and for other solvents that have been used or
recommended is given in Table 24. The 1LV is defined as the
concentration in parts of solvent per million of contaminated air
(measured by volume at 25°C and normal pressure) below which
continuous exposure is considered to be without harmful effect. A high

TABLE 24
THE TOXICITY OF SOLVENTS: SAFE UMlTS AND DILUTION VOLUMES

Solvent Threshold Safe dilution


limit value volurne

Trichloroethylene (trike) 100 330


1,1,1-Trichloroethane 350 640
Methyl ethyl ketone (MEK) 200 1270
Acetone 1000 310
Ethanol (alcohol) 1000 390
Methanol (wood a1cohol) 200 2800
Iso-propanol (IPA) 400 740
White spirit (estimated values) 500 230
222 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TLV indicates a relatively safe solvent; greater than 200 ppm might be
taken as an indication of safety, whereas a value of 100 ppm or less
indicates that stringent action must be taken to prevent its escape into,
or use in, the ambient atmosphere. Also recorded in Table 24 are the
safe dilution volumes (SDV) of the same solvents. The SDV is the
amount of fresh air measured in cubic metres needed to dilute one litre
of solvent to its TLV. From the SDV it is easy to calculate the
ventilation requirements in the number of air-changes per hour. A
SATRA publication (Huggett, 1971), from which the TLV and SDV
figures have been taken, gives the formula:

Number of air changes per hour = (SDV) I S/8R

where 1= number of litres used in an 8 h day, R = cubic capacity of


room, (m3), S is a safety factor. The safety factor recommended in the
SATRA publication is 2 for a work-place smaller than 150 m3; 3 for a
cubic capacity of 150-600 m3 and 4 for anywhere greater than 600 m3.
The normal planning figures for air changes per hour used by architects
and heating/ventilating engineers is 4 and, if the above formula gives a
figure larger than 4, then solvent extraction at source is essential.
It is obvious from Table 24 that the preferred use of tri-
chloroethylene should refer to its use in fully enclosed systems such as
those used for vapour degreasing and where, on a small scale opera-
tion, solvent is used on the factory bench, it should be replaced by
1,1,1-trichloroethane. The equivalence of the cleaning efficiency of
these two solvents is indicated by the closeness of their solubility
parameters of 18·7 and 19·8 (MJ.m-3)~ but 1,1,1-trichloroethane is by
at least 50% the more expensive of the two materials. Similarly,
methanol which is recommended (Shields, 1976) für polycarbonate,
polystyrene and polysulphone plastics is better replaced by iso-
propanol. For cleaning GRP and CFRP, methyl ethyl ketone should be
replaced by acetone which also has the advantage of being cheaper.
All solvents which are efficient for cleaning surfaces will also remove
the natural body greases from the hands. Whilst direct handling and
wiping of components with solvent soaked cloths is to be deprecated,
there are occasions when it will be done. Because barrier creams will
themselves contaminate the surfaces to be cleaned, the hands are best
protected with cheap disposable polyethylene gloves.
Because of their low thermal capacity, small metallic parts are more
SURFACE PREPARATlON 223

efficiently c1eaned by condensing vapour on to them in a chemical


extraction apparatus rather than hanging them in the vapour as is the
normal procedure for vapour degreasing.

CHEMICAL ETCHING

Aluminium
For metal-to-metal joints of the highest strength and greatest durabil-
ity, a rigorous treatment with chemical etchants is essential. There is a
very wide range of treatments recommended for the different metals,
and their variety in part reflects the behaviour of different alloys of the
same basic metal and, in part, increasingly rigorous environmental
requirements and test procedures. Thus, the gradual replacement of
the chromic-sulphuric acid for aluminium by anodizing treatments is
consequent upon the discovery that the nature and thickness of the
oxide surface with the chromic-sulphuric acid process is more depen-
dent on the washing which follows than on the etch itself; that under
conditions of high humidity with water vapour reaching the interface,
the oxide layer thickens with new oxide which is less coherent than the
originallayer and, for reasons which are not c1ear, these faults became
apparent with the newer toughened epoxy adhesives curing at 120°C
although the chrome-sulphuric treatment was and still is, very effective
with nitrile-phenolic polyvinyl formal/PF and epoxy-novolac adhesives
cured at 17 S°c. A summarized discussion of these facts, together with
an indication of the evidence is given by Cotter (1977). The choice of
treatments for aluminium and its alloys revolves round the scale of
operations, the environmental durability which is demanded and the
adhesive to be used; hence there is a need for multiple recommenda-
tions for these materials. The use of alloys c1ad with pure aluminium or
of unc1ad alloy need not influence the treatment procedure to be
adopted although, in fact, the US manufacturers avoid c1ad alloys and
have now switched to a phosphoric acid anodizing process (see
ASTM D 3933-80) whereas European manufacturers use c1ad alloys
and a chromic acid anodizing process.
With this rather lengthy preamble, Tables 25 and 26 indicate satis-
factory treatments für aluminium given certain fixed points of refer-
ence, thus obviating the need for considering alternatives among a
number of apparently equal possibilities.
224 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

TABLE 25
CRITERIA FOR SURFACE TREATMENT OF ALUMINIUM

Conditions of Type of adhesive Availability Etch treatment


service to be used of electrolysis recommended in
tank Table 26

Phenolic-containing {Yes Aor C


adhesive cured at No B
Hot, wet high temperatures
environment
Modified epoxies {Yes C
cured at 120°C No Adhesive not
recommended

Phenolic-containing {Yes Aor C


Outdoors in adhesive No B
temperate c1imate
Modified epoxies {Yes C
cured at 120°C No B

Less severe than Any adhesive {Yes A or C


above c1asses No BorD

Notes on Table 26 (pp. 226-7)


In view of the multiplicity of recommendations, the following points of
explanation are necessary to expand or justify the selections made.
Treatment A
(i) This is a UK Ministry of Defence specification DTD 910C. The
specification further states that the ion concentration of the
anodizing solution must not exceed the limits specified and that
free chromium trioxide should not exceed 30-36 g.litre- 1 •
Chromium trioxide dissolves in water to give chromic acid, but
the equilibrium:
CrÜ3 + HzO ~ H ZCrü4
is easily pushed to the left, particularly in the presence of
sulphuric acid or if the solution is heated. The chromium
trioxide should be dissolved in the bulk of the water, retaining
two small portions, say 100 rnl each, for the sodium chloride
and the sulphuric acid both of which should be added as
solutions when the chromium trioxide is completely dissolved.
Should chromium trioxide be precipitated either on standing
SURFACE PREPARATION 225

or with use it may be ignored provided that it does not exceed


one third of the total present. This is a specification require-
ment and, if there is doubt about the amount, filtration through
a sintered glass crucible is possible on the small scale but
chemical analysis of the liquor may be required when a precipi-
tate accumulates in a large tank.
(ii) The bond strength obtained with some adhesives has been
shown to be sensitive to chloride ions present in ppm. By
adding 200 ppm (i.e. 0·2 g.litre- 1) any effect of impurity is
swamped and hence constancy of performance obtained.
Treatment B
(i) This is the UK Ministry of Defence Specification DTD 915B.
The chromium trioxide should be dissolved in half the water
and the sulphuric acid diluted into the other half. The diluted
acid should then be added to the solution of chromium triox-
ide. This procedure should avoid the precipitation of the oxide
or failure to dissolve it.
(ii) The washing procedure can be hastened, or made more
thorough by using warm water. If this is done, it is important
to maintain the temperature below 65°C or 60°C for safety.
(iii) Freshly made solutions do not give such stable coatings as a
well-used or 'ripened' tank. This effect has been traced to the
need for the presence of copper ions dissolved from aluminium
alloys.
Schwartz (1977) recommended leaving a small piece of
aluminium alloy in the bath for 24 h prior to use to 'seed' the
tank.
Treatment C
This phosphoric anodizing treatment is most favoured in the USo It has
an additional advantage on the large scale that the spent solution is
more easily disposed of than one containing chromium. However, the
absence of chromium at any stage of the treatment renders the use of a
chromate primer essential. The favoured material is strontium chro-
mate.
Treatment D
This is a rapid treatment given by Martin (1967). It should be perfectly
satisfactory but does not appear in any official specification possibly
because rapid treatments pose difficulties of control on a large scale.
~
0'1

TABLE 26
SURFACE TREATMENTS FOR ALUMINIUM
i
cn
~
Treatmentt Preparation Etch or anodizing solution Anodizing conditions Washing
~
Chromium trioxide 100 g Running tap water 15 (3
A Vapour degrease for Workpiece anode; inert
10 min. Immerse 5 Sulphuric acid 0·2 g* cathode. Solution at min. Rinse in de- ~
min at 40°C in: (specific gravity 1-84) 40°C. Raise voltage ionized or distilled
Trisodium phos- Sodium chloride 0·2 g to 40 V over 10 min. water. Dry in clean z
phate 25 g Water to 1litre Maintain for 20 min. air at 70°C. ~
TeepolS ml Raise to 50 V over 5
Water to 1litre min. Maintain for 5 ~
min.
~
o
B AsA Chromium trioxide 50 g
Sulphuric acid 80 ml (spe-
cific gravity 1-84) AsA
Water to 1litre
Immerse in solution at
60°C for 30 min.
c AsA Ortho-phosphoric acid 60 Solution at room
ml (specific gravity 1·7) temperature. Raise AsA
Water to 1litre voltage to 10 V over
2 min. Maintain for 5
min.

D Immerse for 30 s at room


temperature:
Sodium fluoride 10 g
Nitric acid 106 ml (specific
gravity 1·42) AsA
Water to 1litre i9
Then passivate for 1 min at
60°C in: ~
Q
Sulphuric acid 163 ml "C
(specific gravity 1· 84)
Sodium dichrornate 75 g
Water to 1litre ~
~
* 2 drops of concentrated sulphuric acid from a pipette will be sufficiently accurate, if litre quantities are mixed. ~
t See text for explanation of treatments.

~
228 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Ferrous Metals
Chemical treatment of the wide group of the ferrous alloys is not
straightforward because of the precipitation of free carbon on the
surface, known colloquially as 'smutting', and the consequent need for
'de-smutting'. Additionally, it is impossible thoroughly to clean and
etch mild steel and to finish with a washing process without corrosion
occurring. Although the stainless steels vary greatly in composition and
structure, there is little published information on the effect of composi-
tion on either adhesive bond strength or interaction with etchant
solutions. The chemistry of the processes seems to be as follows:

(i) Any strong reducing acid will serve as an etchant, but sulphuric
or hydrofluoric acids are preferable.
(ii) With stainless steels, there is an induction period before etch-
ing begins. This may vary with the thickness of the oxide layer
already present and with its nature which, in turn, depends on
the alloy composition and heat treatment. Whatever the cause
of variation, induction periods can be up to or even exceeding
an hour even at lOO°e.
(iii) Whilst mild steel is conveniently etched at 20°C, stainless steels
are likely to require temperatures in excess of 70°e. The time
of immersion and at what temperature is best determined by
experimental bonding and measurement of bond strengths.
With a specific maraging stainless steel, Firth-Vickers
FV 520B, Allen and Alsalim (1977) recommend 20 min at
lOO°C in either hydroftuoric acid as given below or in sulphurie
and oxalic acids.
(iv) The vigour of the reaetion, onee it has started is redueed if an
oxidizing agent is present and this seems without effect on the
duration of the induetion period.

Recommended Procedure
Any obvious rust or mill scale on mild steel should be removed by
brushing and all steels should be vapour degreased. Mild steel should
then be grit blasted. Whether or not stainless steel is grit blasted will
depend on the nature of the steel, the heat treatment it has reeeived
and the precision of machining. However, all stainless steels ean be
safely wet blasted with an aqueous suspension of alumina or gamet.
Etch in 4% solution of hydrofluoric acid at room temperature for
mild steel and at a suitably elevated temperature for stainless steel
SURFACE PREPARATION 229

such that the induction period is not too long nor the re action too
vigorous. Etch for a time measured from the beginning of the re action
for 10 min for mild steel and up to 20 min for stainless steels.
Immediately rinse in water and then transfer to a desmutting bath at
70°C composed of:
Sulphuric acid, specific gravity 1·84 57 ml.
Chromium trioxide 100 g
Water to 1litre
(See note on p. 224 on preparing chromic acid solutions).
It is important that desmutting is carried out immediately after
etching. Desmutting should occur very rapidly and certainly within a
very few minutes. The metal is then rinsed and washed in running
water and transferred to isopropanol and then in the case of mild steel
to a second bath of dry isopropanol before drying. If water is not
replaced by a dry solvent before drying, mild steel will rust in seconds.
Quick lime or a molecular sieve desieeant ean be added to isopropanol
to ensure dryness and filtered off before use. Store in a desiccator until
needed or apply a primer eoating at onee. (NB Primers are diseussed
later in the chapter.) Some authorities recommend vigorous brushing
to remove smut, but the ehemical method given above has been found
more effieient. If etching has been efficient, a microscopically rough
surface will exist, depending for its roughness on the phase structure of
the alloy. It is easy to brush earbon into microseopic valleys and
impossible to brush it from them.

Titanium
Titanium alloys are used for aireraft struetures and the metallographic
strueture of different alloys leads to different surface struetures when
etehed. Maximum bond strengths with a newly-introdueed alloy will
only be obtained as a result of experiments in whieh the eoneentration
of reagents and time of treatment are varied over a suitable range. In
order to obtain this desirable maximum strength and enduranee the
etching must leave a eoherent layer of titanium oxide (rutile) on the
surface and the treated alloy must have a rough surface. This is to be
achieved by using an etehant in whieh the rate of attack on the two or
more phases of the alloy differs greatly. Titanium alloys are subjeet to
hydrogen embrittlement and many reeommendations employ redueing
acids (e.g. hydrofluorie acid) together with an oxidant to rninimize the
uptake of hydrogen. These problems are avoided by the use of an
230 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

alkaline-peroxide etchant (Allen et al., 1974). Industrial evaluation


including exposure of unstressed lap-shear test pieces to high humidity
(Westland Helicopters Ltd, 1974, 1975) and stressed double lap-shear
at both temperate and tropical stations (Cotter, 1977) have both
indicated the value of this treatment in giving a surface leading to
durable bonds.
After vapour degreasing and wet blasting with alumina, immerse for
20 min at 65-70°C in:
Sodium hydroxide 20 g
Hydrogen peroxide 30% (100 vol) 22·5 m1
Water to 1litre
The alkali is dissolved in water and the temperature brought to 65°C
ready for immersion of the metal pieces. The hydrogen peroxide
solution is added with stirring immediately before the metallic compo-
nents. Remove the components when their surfaces have darkened to
approach black. This normally takes 10-20 min. Wash in hot water for
at least 10 min, dry in warm air and, preferably, apply primer coat
immediately.
The phase structure of the 90Ti-6AI-4V alloy at present in use is
inftuenced by the heat treatment the alloy has received. This in turn
influences the surface structure obtained with any etchant. It will also
affect the time required for the alkaline-peroxide etch to produce the
desirable rough coating of black rutile. If heat treatment has not been
such as to produce a uniform surface structure, surface treatment will
not produce uniform bond strength or durability.

Other Metals
Various etchants have been recommended for the remaining metals of
engineering practice, but it is doubtful if sufficient work has been
reported to differentiate between them or to assess their effect on the
durability of the bonds formed with different adhesives. Strong, dura-
ble bonds are uncertain with copper because of the ease with which a
weak, friable oxide is formed. Even when coated with an adhesive,
oxygen can diffuse to the interface and eventually cause failure. Brass
has an oxide film almost entirely of zinc oxide and, as with zinc
'galvanized' iron, it can hydrate or form salts with the tackifiers added
to some contact adhesives. Cadmium is met with as a plating; if a
strong, durable adhesive bond is essential, it should be replaced by
chromium, the surface of which can be treated as stainless steel.
SURFACE PREPARATION 231

All the above met als are best treated for adequate bond strength by
a sequence which involves degreasing followed by some abrasive
treatment which may vary according to circumstance from emery paper
through to wet blasting with alumina grit. After this roughening-
cleaning it is vapour degreased and then dipped for a full 5 min in:

Trisodium orthophosphate 25 g
Teepol 5 m1
Water to 1litre

After this, it is thoroughly rinsed in warm water, acetone and dried in


warm air. It is doubtful if chemical etching is really worthwhile for
metals other than the steels, aluminium and titanium.
For those who wish to apply a full etch procedure to zinc coated or
cadmium plated metal, McIntyre (1972) gives appendices with very full
specification details in his paper on bonding to cadmium and zinc.

PRIMING LAYERS

At the beginning of this chapter, six functions were listed as fulfilled by


priming coats. Two important ones of these are discussed below in a
section on coupling agents, the remaining four need only slight amplifi-
cation of the subject matter of the list.
Since the purpose of surface treatment is primarily to get a surface
free of contaminants which would act as weak boundary layers be-
tween adherend and adhesive, item (i) of the list is an obvious action to
take immediately cleaning and surface preparation has been com-
pleted. The adhesive manufacturer often supplies a suitable material in
dilute solution in a solvent which can be conveniently and rapidly
applied by spraying. The solution can be simply the adhesive to be
used but dissolved in solvent instead of supplied as a viscous liquid,
paste or film. If it is heat-curing, care must be taken not to allow curing
to occur when drying the primer, i.e. removing the solvent from it. It
should be allowed to dry in a current of warm air and not stoved. For
room temperature curing adhesives, usually two-part formulations, the
primer may consist of the resin without curing agent or without
accelerator, reliance being placed on diffusion of adequate reagent
from the bulk adhesive when it is applied. The use of solvent over-
comes any difficulty in the rate of wetting if the bulk adhesive is solid
232 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

and, when heated becomes a viscous liquid already starting to cross-


link and hence to become even more viscous (item (ii)).
Of less importance for structural adhesives is item (iii) except where
concrete is one of the adherends. Even in this case, however, penetra-
tion of the adhesive into the substrate may be a very desirable and, as
epoxy adhesives will be those normally involved, viscosity can be
adjusted to avoid any possibility of a starved joint.
Item (iv) concerns corrosion inhibition. The reagents most com-
monly employed are strontium or zinc chromates. As only those
particles in direct contact with the steel or aluminium surface are
effective inhibitors, it is reasonable to apply them in a suitable binder
in as thin a layer as possible. Etch primers, such as are used in paint
technology, are unsuitable for use with adhesives and the etching must
be carried out as a separate operation as has been discussed in detail
above. Where a chromic acid process has been used to prepare the
surface, some chromate becomes incorporated therein and the use of a
chromate-containing primer seems superfluous.

Primers as CoupUng Agents


The last two items of the earlier list involve the behaviour of primers,
in part chemical, in part physical, as forms of adhesives sticking the
bulk adhesive to the actual metal, wood or other adherend. When
dissimilar materials are to be joined, an adhesive suitable for one may
not be so for the other. In these circumstances a primer for one of the
substrates is chosen to be both suitable for it and compatible with the
adhesive. Phenol-formaldehyde condensation products are often thus
used particularly when the adhesive itself contains such material dis-
persed in another polymer, e.g. Redux 775 (Bonded Structures, Ltd) or
various epoxy-phenolic or nitrile-phenolic adhesives. Most important
members of the dass of coupling agents are the reactive silanes
(siloxanes) and titanates. These materials have achieved their main use
in the treatment of glass fibres for the manufacture of glass reinforced
plastics.
Adhesives which incorporate phenol-formaldehyde res ins are much
used as structural adhesives for metals. They comprise the original
polyvinyl formaljPF, and polyvinyl butyral/PF together with nitrile
phenolics and epoxy phenolics. The ratio of nod-phenolic polymer to
the PF resin in the structural adhesives determines the morphology of
the material, whether it is two-phased, and which phase is continuous
and which discrete. This, in turn, influences the flexibility of the
SURFACE PREPARATION 233

adhesive and hence the ratio of peel strength to shear strength. Where
high peel strength is desirable, the composition may be such that
adequate wetting of the substrate by the PF component does not occur.
It is the PF component, by virtue of its capacity to form hydrogen
bonds, that unites with the (hydrated) oxide of the metal surface or the
cellulose hydroxyl bonds of wood. Its reaction is ensured if it is
employed in an alcoholic solution and without the presence of the
other polymeric component.
Siloxane coupling agents are applied to the metal or glass surface as
a monomer. A condensation reaction occurs as it is dried on and
baked, giving a very thin resinous layer whieh is attached through
primary valency groups to the oxide of the metal or to the structure of
the glass. The high energy oxide surface is now hidden and replaced by
a surface of relatively low energy. For water to displace this resinous
coating of siloxane from the metal or glass surface, it is necessary for
hydrolysis of a silicon-oxygen valency linkage to occur. This is a
relatively slow process and requires excess water to be present at the
molecular site involved. However, because this bond exists at the
interface, it is energetieally difficult for water. to get to the interface,
even though the siloxane resin itself will allow the passage of water.
Hence, the bond, although theoretically capable of breakage by hyd-
rolysis is, in fact, very stable under external conditions of high humid-
ity. The surface of the siloxane resin, being of low surface energy, is
hydrophobie rather than hydrophilic so that an adhesive applied to it is
energetically not very likely to be displaced by water even when there
is no chemical reaction between the siloxane and the adhesive. To take
full advantage of the stability of the siloxane-metal oxide surface, it is
desirable to build in an equally stable chemical connection between the
adhesive and the siloxane. It is because of the existence of a truly
chemical connection as distinct from physical (or van der Waals's)
adsorption between oxide surface and siloxane on the one hand and
siloxane and adhesive on the other that these reagents are known as
coupling agents. The chemical bond to be inserted between the silox-
ane and the adhesive depends on inserting into the siloxane monomer
molecules a chemical group which is functionally reactive towards
certain chemical features of the adhesive molecule. Thus, epoxy adhe-
sives frequently depend on polyfunctional amines for their cross-
linking reaction. It is therefore possible to use either epoxy groups or
amine groups on the siloxane, for in one case they will react with an
amine on the hardener molecule leaving the other amine groups to
234 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

react with the epoxy adhesive while in the other case, direct reaction
with some of the epoxy groups of the adhesive will occur. Table 27 lists
the silane derivatives that are available for siloxane coupling agents
together with the type of adhesive for which they are suited.
Although siloxanes have been available for a number of years and
have given excellent performance when used as coupling agents in
GRP, their use in adhesives technology has not become as widespread
as was expected. One of the reasons for this is the excellent perfor-
mance of many adhesives in conditions of high humidity without the
use of a siloxane coupling agent. If good stability is obtained in hot,
wet tropical conditions as Cotter (1977) shows to be true with, for
example, nitrile-phenolic or a modified epoxy, there is no point in
adding to the complexity and cost of the bonding operation. On the
other hand, siloxane coupling agents do not appear to have been used
with polyvinyl formal/PF adhesives which Cotter reports as showing
properties inferior to epoxy-novolac, modified epoxy, epoxy phenolic
and nitrile phenolic adhesives under stress at a hot, wet tropical
exposure site. In general, adhesives incorporating phenolic res ins or
the corresponding functional groupings do not need silane primers
because they would not profit from them. The exception to this among
the adhesives tested in the UK Ministry of Defence Procurement
Executive exposure trials reported by Cotter is the vinyl phenolic (i.e.
polyvinyl formaljPF resin adhesive). It is likely that the actual adhesive
used in this case is relatively low in PF content. Hockney (1972),
referring to the same series of trials reports a high peel strength for it.
A higher content of PF, though reducing peel strength, would increase
durability. Cotter's review does show the poor performance of an
epoxy-nylon in a hot, wet environment. This structural adhesive un-
doubtedly would be improved by the use of a siloxane primer. Indeed,
Schrader and Cardamone (1978) have demonstrated this in the
laboratory. Tbey used I'-aminopropyltriethoxy silane as the coupling
agent on titanium adherends which were made with an epoxy-nylon
adhesive into lap-shear joint test pieces. Such joints were immersed in
boiling water for 24 h. Tbe immersed joints were tested wet at the
temperature to which they had fallen on removal from boiling water to
the testing machine in the open air. Tbe presence of the coupling agent
improved the strength retention of the joint from 54%-63%. Simi-
larly, an epoxy-nitrile adhesive joint retained 68% and 80% in the
absence and presence, respectively, of the same coupling agent.
Kinloch et al. (1976) also used the fall in strength of a bonded joint
TABLE 27
SILANE COUPLING AGENTS FOR USE WITH ADHESIVES (CASSIDY AND Y AGER, 1972)

Silane Trade designation May be used with V>


and manufacturer e
:;0
'Tl
;J>
l' - Methacryloxypropyltrimethoxy silane Union Carbide Unsaturated polyesters ()
{A174 m
Z6030 Dow Corning '"C
1'-Glycidoxypropyltrimethoxy silane Union Carbide Epoxies, two-part :;0
{A187 m
A6040 Dow Corning polyurethanes or other '"C
;J>
isocyanate-curing adhesives :;0
l' - Aminopropyltriethoxy silane A1100 Union Carbide Epoxies, isocyanate- ~
curing adhesives 0
Z
N-ß-(Aminoethyl) ')'-aminopropyltrimethoxy silane {A 1120 Union Carbide As above
Z6020 Dow Corning
l' - Mercaptopropyltrimethoxy silane A189 Union Carbide Nitrile-phenolics, epoxies

N
V-l
VI
236 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

in boiling water as an indicator of the efficiency of a coupling agent,


the ')I-glycidoxy compound instead of the ')I-amino, and the adherends
were gl ass instead of metal. They report an optimum system in which a
simple polyamide cured bis-phenol A resin was replaced by the use of
the coupling agent and a spiroacetal amine curing agent. In this case, a
very large glass structure which, from its nature involved room temper-
ature curing, had started to fall after only months of exposure in a
temperate c1imate. Laboratory experiment confirmed the incipient
instability of the system without a coupling agent but also gave a
confident basis for the prediction of decades of life when the siloxane
coupled system was used.
The indications are therefore that if an adhesive known to be
humidity sensitive has for other reasons to be used under conditions of
high humidity, there will be a gain in using an appropriate siloxane
coupling agent. This judgment can only be expected to be valid when
the coupling agent has a functional group on it capable of reacting with
the adhesive and the substrate possesses an oxide surface. The use of a
coupling agent will not, of itself, convert an unsuitable highly moisture
sensitive adhesive into an environmentally stable system.
Chapter 7

Service Life

The estimated service life of a joint made with structural adhesives


necessarily depends on controlled laboratory or environmentally ex-
posed trial joints for which the load applied is precisely known. In
service, the applied load will in general be much lower than that
applied in testing. The load used during testing may therefore be
regarded as applying an accelerated test whereby the time-scale is a
fraction of the service life expected or the increased load a safety
factor incorporated into the estimated life. In the former case it is
desirable to establish the relation between test life and service life; i.e.
the degree of acceleration given by the test procedure. This is not
easily done with confidence. Frequently, it is attempted by establishing
the actual relation in the short-term and then assuming the relation to
hold over very much longer periods of time. A few dead-Ioad experi-
ments, to be discussed below, have been run for up to 6 years and
cyclical loading has been applied for numbers of loading cycles equi-
valent to those expected over the planned service life. The latter
instance assumes that the performance of the joint when cyclical
loading is intermittent and is interspersed with static loading applied
between periods of cyclical loading, is at least as good as when the
dynamic loading is continuous.
An important concept, familiar to engineers studying fatigue in
metals, is that of the endurance limit. This refers usually to the peak
cyclicalload wh ich the joint can sustain indefinitely. It is also used with
respect to static loading. It is not, however, a stress which can be
identified with certainty and will be influenced by the type of failure to
be expected if it is exceeded. Where attempts have been made posi-
tively to identify this stress by short-term experiments the results
237
238 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

obtained, ab out 40% of the ultimate tensile strength (UTS), seem very
high (Lewis et al., 1972). Such figures are not supported by long-term
trials and could only be employed where a service life of comparative
brevity is required or where the times of application of stress are brief
compared with the service life of the structure. As will be recounted in
the following paragraphs, the response of an adhesively bonded joint
to stresses which cause creep is particularly difficult to fit with the
concept of an endurance limit. It is, therefore, probably best reserved
for describing response to dynamic stressing.

THE CREEP OF ADHESIVE JOINfS

In normal design under static loading a continuous extension or shear


in the direction of the principalload is not expected whilst the adhesive
is maintained below its glass transition temperature, Tg • Overloading is
far more likely to lead to stress rupture than to creep. However,
components at elevated temperatures could be maintaining the adhe-
sive elose to or at Tg , under which conditions some creep might occur.
A complex structure cannot be stress-free even in the absence of
external loading. Manufacturing imprecisions and minor stresses not
allowed for in the design may give a structure which has initially some
peak stresses. These stresses redistribute themselves with time by a
creep mechanism. Where adhesive joints form part of the structure it is
reasonable to assurne that they will participate in this stress redistribu-
tion.
Allen and Shanahan (1975,1976) studied the tensile creep of lap-
shear joints at temperatures in the neighbourhood of Tg and below.
They found that actual creep under load was preceded by a delay or
induction period which was temperature and load dependent. Follow-
ing this delay period, steady state creep took place which was logarith-
mic with respect to time. Finally, this steady creep gave way to an
accelerated creep terminating in stress rupture. No detectable move-
ment occurred during the delay period. Evidence that has accumulated
since this work was done suggests that the stress distribution in the
lap-shear joint considerably inftuences whether or not tensile creep
occurs. Allan and Shanahan worked with steel adherends joined by
Redux 775 (Ciba-Geigy) as the adhesive. Attempts to reproduce the
results with aluminium adherends of similar size foundered, the joints
breaking without showing previously any tensile creep. The explana-
SERVICE LIFE 239

tion appears to be that with the steel adherends the ratio of the
maximum shear stress to the maximum c1eavage stress allows tensile
shear but that the lower modulus of aluminium imposes a relatively
larger c1eavage stress from which failure propagates. The relevant
stress concentrations are, of course, those at the ends of the overlap of
a lap-shear joint. In a structure the c1eavage or peeling stress may be
minimized by chamfering the overlapping adherend or otherwise con-
touring the shape to fit the stress. Moreover, a joint in a structure
should not be designed to sustain a load which, though nominally in
shear, gives rise to a c1eavage stress sufficient to cause rupture. In
aircraft fuselages where stringers and stiffeners are bonded, their ends
are the sites of definite c1eavage stresses and may be, by some
manufacturers, reinforced against c1eavage-induced failure by the in-
sertion of a single rivet. This should not be necessary in properly
designed and manufactured structures.
It is difficult to anticipate conditions under which creep may be a
problem but, obviously, they will be those where shear predominates
in the stress concentration at the end of the overlap. Before tensile
creep commences, the delay period occurs and this may extend far into
the service life of the structure. There appears to be no information on
this but the nature of the delay or induction period suggests that other
changes in the adhesive due to temperature and humidity will, in fact,
hasten the stress redistribution which is believed to accompany the
onset, if not the rate of creep. The stresses at the ends of an overlap
joint are representedl in Figs 9, 28 and 29. By the time creep is
observable, these end stress concentrations, although they still exist,
will be less marked whilst the stress in the central section will be raised
nearer to the mean stress. If, as Hahn (1961) believed, creep could
only start when the peak stresses have been modified, then the
adsorption of moisture, plasticizing the resin of the adhesive (Iowering
its T J and lowering its modulus will assist redistribution of peak stress
and will shorten the delay period before creep commences. Allen and
Shanahan, however, think that primary chemical bonds are severed
during the delay so that some critical loosening of the network struc-
ture is involved before movement can occur. If this is true then
adsorption of moisture by plasticizing the adhesive resin will minimize
the effect of stress on individual chemical linkage and hence certainly
not shorten delay time though it will still encourage creep once the
delay period is passed. The original query posed as to whether the
delay period extends appreciably into the service Iife is, therefore,
240 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Failure by
stress rupture
I I
elay time, td ~Period of logarithmic creep
I
-o!

slope =Y/ln(t)

Log (time)

FIG. 107. Schematic representation of creep versus time (after Allen and
Shanahan,1976).

unanswerable at present. If the stress is of the nature and magnitude to


cause creep, then it must be assumed to start at the time of first loading
and allowed for accordingly.
Figure 107 shows the nature of the creep relation to time. If 'Y is the
percentage shear creep strain, d'Y/dt is the creep rate. But, because of
the logarithmic nature of the relation between rate and time, the creep
is characterized by its rate at time (to) where the creep curve intercepts
the log (time) axis. The shear rate at time, to, i.e. y(to), is a linear
function of the load applied to the joint provided that load is above a
certain minimum required to induce creep. The magnitude of this
minimum with reference to the short-term breaking stress has not been
investigated.
Given that the high stress is to be applied and creep expected, the
results reported by Allen and Shanahan suggest that under given
conditions but with varying load, the creep becomes catastrophic and
runs into stress rupture after the percentage shear strain reaches a
certain value substantially independent both of time and load. This can
be seen from Fig~ 108.
If creep is likely then increasing the length of overlap is advantage-
ous. Long overlap shear joints are normally avoided because increasing
the overlap does not proportionally increase the short-term strength
but the increased length does both delay the onset of creep and reduce
its rate.
It will be appreciated that under conditions of static loading an
SERVICE LIFE 241

271

325 KN

~ 50
~
Q) 2-89
"-
c: f
"eu; 2-53 2-35
f f

199
181d d

217 d

Time (min)

FIG.108. Delay times and creep of 1 cm overlap-shear joints at 75°C with


epoxy-novolac adhesive (Bonded Structures Ud); d = discontinued, f = failed
(after Allen and Shanahan, 1976).

endurance limit may refer to the load below which response is only the
instantaneous elastic response, or to the load for which the induction
period for creep is greater than the foreseeable employment of the
structure, or to the load which, although causing creep extension does
not lead during the life of the structure to stress rupture, it being
assumed that the creep deformation does not impair its functioning. It
is this ambiguity which limits the use of the term 'endurance limit'.

TIME-TO-FAILURE (UNDER STATIe LOADING)

For short-term operation, for example in rocketry, the time-to-failure


may be an important parameter enabling stressing to be high and the
use of the adhesive efficient because a limited working life is designed
into the structure. In general, however, the time-to-failure is much less
important for it is the rate at which the strength of the joint declines
which determines the time at which the known stress will be insuffi-
ciently smaller than the decayed strength to be considered safe. How-
ever, because time-to-failure is adefinite and accurately determined
figure, it possesses attractions in attempts to predict probable service
life. Theoretical foundations have sought to explain the dependence of
242 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

time-to-failure on load and temperature, predicting with some success


the recorded failure pattern, but over a very small time scale up to no
more than 6 min (McAbee et al., 1970). For this to be useful in other
than rocketry or missiles, it is essential to have a confidence in
extrapolation to periods measured by years rather than minutes.
The equation developed by McAbee et al. is a logarithmic one:
log (tfT 2 jS) = C+ AH*j2'3RT- bSjT
in which AH* is the heat of activation of the process postulated as
behind the phenomena, T is the absolute temperature, S is the applied
stress, tf is the time-to-failure b, C are constants and R is the gas
constant.
At constant temperature this equation becomes
log (tf/S) = A' - B' S
where A' and B' indicate new constants.
With the very limited data available in Hockney's reports of long
term trials carried out by the Ministry of Defence in Australia (Hock-
ney, 1973; Cotter, 1977) it can be shown that the failure pattern of a
nylon epoxy adhesive in double-Iap shear joints in a hot, wet c1imate is
not inconsistent with this equation, which in the absence of further
information, could be used in limited extrapolation.
Figure 109 uses the median time-to-failure to illustrate the effect of
stress but it should be realized that the distribution of failure times is

1·5

0·5

o 10 20
Percent stress
FIG. 109. Equation of McAbee et al. (1970) illustrated by failure times of an
epoxy polyamide adhesive (Courtesy, Ministry ofDefence, RAE, Farnborough).
SERVICE LIFE 243

always very wide. Tbe median time at a stress which was 20% of the
initial ultimate tensile strength was 72 weeks but the actual failures
started after 12 weeks and continued beyond 90 weeks. * Properly to
predict the failure pattern, the distribution of failures under constant
conditions must also be known. It should then be possible from
relatively short-term measurements to predict the percentage failures
to be expected before any given period of elapsed time and how this is
likely to be affected by stress or temperature.

CYCLES-TO-FAILURE

Rather more information exists on the number of cycles of an alternat-


ing stress which is required to produce failure, thanks to the work of
Matting and Draugelates (1968). Using sinusoidal stressing of a single
overlap shear test piece, 60 mm wide with an overlap of 15 mm, a wide
range of variables was studied by these authors. Figure 110 is typically
selected from these to illustrate the properties of a joint bonded with
Redux 775 alternated between a lower stress level (average shear
stress on overlap) of 1·08 MN.m- 2 and an upper level indicated in the

~ 15·0 ,
'l' ......
7' ,
,
E 12.5
z "- ,
:I , ...:z....11" ....
"-
.10·0 --~120 ..
)(
IV ... >==-< 6
E ~"125
... t---I
oll 7·5 '~4',
oll
-00 5 2
co [I -'>
<-
ti 5·0
<-
IV
~ 2·5
In 1.1 (Shear stress minimum)
o I
104
Number 01 cycles

Fro. 110. Results of dynamic cyc1ing tests on aluminium sheet, lap shear joints
with Redux 775. Range shown by stated numbers of tests r---!':----i; tests in which
metal adherends failed 0; tests discontinued at indicated number of cyc1es 0
(adapted from Matting and Draugelates, 1968).

* We are indebted to the Ministry of Defence Procurement Executive (RAE,


Farnborough) for supplying the data on which Fig. 109 is based.
244 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

diagram. The test frequency was in the region 33·3-38·3 Hz depen-


dent, due to the method of exciting the vibration, on the stress level.
The duration of the longest test, 4· 3 x 108 cycles, was 3! months. The
authors recommend that a limit can conveniently be set at N =
1·2 X 107 cycles allowing, on the strength of the evidence they present,
extrapolation to 5 x 108 cycles. The band of results in Fig. 110 is
exactly analogous to a Wähler line familiar in metal fatigue testing.
The curve seems to level off at 5·6 MN.m- 2 below which loading no
test piece failed. This may be regarded as an endurance limit. At
oscillating loads with the maximum stress below the endurance limit,
the joints will, of course, eventually fail, but failure will not be related
to the stress level and is also likely to be initiated by adherend (metal)
failure. This limit, 5·6 MN.m- 2 may be regarded as an important
design parameter although its magnitude is, of course, tied to a
particular test piece and adhesive thickness as weIl as to characteris-
tics of the testing apparatus. 5·6 MN.m- 2 might correspond to between
10 and 15% of the strength which the joint would sustain in the usual
tensile test before cycling. This figure can usefully be compared with
continuous static loading on an adhesive of the same type (it may
indeed be the same product) exposed to a temperate climate. A
few failures only seem to have occurred and these aIl loaded at 20%
of the UTS. Tbe majority of these survived 6 years at 20% stress and a11
exposed, stressed joints survived 6 years at 10% UTS (Hockney, 1973;
Cotter, 1977).
Matting and Draugelates use for the actual Wähler curve, a line
drawn through the mean value of the groups of replicated test pieces
and point out that it becomes a straight line on a double logarithmic
plot expressed by the relation:
10gN=log C- Klog S
where S is the maximum of the stress cycle, N is the number of cycles,
K is a constant derived from the slope of the straight line. This
expression can be rearranged as N = CS- K and by reading the N value
for S = 1·0 the equation for the Wähler line is easily obtained and is
valid as far as is indicated by the linearity of the double logarithmic
plot, i.e. as far as the endurance limit which, as stated, was found to be
5·6 MN.m- 2 for Redux 775. A modified epoxy, Araldite 106, gave a
Wähler line valid up to 8 X 108 cycles with an endurance limit of
6·0 MN.m- 2 . For FM 1000, an epoxy nylon (Cyanamid), the endurance
limit may be estimated as much higher, at 12 MN.m- 2 , but only a value
SERVICE LlFE 245

of 7 MN.m- 2 can be used for design purposes because the extrapolated


number of cycles falls within the fatigue limit of the metal adherend
used in the experiment.

Inßuence of Temperature
As the ambient temperature is raised towards the Tg of the adhesive,
the Wöhler line is transposed downwards on the stress scale. With
i

Araldite 106, the change is marked even over the range 20°-40°
whereas Redux 77 5 has to be raised to 60°C before the change from
20°C is appreciable.

Inßuence of Test Frequency


lust as metal durability is greater at higher than at lower frequencies,
so also is the durability of adhesively bonded metals. If plotted as a
Wöhler line, that for Araldite 106 is raised by about 10% as the
frequency of test is increased from 12·5-36·7 Hz. This implies that
time-to-failure may be greatly increased depending on the slope of the
line, i.e. the value of K in the expression N = CS- K • With Araldite 106
it amounts to about a four fold increase. It is not immediately obvious
why, at the same maximum load, the higher frequency should be so
much less damaging and this must be discussed. Matting and Ulmer
(1963) calculated the stresses at the end of the overlap but came to the
conclusion that purely elastic calculations based on one-dimensional
stress distribution overstated the normal stresses with respect to the
shear stress at the end where, in any case, their method gave a stress
discontinuity (i.e. it placed the maximum at the overlap end instead of
just inside the terminus as do the methods outlined in Chapter 2).
They therefore developed a stroboscopic optical method of strain
measurement and showed how the displacement at the end of the
overlap varied as the test proceeded and altered with test frequency.
Figure 111 shows how the centres and ends of the overlaps move
during each cycle. The movement at the centre in response to the
imposed shear is always greater than at the ends. The two adhesives
used in this investigation behaved very differently. Araldite 106,
nominally the less rigid material (shear modulus 70 kgf mm- 2 ) showed
greater end movement at the lower frequency whereas Redux 775
(shear modulus 200 kgf mm- 2 ) showed greater end movement at
the higher frequency. In fact, at lower temperatures which, on the
temperature-frequency equivalence principle correspond to the higher
246 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

20 R~dux 775 , Mov~m~nt of c~ntr~ __ - - - -


I' _---------f---yOf OV~rlap~_~ - - - -
E - '/(Q'217HZ) _----~l
:1
10
---------1-------- I
:
I
I
: Mov~m~nt du ring ~ach
I at 38'3 Hz
cycJ~

o 5 15 18x104
cycl~s

Araldlt~ 106
30
MovlZm~nt during ~ach
cycJ~ at 0·217 Hz
20
E
:1 Mov~m~nt during lZach cycl~ at 38·3Hz
(mov~m~nt at c~ntr~ of oViZrlap similar)
10

o 5 10 14 X 10 4 cycl~s
FIG. 111. Movement of centres (broken lines) and ends (fulllines) of overlaps
during dynamic cycling of joints bonded with Redux 775 and Araldite 106
(adapted from Matting and Draugelates, 1968).

frequencies, the real part of the modulus measured by torsional


pendulum is very much the same for both materials. Reference to Figs
99-101 shows the relevant diagrams with plots corresponding to real
and imaginary part of the modulus against temperature. However, the
loss factor at ooe for Redux 77 5 is only ~ and at 200 e only 1/20th of
that for Araldite 106. At frequencies corresponding to these tempera-
tures, therefore, the strain transmitted will be lower with Araldite 106
due to the position and magnitude of the loss peak.
It is obviously of importance, therefore to consider the frequency of
test, or of use, in relation to the frequency at which the peak loss
occurs in the adhesive, i.e. the mean relaxation time of the adhesive
compound itself and the temperature at which the given frequency is
operative.

Inftuence 01 Amplitude
Variation in the amplitude of the applied shear is, of course, a
consequence of variation in the loading applied to the joint. The
response of the joint to increased loading at a fixed frequency would
SERVICE LIFE 247

be expected to be similar to its response to increased frequency. This


follows because if the overall deformation is increased in the same
cycle time, the rate of deformation must also have been increased.
Since the shear modulus of the adhesive is approximately linear for
working stresses, doubling the load will double the strain and hence
double the rate of strain. The response of the joint to this situation will
depend on the relation of frequency, implied by the rate of strain, to
the relaxation time of the adhesive. The decrease in time-to-failure
with increasing load is, of course, well attested with Wähler-type
curves and, allowing for the scatter of results, the relationship appears
to be linear with log (time) as long as the load is above the endurance
limit.
From the design viewpoint what is of interest is not necessarily the
expected life or time-to-failure, but the amount by whieh the strength
of a joint falls during a given number of cycles and what the effect on
the strength is of increasing the load. A limited amount of data exists
showing that as the load increases the strength remaining after 500
cycles decreases at first very slowly but after a certain load has been
exceeded then very rapidly. Figure 112 refers to experiments
with aluminium lap-shear joints ""hieh were cycled to a constant

~
+-'
Cl
c
~1·01---_
tl

In
.go.g
~
o
o
10
'-
c:I
+-'
moos
~

t»c
c:I
'-
tl
+-'
c
~oo7 ! • ! •

o 20 40 60 80
Maximum stress as percentage of UTS

FIG. 112. Effect of cyc1ic stressing on single-Iap shear joints; aluminium and
Redux 775.
248 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

~ ,
1\ ",
I \ "',
"
I \
,"
\ \
\ \
\ \
\ \
,
\

FIG. 113. Effect of cyc1ic stressing on lap shear joints; schematic.

maximum load in a machine with a constant rate of loading. The wave


form was sinusoidal. It is interesting to note that if the two arms of the
curve were approximated by straight lines they would intersect at a
load corresponding to about 40% of the ultimate tensile strength as
determined by the conventional 'static' tensile test. This is very much
of the same magnitude as the endurance limit quoted by Lewis et al.
(1972) for structural adhesives though they did not use Redux 775 in
their investigation. The data is even more limited on the eifect of the
number of cycles but the littIe there is suggests that the shape of the
curve is general. From the limited knowledge available Fig. 113 has
been constructed to indicate the probable response of joint strength to
changes in loading (amplitude) and time of cycling. Even less is known
about the response to loading which is varied during the period of
cycling. Matting and Draugelates depend on the Miner (1945) relation
which assumes that damage commences with the first cycle and pro-
ceeds linearly. If the loading is regarded as varying in randomly
imposed steps, each step, i, enduring for ~ cycles and having a loading
corresponding to the failure on the Wöhler line of Ni, the hypothesis
adopted states that faHure will occur when:
~~
L..-=1
iNi
Matting and Draugelates express doubts about the value of this
hypothesis and their own results based on two and four staged cycling
with both increasing and decreasing loading give 'total damage' values
SERVICE LIFE 249

lying between 0·2 and 2·2 with a majority between 0·7 and 1·4. These
authors quote other work (Späth, 1965) suggesting that the summation
leads to values in excess of unity when the load-steps are in a
decreasing sequence and less than unity when the loads rise with each
step. The results reported by Matting and Draugelates were too few
adequately to support or to refute this extension. The problem is
compounded by the fact that there is no true theoretical basis for the
Wähler line and these attempts at extending its use depend on the
wholly artificial hypothesis of linearly related 'darnage' . Work at The
City University has shown this hypothesis of linearly related 'darnage'
to be untenable. Against the body of information obtained by cyc1ing
at various load ampHtudes for 500 cyc1es and then measuring the
retained lap-shear strength, a number of joints were subjected to 5
cyc1es at 70% of their ultimate strength and 495 cyc1es at 35%. As will
be realized from Fig. 111 the loss of strength for 495 cyc1es at 35%
would be minimal. If the 5 cyc1es at 70% preceded the 495 at 35%, no
effect of stressing to 70% was noticeable whereas if they followed the
495 cyc1es the effect was almost as severe as if all 500 cyc1es had been
taken to the higher loading. Späth is quoted by Matting and Drauge-
lates as postulating an induction period before the onset of damage.
This presumably relates to crack initiation before crack growth pro-
ceeds as a steady process. It would appear that whereas crack initiation
can occur during 495 cyc1es it does not occur during 5 cyc1es even
when these are highly stressed cyc1es. However, having occurred, crack
growth at high stress levels is rapid. In service applications, therefore,
any component which has been subjected to even low amplitude
vibration is likely to be more susceptible to weakening when heavily
loaded than is a previously unloaded joint.

Inßuence of Moisture
Moisture has a profound effect on the service life of adhesive joints
made with some adhesives between some adherends. This point has
been made in the section on Water Sensitivity in Chapter 5. There are
four aspects of the problem which need to be drawn together when
looking at joints which have been exposed to moisture or are to be
designed to withstand such exposure. These comprise the ability of
moisture to diffuse through the adhesive, its susceptibility to displace-
ment from the substrate, the corrosion susceptibility of the substrate
and, lastly, the röle of any primer or sealing agent in modifying the first
three considerations.
250 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Diffusion and Absorption of Moisture


For moisture to affect an adhesive joint between two metal adherends it
must enter the joint by diffusion into the adhesive from an exposed
edge. It is often argued that it can 'wick' along the interface. This
assurnes either that the adhesive is displaced from the substrate at the
exposed edge, the rate of 'wicking' being the rate of displacement, or
that shrinkage of the adhesive away from the adherend occurs, as
could happen for example in a tubular lap joint without immediate
failure. The area avaiIable for entry of moisture may be very small in
absolute measurement as it is determined by the thickness of the
glue-line and the width or length of the overlap or butt, but it is very
large indeed compared with the entry for wicking confined to the
interface. The areas stand in a ratio of, say, 0·5 mm-30 nm if these
figures are taken as the glue-line and the interfacial interatomic or
molecular distance respectively. 'Wicking' as a primary mode of entry
is extremely improbable and entry is normally by diffusion through the
exposed boundary surface. The rate of transport of moisture through
the adhesive to the interface is given by the permeability which, in turn
is the product of the diffusion constant and the solubility. It follows
that high permeability can occur with an adhesive in which diffusion is
high or in which water is relatively soluble, or both. For example,
silicone sealants, which are often regarded as moisture resistant be-
cause the solubility of water in polysiloxanes is so extremely small, are,
in fact, very permeable to moisture because of a very high diffusion
constant.
The rate of diffusion of water in an adhesive is important if water
can displace the adhesive from its substrate, or if there is appreciable
solubility, for this determines its rate and pattern of saturation. Where
the adherend itself absorbs moisture, as does wood, the rate of diffusion
through the adhesive from one adherend to the other may be of
importance in minimizing stresses arising from incipient dimensional
changes. A full account of the method for calculating the pattern of
distribution of water in an adhesive joint is given by Althof (1981) and
also by Comyn (1981). Both these authors also consider the effect of
absorbed water and the pattern of its distribution on the properties of
the joint. The maximum stresses in a lap-shear joint are at the ends of
the overlaps and, of course, since here in one of the boundary surfaces
for the entry of moisture the moisture first attains moisture equilibrium
at the position of maximum stress. This results in a reduction of the
SERVICE LIFE 251

end stresses together with an increase in the strain due to changes in


the modulus of the adhesive. An iterative computing pro gram has been
developed and published (Althof, 1977) by the Deutsche Forschungs-
und Versuchsanstalt für Luft- und Raumfahrt (DFVLR) to give the
shear stress distribution in the adhesive with an inhomogeneous mois-
ture distribution enabling the long-term behaviour of joints to be
predicted in moist environments. The essential inputs for this pro gram
involve the mechanical elastic-plastic properties of the adhesive under
dry and wet conditions determined by experiments with bulk adhesive
material as well as with properties of lap-shear joints with thick
glue-lines.
The alternative approach involves the determination of the water
uptake by the adhesive, checking at the same time that the diffusion
leading to saturation obeys Fick's two laws, deducing the diffusion
constant from the rate of uptake of water and thence calculating the
distribution of water in the joint. These steps are outlined by Comyn
(1981) together with data on a number of epoxide-curative combina-
tions. The· epoxide resins were all based on the commonly used bis
phenol A with various amine curing agents.
Many structural adhesives are used as films carried by a fabric of
textile or glass. This, of course, provides further interfacial surfaces but
there is no evidence that wicking of the adhesive occurs either along
these interfaces or between the individual filaments which form the
yarn from which the cloth is woven. These facts follow from adhesion
between adhesive and the carrier filaments and complete penetration
of adhesive into the structure of the yarn under the pressure and
temperature of the cming process. Contrary to expectation the pres-
ence of the carrier usually interferes with the diffusion process, slowing
it down as moisture diffuses less readily through the polymer of the
yarn than through that of the adhesive.
As a guide to the entry of moisture into a typical lap-shear joint
Althof (1981) quotes an epoxy adhesive joint, 10 cm wide and 3 cm
overlap in which, after 60 days exposure to 95% relative humidity
(RH) at 70°C showed the equilibrium water content of 4% at the
boundaries with virtually zero in the centre of the overlap. Moloney et
al. (1981) performed a similar exercise with a lap-shear 2·5 cm wide
and 1·25 cm overlap bonded with FM 1000 and exposed to saturated
water vapour at 50°C for 1000 h (approx. 40 days). The overall uptake
of moisture of the joint was 69% of the true equilibrium value at
252 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

saturation but only 60% when a carrier was employed. The centre of
the overlap reached 20% of saturation but only 2% in the presence of
the carrier.
In normal service conditions, exposure to high humidity is intermit-
tent and periods of absorption are followed by periods of desorption of
moisture. Whilst only exceptional circumstances will lead to a complete
drying out of the joint, it is also true that only continuous immersion in
water will lead to complete saturation. Even under conditions of
tropical rain-belt climates intact joints will not be saturated
throughout.

Displacement of Adhesive
Apart from changing the physical properties of the adhesive resin, the
absorption of water and its transport to the interface with the adherend
can, in certain cases, lead to displacement of the adhesive by moisture
and hence failure of the joint. From what has beeil given above, it
follows that if displacement occurs it will do so from the edges of the
joint working inwards as diffusion proceeds. The conditions leading to
displacement or the possibility of displacement involve the nature of
the adhesive bond operating in given combinations of adhesive and
adherend. If this bond is purely physical, Le. it is a simple physical
adsorption, then the relative energies of adsorption of water to substrate
and adhesive compared with the energy of adsorption of adhesive to
substrate determine the equilibrium situation. This equilibrium will
tend to be shifted from the position where adhesive and substrate are
firmly bonded to the position where a film of water preferentially exists
at the interface (and hence the joint fails) if the adhesive absorbs more
than a few percent of water and the substrate is a high energy substrate
such as a metal oxide. If the bond between adhesive and adherend
involves chemisorption rather than physical adsorption, then displace-
ment can only occur after hydrolytic destruction of the chemical bond
uniting adhesive and substrate. Occasionally, separation of adhesive
from a primary coat occurs. The primary coat remains firmly bonded to
the adherend but moisture or the combined effect of moisture and
stress causes the adhesive to separate from the primer. This is not
common but has been known to occur in structural adhesives where a
phenol-formaldehyde resin has been used as a primary layer. The
cause may be associated with too great a degree of eure given to the
primer making it difficult for the main adhesive properly to interact
with it. Separation of paint layers from woodwork leaving the priming
SERVICE LIFE 253

layer intact is, of course, weIl known. This can usually be traced to
prolonged delay between priming and painting with consequent con-
taminations of the surface giving a weak boundary layer between coats.
When displacement or partial displacement of a structural adhesive
from its primer occurs, it is frequently a reversible phenomenon and
drying out frequently restores tensile strength. This reversibility par-
ticularly over many cycles, will lead to a gradual and permanent loss of
strength but the period of time over which this occurs will normally be
many years.
Displacement of adhesive from a metal substrate will rarely be
reversible. Examination of joints where this has occurred shows that in
the moist environment, displacement is followed by corrosion. Succes-
sively, therefore, a band of corrosion steps into the interface and the
area of bonding supporting the load diminishes until the bond can no
longer sustain the load. It is in such circumstances that the length of
overlap in a lap joint can determine the endurance rather than the
strength of abond. The increments in load bearing capacity obtained
by increasing the length of overlap diminish with increasing length and
there comes a point at which further increase is uneconomic in design.
However, although little or no strength increase is obtained there is a
gain in the life of the joint ascribable to the increased length.
Permanent immersion in water is not an acceptable environment for
organic adhesives and immersion in water at elevated temperatures,
although never envisaged as a service condition, has been used in
acce1erated tests. Figure 114 shows the effect of immersion in running
water at 70°C of three structural adhesives (Eichhorn, Hahn and
~30nr------r------r------,------.
N
'E
z __-. - __ - . 'r' - - amidcz-imidcz
~ =._-._.~- "
~=--

~20
.c
- - -=-- } czpoxy- phcznolic
:::::::::I~-=___==',_""",~'_=l'--="'-_
+'
t:n
C
~ • • • • - • -
"
. - . .::- -. - cyclo-aliphatic czpoxy
~10-~-----+------4=--~~--" r~'\'.\-----
1:

I----+----r-----+', , \ __ amidcz-imidcz
, \
0'+-::-____.1..:-____--'-;::--____-'-;---"__----:. cyclo-aliphatic czpoxy
10-2 10-1 10° 101 102
Pczriod of czxposurcz (wczczks)
FIG. 114. Strength properties of joints after exposure to various environments:
- . - . - . - running water at 20°C; - - - - - - running water at 70°C; - -
c1imatic cyc1ing to DIN 50016 (after Eichhorn et al., 1973).
254 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Fuchs, 1973), an imide-amide copolymer, an epoxy-phenolic and an


epoxy based on a cyclo-aliphatic compound cured with anhydride,
rather than the usual bis-phenol A cured with amine. The superiority
of the epoxy-phenolic in these particular circumstances is apparent; the
other two adhesives are obviously being displaced by the water. The
adherend was an acid-resisting stainless steel.

Effect of Moisture on Joint Strength


The effect of moisture on joint strength, and whether it is reversible or
permanent, will depend therefore on which of the above two mechan-
isms predominates. If the nature of the adhesive, the surface treatment
of the adherends and the resultant interaction between adhesive and
adherend are such that water does not displace the adhesive from the
adherend, the effect of water or high humidity on the strength of the
joint will be the result of absorbed water lowering the glass transition
temperature and the modulus of the adhesive. Whether the former or
the latter effect is paramount will depend on the range through which
the Tg is lowered in relation to the working temperaure. Obviously,
with an adhesive such as FM 1000 for which. Tg = 34°C, and which
is water-sensitive with a relatively high equilibrium uptake (c. 14%),
only a comparatively smalilowering of the Tg will be needed to bring
altered properties at temperatures likely to be ambient. Other adhe-
sives with Tg > 100°C would be expected to be less prone to loss of
strength in wet conditions. In practice, although immediate or even
rapid displacement of adhesive from the adherend surface is, for
appropriately prepared surfaces, rare, failure by separation does occur
on prolonged exposure. The time necessary for this failure mechanism
to become paramount is usually much greater than the time necessary
for the concentration of water in the adhesive at the interface to reach
its equilibrium value, an indication that some kinetic process, perhaps
an hydrolysis, occurs at the interface. This can be illustrated by
reference to the work of Butt and Cotter (1976) from whose paper Fig.
115 is adapted. (It has been redrawn to a log scale to facilitate
comparison with Fig. 114). The figure shows conclusively the inadequ-
acy of preparing aluminium surfaces merely by solvent degreasing and
the superior properties of a chromic-sulphuric acid etch. Since the
adhesive and its curing conditions are identical, the simple ingress of
water with a reduction in T g does not account for the fall in strength of
the solvent degreased joint. An additional indication of interfacial
displacement is that the nature of joint failure changed from wholly
SERVICE LIFE 255

N
I
20'·~·-~------~--------'----"

Chromic- sulphuric
/ etch

(:~
:2, 1 0 1 - - + - - - - - - - + Alkaline etch
'0
'0
t1I
.Q
Cl

~ 5~~-------+-------
t1I
~
aJ

10°
Period of exposure (weeks)

FIG. 115. Strength of lap shear joints after exposure to 97% relative humidity
at 43°C. Adhesive; epoxy-polyamide, FM 1000 (Cyanamid International
Corp.) (adapted from Butt and Cotter, 1976).

cohesive within the adhesive to wholly adhesive at the interface as the


joint strength deteriorated. The situation is typical of one in which a
moisture sensitive adhesive is employed without a primer and, in the
case of the solvent degreased substrate, displacement of the adhesive
appears the primary cause of failure.
Where exposure to water or high humidity is expected and, for good
reasons the adhesive chosen is not outstanding in its resistance to
displacement or attack by water, it is customary to seal the edges of the
joint or to coat the assembly with a polymer of low moisture permea-
bility. This is frequently done in the assembly of honeycomb sections
as ingress of water, either as liquid or as vapour subsequently condens-
ing with lowering of temperature, can be trapped and remain in the
cells with catastrophic results. Polysulphide or butyl rubbers can be
used for this purpose though adhesion additives are desirable to make
sure that the bond to the metal is permanent. They must be applied
when the joints they are intended to protect are manufactured. Appli-
cation at some later date is bound to be less satisfactory due to
deterioration of metal surfaces from their condition of optimum bond-
ing when first prepared.
256 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Undoubtedly, where the adhesive necessarily chosen is liable to


displacement by water, the very best practice would be to use a
coupling agent as a primary coat and, if the type of joint or structure
warrants it, also to use a sealing coat although it is the former which
will give the most certain improvement in duration by postponing
displacing for a very long time. Tbe coating will further improve the
joint by slowing down the saturation of the adhesive by water and
hence the possibility that the coupling agent will eventually hydrolyse.
However, if there are periods of relative dryness the adhesive may
never become fully saturated with water and hence breakdown of the
coupling agent will be indefinitely postponed. Coupling agents are
discussed in Chapter 6.

EFFECTS OF TEMPERATURE CHANGE ON


JOINf STRENGm

Information relevant to specific adhesives is of two types. Tbere are


figures relating to the strength of a joint as a function of temperature
and as a function of time during which a given temperature is sus-
tained. The former data ass urne the joint to be held at given tempera-
tures only for aperiod of time sufficient for temperature equilibra-
tion. Since most of the data refer to lap-shear joints of metal
adherends, it records the effect on both modulus and strength of the
adhesive in modifying the strength of the joint. Only at temperatures
elevated to near the limit of use of organic adhesives do changes in the
modulus of the adherend affect joint strength. With carbon fibre
reinforced plastic adherends, changes in stiffness of the adherends may
make the joints more sensitive to temperature. A secondary effect
exists as sub-normal temperatures are used. Tbe bonding having been
carried out at elevated temperature, shrinkage stresses in the adhesive
arise partly from the cross-linking process but also from the differences
between the coefficients of thermal expansion of adherend and adhe-
sive. At very low temperatures, stress arising from thermal origins can
become overriding, causing failure in extreme cases but always causing
some loss of strength. With joints tested in torsional shear, a less
complicated situation exists and changes are wholly those of the
adhesive even the effect of shrinkage stresses being less pronounced.
Changes observed when the joint is held at elevated temperatures
for prolonged periods arise from chemical change and are sensitive to
SERVICE LIFE 257

the presence or absence of oxygen in the ambient atmosphere. It is


usually, and with justice, assumed that the changes are those occurring
in the adhesive but, in air, oxide layers or metals can thicken and
hence weaken the joint. Shields (see Foulkes et al., 1970) used a
fluoroalkylenated aromatic polyimide as an adhesive for steel sub-
strates, heating the joint at 310°C for various periods of time. After
breaking the joints he found the oxide layer on the steel had increased
from something less than 1 nm to possibly 15 nm. Such a change in the
interfacial region must lower the strength of the joint although the
experimental circumstances necessary here precluded strength meas-
urements of the joint.
Figure 106 shows the strength profile of a typical structural adhesive
intended for use over a wide temperature range as determined by
torsional shear (Wake, 1970). This may be compared with the even
greater temperature range used by Althof (1966), but with lap-shear
joints, the profile for which is given in Fig. 116 for the same adhesive,
Hidux 1197A (Ciba-Geigy), an epoxy-phenolic. Althof's results bring
together those from joints made with different thicknesses of adherend
and length of overlap by the use of a form factor f. All joints used to
construct Fig. 116 had a form factor f = 0·1 where f = S~l-\ S is the
thickness of (equal) adherends varied from 1'5-3'0 nun by steps of
0·5 nun and I is the length of overlap. This implies that overlaps
ranged from 1'2-1'7 cm. The lap-shear strength is then given as T = af~
where a is the breaking load per unit area of overlap for a form factor
of unity.
Figure 117 gives the strength profile of an adhesive particularly
designed for use at high temperatures (Imidite 850, Narmco Materials

N
'E 20f__--+
'- z
~ ~
~
on ~
-
.. +'
~ ~10f----+---~---+---4---­
on ..
c: '-
.. +'
.... on
-200 -100 o 100 200 300
Trzmprzrature of test (OCl

FIG. 116. Tensile lap shear strength of joints of thin aluminium sheets of
various thicknesses (see text) as a function of temperature (adapted from
Althof, 1966).
258 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

100-.!"
-----0"____
lji-
~ 80-
• O~O"
a.

.~
" Shields
o Buck and Hockney
• Reinhart and Hidde

~ I I I I
%~----~10~0----~2~00~--~30~0~---4~0~0----~5~00~

FIG. 117. Percentage of strength at room temperature retained at elevated


temperature by polybenzimidazole (Imidite 850, Narmco Materials Corp.) on
stainless stee1.

Corp., a polybenzimidazole). It has been drawn to show the reproduci-


bility of strength retention between different workers using different
methods of test although the actual strengths recorded with the various
adherend dimensions were different. The room temperature lap-shear
strength was about 20 MN.m- 2 (Shields, 1967; Buck and Hockney,
1969; Reinhart and Hidde, 1966). There is far less data on the effect
of temperature on the peel strength of structural adhesives. In general,
however, peel behaviour will exaggerate the changes shown by lap-
shear joints but whereas torsional shear tests show a rise in strength to
a plateau at sub-normal temperatures, peel strength will fall after
reaching a more or less broad peak. The fall occurs as the failure mode
becomes increasingly brittle. This fall occurs also in lap-shear results
when the adherends are insufficiently stiff. A useful bibliography on
the Low and Cryogenic Temperature Properties of Adhesives includes
informative abstracts of the papers quoted (Sims, 1972).
The other aspect of the effect of temperature, namely that of
strength retention over aperiod of time at an elevated temperature, is
illustrated by Fig. 118. Figure 118 has been drawn with data taken from
Burgman and his colleagues (Burgman et al., 1968) for two polyimide
adhesives from the Westinghouse Co. It shows that the usefullife at
371°C is less than 100 h but increases to over 200 h at 329°C and to
well over 1000 h at 288°C. The property measured is the strength at
the exposure temperature. Figure 119 shows the effect of prolonged
high temperatures on the same epoxy-phenolic as was used by Althof
for the figures quoted in Fig. 116. It should be emphasized that the
SERVICE LIFE 259

14~----~-------r

°1~O~-1-----1400~-----10~1------1~O~2-----1~03
Period of exposure to elevated temperature (h)

FIG. 118. Thermal ageing of stainless steel joints made with polyimide adhe-
sives 18 and 140 (Westinghouse Corp.) (adapted from Burgman et al., 1968, by
permission John Wiley & Sons Inc.).

exposure periods to high temperature do not involve any stress im-


posed on the joint by loading. The joints in Figs 116-19 are load free
though not necessarily stress free and the strengths shown will differ
considerably from the average stress at which the joints would break
when loaded and subjected to high temperature. As Althof states
(1966), adhesives which are not heat resistant should not be used at
temperatures above 50°C as no long-term strength is to be expected.
With structural adhesives intended for use at elevated temperatures,
however, the room temperature durability should be obtained at 50°C

FIG. 119. Effect of storage at elevated temperatures on room temperature


strength of stainless steel lap shear joints with an epoxy-phenolic adhesive
(Hidux 1197 A, Ciba-Geigy) (adapted from Althof, 1966).
260 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

~25
'l'
E 20
z 'e
~
60
-g 15 -
.3=:::--
150 I
.Q 175
I
.
~~
'- 200
.. 10
J:
on
!!. 5
.jjj
c:
{!. 0 -
Time to lailure (h)

FIG. 120. Effect of elevated temperature on sustained load. Stainless steel


adherends, epoxy-phenolic adhesive (Hidux 1197A). The lines are drawn as
lower envelopes of the range of the data points to the same scale as Fig. 119
(adapted from Althof, 1966).

and perhaps up to 100°C but above this, the time-to-failure will


depend markedly on temperature as weIl as on the imposed loading.
Figure 120 adapted from Althof, shows the loads which can be
expected to be sustained for given periods with the epoxy-phenolic
adhesive Hidux 1197A (Ciba-Geigy). The lines plotted are the lower
enve10pes of the plotted data points.

SERVICE LlFE AS INDICATED BY CLIMATIC


EXPOSURE TRIALS

So far the expected life of an adhesive joint has been discussed in


terms which describe how its strength is affected when it is held,
unloaded or loaded, for periods of time subjected to known, steady
temperature and humidity or subjected to a sinusoidally varying load
at a steady frequency. Such conditions usually bring about deteriora-
tion at a far greater rate than experience of service conditions shows
normally to obtain. Nevertheless, it would be intuitively expected that
a combination of diurnally cycling temperature together with sunlight
and intermittent rain would constitute a regime more aggressive than a
steady laboratory atmosphere even though it is of high (or low)
temperature and saturation humidity. The comparison of the results of
outdoor exposure to tropical and temperate cIimates with those from
laboratory testing would be expected to give a valuable indication of
actual service behaviour. Such trials have been few and among the
SERVICE LIFE 261

TABLE 28
MEAN ANNUAL WEA'IHER PARAMETERS IN RAE TRIALS

Mean annual Temperate Tropical


parameter
RAE, England Innisfail Cloncurry

Temperature, oe 9 24 26
Rainfall, cm 48 323 28
Relative humidity 83 34

more important of these, incompletely reported. Although interim


details of exposure conditions and some results had been published as
they were obtained (Hockney, 1970, 1972, 1973), the first reasonably
complete account of the trials organized by the Royal Aircraft Estab-
lishment was given by Cotter (1977). These comprehensive exposure
trials were carried out in England and at two tropical sites, one 'hot,
wet' at Innisfail and one 'hot, dry' at Cloncurry in Australia. Cotter
records the mean annual weather parameters of the three sites as
shown in Table 28.
It will be appreciated that these figures are intended to be indicative
of site differences but, naturally, they hide considerable variation. For
example, the rainfall at Cloncurry is concentrated into two or three
months whereas it is uniformly distributed at Innisfail and irregularly
distributed in England. The test joints were double-Iap shear joints
made with clad aluminium alloy to BS 2L73 (high purity aluminium
surface) etched in chromic-sulphuric acids or chromic acid anodized to
UK Ministry of Defence Standard 03-2/1 or Defence Specifications
151 (Type 2) respectively. A range of adhesives, categorized by
chemical type and not identified by trade name, was investigated and
the joints were exposed unloaded and loaded to given percentages of
the proof breaking load. Remaining strength or the time-to-failure
were recorded on test joints removed from the trial at 0·5, 1,2,4 and
6 years. Additionally, strips of 12 gauge alloy (2·8 mm thick) were
bonded with 24 G alloy (c. 0·6 mm) and were peeled at 90°. Such test
pieces were exposed unstressed.
Minford (1982) reported results from tropical exposure trials or-
ganized by the Aluminium Company of America (Alcoa) lasting 12
years although, in these trials, only unstressed joints were exposed.
They were single-Iap shear in form, one series manufactured from
262 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

1· 63 mm Alclad 2024-T3 alloy (high purity aluminium surface) and


one from 3·18mm 6061-T6 aluminium alloy. Named adhesives were
used; a two-part room temperature curing polyamide-epoxy (National
Starch, 82-0688/84-9023) and 3M Company's EC 2086, a one-part
heat cured epoxy-nitrile. Exposure was at Suriname, South America at
two jungle sites differing in the degree of cover afforded. One site was
open to the sky and the other beneath the jungle cover. The latter led
to the presence on the exposed joints of heavy clinging deposits of
mould, fungi and mosses and the complete absence of direct sunlight.
This is referred to in Table 29 as the secluded site and the other as the
open site.
The joints were exposed in two forms; those without and those with
a sealing film of a polysiloxane rubber. Previous Alcoa exposure trials
at maritime sites in Panama had shown a distinct advantage in sealing
with a silicone rubber, presumably due to the exclusion of salt spray.
The jungle trials showed no such advantage nor, in fact, would any
have been expected by the present authors. A lap-shear joint of the
size used in the trial and made entirely with silicone rubber would
equilibrate at the interface at the centre of the joint in a matter of
hours due to the very high permeability of silicone rubber to moisture.
If periods of drying-out were operative then advantage might be
expected by sealing with polysulphide or butyl rubber but even these
would be ineffective in the presence of growths of mosses.
An attempt is made in Table 29 to summarize the experience
obtained with unloaded joints in these two important exposure trials
but, such is the nature of a summary in this context, that caution must
be exercised in drawing conclusions from the table. The results
selected refer to joints made with alloys clad with high purity
aluminium in all cases with etched surfaces.
The introduction of stress during the climatic exposure markedly
inftuences strength retention. Figures 121-3 illustrate this and refer, in
numerical order, to adhesives F, G and H of Table 29 at the hot, dry
and hot, wet tropical exposure sites over 4 years. The stress applied
was tensile and the level is expressed as a percentage of the proof
ultimate tensile breaking stress. It was continuously applied by dead
load.
From the authors' discussions and the published data about these
trials the following summaries have been compiled referring to the
adhesives employed.
A. Epoxy-polyamide. Time-to-failure studies on this adhesive are
SERVICE LIFE 263

reported earlier in this chapter. Undoubtedly, the relatively high water


absorption of this type of adhesive gives a joint which performs badly
in humid conditions and failure occurs at the interface. This interfacial
sensitivity is reflected in two ways. Firstly, the rate at which the bond
with aluminium deteriorates is very sensitive to the surface preparation
of the adherend; secondly, bond failure is frequently accompanied by
corrosion of the adherend if it is a susceptible aHoy. Although the
corrosion undoubtedly occurs after the adhesive has been displaced
from the surface, the fact remains that joints with corrodible aHoys do
not last as long as those clad with pure aluminium. Simple vapour
degreasing which can be used with some adhesives (even though it is
not advised for the strongest and most durable bonds) is definitely
inadequate with epoxy-polyamides if exposure to weathering is likely.
In the absence of stress, an epoxy-polyamide (National Starch 82-
0688/84-9023) behaved unexpectedly wen in the secluded jungle evi-
ronment.
B, C and D. Modified epoxies. Unfortunately, the term 'modified'
encompasses two classes of epoxy resin. The commonly used starting
material bis-phenol A may be modified by admixture with other
phenols or alcohols before or after reaction with epichlorhydrin, or the
epoxy resin may be modified by the introduction of a rubbery compo-
nent either by blending or by forming the epoxy in the presence of the
rubber. The composition of Band C may be of the former type (it is
not disclosed); D is of the latter type. Adhesive D did not perform so
weH as the epoxy-polyamide (A type) in the secluded jungle atmos-
phere and one must assume that some biological attack on the nitrile
rubber is more severe than on the polyamide even though both contain
nitrogen which tends to support growth in some form or other.
F, G and H. Phenolic-modified adhesives. All these adhesives per-
formed weH in the absence or presence of continuous stress up to 20%
of the proof strength. This comment should not be taken to imply that
stresses above that level could not also be borne for prolonged periods;
merely, that this stress was the highest applied during the trials which,
as far as Cotter has reported results (1977), have extended for 6 years.
Nevertheless, stresses in excess of 20% have not, hitherto, been used
for design purposes and the stress distribution in the double-lap joints
used in this trial will certainly involve end stresses greatly in excess of
this figure which is that of the average stress over the bonded areas.
Adhesives G and H were both applied over a primer (Hockney,
1972) which is, presumably, a phenolic primer. It is not surprising
TABLE 29
STRENGTH RETENTION PROPERTIES OF UNLOADED, EXPOSED JOINTS

Adhesive Time in years for 50% loss of strength or comment on performance


type
Open Secluded
Temperate climate Tropical: hot, dry Tropical: hot, wet jungle jungle
Double-lap Double-lap Double-lap Single-lap Single-lap
shear 90° Peel shear 90° Peel shear 90° Peel shear shear

A Epoxy- Indefinitely* Indefinitely Indefinitely >6 1·5 2 2 >12


polyamide
B 2-part, room 4·5 Strength Indefinitely Indefinitely Strength 3
temperature levelled after levelled after
curing, mod- 30% loss 30% loss
ified epoxy
C 2-part, heat cur- Strength Dropped 10% Indefinitely Strength Strength Strength
ing modified levelled after in 4 years. dropped levelled after levelling af-
epoxy 20% loss Loss con- 15% then re- 20% loss ter 15% drop
tinuing at covered in 4 years
same rate
D I-part, heat cur- S >12t
ing nitrile
modified
epoxy
E Poly vinyl Indefinitely Varied +20% Indefinite Strength Indefinite Strength
formallphe- -10% dropped dropped
nolic 50% in first about 20%
year but re-
covered
thereafter
F Epoxy-phenolic Strength 6 Initial fall in Strength Rapid fall of Strength
dropped strength fol- dropped 20%. Slow dropped
15% in 6 lowed by 30% rapidly fall con- 40% in 2
years levelling then fell a tinuing at 6 years then
further 10% years levelled
over 4 years
G Epoxy-novolac Indefinitely Increased in Indefinitely Increased in Indefinitely Indefinitely
strength strength
H Nitrile-phenolic ( Indefinitely
"
* This indicates that the proof strength was maintained ±10% throughout the reported period of 6 years. It also implies that there is no
indication at 6 years of imminent change.
t This refers, as do all results in the table, to alloys clad with pure aluminium. In the trial (Minford, 1982) from which this result is taken unclad
aluminium alloy joints lost 50% strength in 9 years.
266 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

...
c
~
Co
.c'
e;.
c
~
c'"
.§.

..
t:;;

Cf'
c
o
.c
U

..V
c
u
T i me In yeors

Co

:f
Cf'
c
~
'"
c
:§. -1 5
,-
.=
Cf'
c
-20
\

0
.c
U
-25
0

T,me In yeors

FIG. 121. Behaviour of an epoxy-phenolic adhesive on extended exposure trial


(data from Hockney, 1973).
SERVICE LIFE 267

c.,
u +10
Q;
0-
.<:
;;,
c:
~
7ii
'E
:§.
.=
.,
c'"
0
.s::.
0

Time in years

FIG. 122. Behaviour of an epoxy-novolac adhesive on extended exposure trial


(data from Hockney, 1973).

therefore that their stability in the presence of high humidity and


somewhat elevated temperature should be very similar (cf. Figs 122
and 123). It would seem from the diagram in Cotter's artic1e which
extends the results shown in Figs 122 and 123 to six years and
obviously smoothes them somewhat, that the fall shown in Fig. 122 does
not continue the full six years but is modified and levels off. Unfortu-
nately, the disparity of scales tends to give different impressions from
268 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Time in yeors

.,
'"
c
o
_
oe
U

o
Time in yeors

Fro. 123. Behaviour of a nitrile-phenolic adhesive on extended exposure trial


(data from Hockney, 1973).
SERVICE LIFE 269

what is the same set of data. The fact remains that a 10ss in strength of
less than 20% in six years indicates a substantially stable situation and
one in which the additional severity of the Innisfail tropical site over
the temperate site at Farnborough made but Httle significant differ-
ence in properties although the loss of strength at Innisfail was actually
somewhat greater.
The generally expressed opinion elsewhere in this book of nitrile-
phenolic adhesives as the valuable 'work horses' of structural adhesives
is amply brought out in these trials. Adhesive F (Hockney, 1970) was
used without a primer and was carried on a glass cloth. The initial drop
in strength was shown also at the temperate sites and by the
laboratory-stored controls according to Hockney's data (1973). No
explanation has been advanced for this but it may relate to an
inadequacy in the curing of the test pieces. The loss is not serious and
is rapidly recovered. Figure 114 refers, inter alia, to the behaviour of
an adhesive of this type in running water at 70°C when a loss of about
22·5% in the strength of an unloaded single-lap joint is indicated as
occurring in 10 weeks. In spite of the effect of combining a high (20%
of proof strength) stress with daily temperature and humidity cycling,
as is occurring in a tropical situation such as obtains at Innisfail, such
exposure is far, far less severe than the unloaded laboratory test which
shows in 10 weeks much the same loss of strength as occurs in 6 years
at Innisfail on a loaded joint.
E. Polyvinyl formal/phenolic. Adhesives of this type (e.g. Redux
775, Ciba-Geigy) have now been used in structural engineering
for a longer period than any other adhesive. In particular there is a
wealth of experience in aircraft manufacture. Considerable importance
is therefore attached to the reporting of the effect of exposure
and stress on joints made from material of this type. Apart from
the rather small scale graphs given by Cotter (1977) which cover
six years, recourse must be had to Hockney's detailed figures covering
four years only. These figures present an extremely confused picture
which suggests when joints are stressed they are indeed sensitive to
climatic variation on a short time scale in the sense that conditions can
bring about a loss of strength which later conditions can reverse
provided that the stress is not too high. Hockney (1973) reports the
failure of 28 joints loaded to 15% stress at Innisfail. The failures
occurred irregularly spread over a wide period from 6 weeks to 206
weeks with a median time-to-failure of 73 weeks. However, this figure
is based on joints that failed from an unknown number that completed
270 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

the trial at this loading. It would appear, therefore, that this dass of
adhesive is unable continuously to sustain loads greater than, say 10%
of the proof stress in a wet, tropical dimate though it can be loaded to
15% where the joint is not continuously wetted. Unlike the other
phenolic-containing adhesives, loading to 20% of the proof stress does
not seem advisable even in a dry, tropical dimate. Comparison of
Hockney's figures for peel resistance with the strength of the double-
lap shear joints suggests that ductility is at the root of its behaviour.
Both moisture and elevated temperatures increase ductility and reduce
load bearing though a modest increase in moisture associated with a
similarly modest change in ambient temperature improves rather than
decreases strength.
Chapter 8

Applications

There are today very many examples of structures where a main


load-carrying component transfers its load to another through an
adhesive layer. Adhesive manufacturers have published photographs of
such applications together with details of the adhesive used and usually
with a briefly worded note either of the advantages sought by using
adhesive rather than some other method of joining or the actual
reasons for the choice. The loads borne or stress analyses of vital
sections of the structure are not given and the engineer is left with the
message that the structure is sound from an engineering aspect without
the data wherewith structural features could be carried over into some
other structure or design. This is, of course, usual. Published accounts
of a new bridge may mention the maximum expected loading on the
piers or pull on the cable anchorages but more detailed figures would
be virtually meaningless unless very detailed indeed and supported by
comprehensive drawings. In many cases the shock-Ioaded stresses are
unknown and trial has established the eflicacy of the bonding process
either as superior to welding, brazing, or mechanical fastening, or equally
satisfactory but more economical or better adapted to a given produc-
tion line. Although there are a few case histories of design published, it
is diflicult to use these in a systematic presentation or to avoid the
mere compiling of a bibliographical catalogue in the form of a list of
references with titles.
One of the difliculties is that of laying down effective criteria for the
use of adhesive rather than any other method of fastening. So much of
engineering design in its broader conception rests on a subjectively
appreciated 'fitness for the job'. Engineering elegance in design is
recognized by all engineers but, although it can be analysed in specific
271
272 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

cases, it is impossible to define. It is an amalgam of aesthetics and


stress distribution.
With these difficulties in mind, the following brief alphabeticallist of
applications is offered with such comment as is available on the details
of the design, the stresses, or the reasons for the choice of adhesives
for the structure. In this connection attention must be drawn to a
chapter in a book by Semerdjiev (1970) which discusses a similar
variety of applications of the adhesive bonding of metals. The numer-
ous examples frequently inc1ude fairly detailed drawings dimensioned
in essential aspects though, unfortunately as elsewhere, information on
stress levels is absent. In many cases, of course, this information has
never been obtained or ca1culated and the application stands or falls by
its performance. If it works, it works! An example quoted by Semerd-
jiev is that of bonding tungsten carbide or other cutting materials as
the teeth of cutters in milling machines. Shapes and their correspond-
ing cavities are shown in diagram together with the surfaces to be
bonded. The application is, apparently successful and the actual stress
at the adhesive joint becomes irrelevant.

Aircraft (see also P ABST)


As others have remarked, it is odd that the aircraft industry, in which
safety and reliability command a higher degree of attention than in
almost any other product, should be the industry which departed
soonest from traditional methods and made the most extensive use of
adhesives. Adhesives bonded airframe structures were in British air-
craft during the Second World War and since the War, a particularly
successful turboprop aircraft, the Fokker Friendship, made more use of
adhesives than any other aircraft. Semerdjiev (1970) states that over
55% of the surface area is bonded, panel sizes of 1·3x4·9m being
involved together with spar booms adhesive1y constructed 10 m long.
Testing, inc1uding 14·5 x 106 cyc1es of reverse loading, established the
integrity of the wing assemblies. Experience of safe operating extends
forwards from the 1960's to the present time with later models
increasing their dependence on adhesives compared to earlier craft.
Modern military aircraft involve vital stressed components where
transfer is from CFRP to metal via adhesive bonding. Honeycomb
constructions are extensively used in aircraft, frequently involving
dissimilar materials (metallic and non-metallic) for the skin and core.
Their advantages reside in their stiffness, strength and light weight as
well as the smoothness obtainable on their surfaces with the absence of
APPLICATIONS 273

rivets. These honeycomb panels may be contoured as aerofoils or simply


used as flooring. Carbon fibre reinforced skin with a honeycomb core
and used as flooring can save 300 kg payload on aircraft such as the
Boeing 747.
The new Lear Fan aircraft is the first to be made with a very high
proportion (about 80%) of the structural weight in carbon fibre rein-
forced plastics. This ambitious project uses adhesive bonding for
joining the composite components and panels together as well as for
attaching pre-formed stringers and stiffeners.

Anchorages
Epoxy and polyester adhesives are used to provide bolted anchorages
in concrete. Anchorages are normally designed with stresses such that
the adhesive is frequently subjected to compressive as well as shear
stress and the interface with the metal is similarly protected. Where the
placement is to be to existing concrete, there is little alternative to
drilling a cylindrical hole with resultant shear at the adhesive-concrete
interface. If the anchor can be inserted before pouring the concrete,
then a compressive element can similarly protect the adhesive-
concrete interface from transferring the load wholly by shear. Adhe-
sion along the length of the inserted anchor or bolt is, apparently,
disadvantageous and adhesion of the emerging end can be prevented
by winding PVC insulation tape extending this debonding zone by
perhaps half of the inserted length. Mayfield et al. (1978) used
block-ended bolts 16 mm diameter with 21 mm diameter block ends
25 mm long made of EN 16 steel to investigate the pull-out behaviour
from cylindrical holes 23·5 mm diameter. The adhesive used is not
specified but appears to have been a polyester resin rather than an
epoxy. The usual design formula relating embedded length, I (assumed

TABLE 30
DESIGN AND FAlLURE LOADS FOR BONDED ANCHORAGES (after Mayfield et
al.,1978)

Hole depth Design load Mean failing


(mm) load (kN)
P = (1- 50)/25 (kN) P = 41(l + d) (kN)

75 10 15 35
100 20 30 71
125 30 50 91
274 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

bonded throughout) to load, P, is 1= (50+25P) mm, where P is in


metric tonnes. An alternative includes the diameter of the hole, d, in
the calculation and gives P = 81 (I + d). The authors argue for a safety
factor of 2 to be applied to this expression. These formulae relate to
cylindrical bolts without the block-end. Table 30 compares the design
loads calculated for fully bonded cylindrical bolts with the experimen-
tally observed ultimate loads borne by block-ended, partially de-
bonded anchorages. The advantage is obvious.

Bridges
An interesting application to bridge construction concerns the bonding
with epoxy adhesive of additional reinforcement either to remedy
deficiencies in design or construction or, more usually, to allow for
increases in the designated loading. The first reports (L'Hermite and
Bresson, 1971) refer to work carried out a few years earlier.
Mor~ recently the Transport and Road Research Laboratory of the
UK has reported similar work on motorway bridges (Raithby, 1980;

FIG. 124. Side span of bridge at Quinton Interchange (M5) showing bonded
strengthening plates (Crown copyright).
APPLICATIONS 275

Macdonald and Calder, 1982). Figure 124 shows clearly the size and
position of the bonded mild steel plates. They are 6· 5 mm thick and
254 mm wide. Surface preparation of both plate and concrete was by
grit-blasting after which epoxy adhesive (type not stated but obviously
a low temperature curing two-part epoxy) was 'buttered' on by hand.
The plate was then presented to the concrete surface and held in
position until cured by a cradle with rods screwed into the soffit,
wedges being used to ensure good pressure to squeeze out excess
adhesive. After adequate curing, loading tests, which had been made
before the placement of the plates, were repeated on one bridge which
had shown some cracking. Flexural stiffness was found to have been
increased by about 11 % and the opening of existing cracks was
reduced under load by 35-40%. Raithby also reports parallellabora-
tory tests on fabricated beams which are given in greater detail by
Macdonald and Calder.
Several points may be noted:

(i) There is a corrosion problem present which appears to arise


from the transport of moisture through the adhesive not so
much by diffusion as through the agency of micro-cracks, in
part associated with filler particles in the adhesive. Chromating
the steel seems advantageous without noticeable loss of adhe-
sion.
(ii) Failure of externally exposed experimental beams has shown
that environment factors are important and in continuously
loaded beams can cause separation at the adhesive/steel inter-
face apparently spreading from the ends. Apart from the
absence of protection against corrosion, it is possible that the
end cleavage stresses in these 3·5 m laboratory beams may be
much larger than in the actual applieation. If this is a weak-
ness, ehamfering of the edge, for reasons indieated in Chapter
2 would be worth while.
(iii) On the actual bridges there is, of course, no dead shear loading
although there is a small normal load caused by the weight of
the plate. The loading exists only as transients with the passing
of traffie and possibly temperature indueed changes.
(iv) The reinforeing plates were plaeed in position and the adhesive
allowed to eure whilst the bridge was in use by traffie. (See also
Segmental Construetion.)
276 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Carriages
Carriages for underground railways require fast acting sliding doors.
These must therefore be of light-weight construction. They are fre-
quently of aluminium frame stiffened by adhesively secured panelling.
The surface area for bonding is relatively high and the stress on the
bond therefore relatively low so that it can be accommodated by
semi-structural adhesives. Depending on location, it is possible to use
polyurethanes, nitrile or polychloroprene adhesives but in tropical
situations cross-linking nitrile phenolics would be safer.

Cars
With the recent need to conserve fuel, alternative constructions
minimizing the use of steel in car bodies are being considered by
manufacturers in all countries. Little of substance has been published
about the experimental vehicles now under road trials. The principal
areas where adhesives seem to be utilized are those where light-weight
metals are being fastened to steel framework, aluminium alloys to
themselves and these materials to various load-bearing, plastics mem-
bers or panels. An example of high stress application with shock
loading appears to be involved in the bonding of vertical door posts to
the horizontal members of the main frame. There is, of course,
extensive use of adhesives (and sealants) in non-structural applications
and also in structures where the stresses are minimal. An example of
the latter is the fastening of stiffeners and anti-vibration members
across the covers of luggage boots and engines (bonnets) of the
ordinary family car. The trials with structural adhesives were initially
confined mainly to epoxies and their various modifications with rub-
bery materials but, more recently, the toughened acrylics are being
favoured on account of their combination of rapid curing with toler-
ance of oily surfaces. A difficulty with both types of adhesive could be
the temperature of the stoving tunnel at the end of the finishing
process.

Decking
Ciba (ARL) (1969) (now Ciba-Geigy, Ltd) report the bonding of
concrete pre-cast beams to steel stringers prior to resurfacing a bridge
across the river Trent at Keadby, Yorkshire. The beams, weighing
255 kg each, took the direct weight of the road surfacing (concrete)
and the moving vehicles. The loading was therefore compressive with
additional transient compression and shear. The adhesive was applied as
APPLICATIONS 277

a sand-filled epoxy two-part compound. A major consideration in its


use was the equi-distribution of the loads direct to the stringers
ignoring the previously used (and failing) riveted plates. The load
bearing capacity of the bridge was improved from its previous figure.

Furniture
Chair frames, whether of metal or wood are subjected to very high
torsional and cleavage stresses. Metal frames can be made by shaping
tubes but aesthetics have led to the use of other extruded sections
which do not lend themselves to fabrication by this means. Electro-
plated castings have been used as junction pieces where vertical and
horizontal members meet in the conventional or modified chair shape.
Into these castings the straight members are inserted and held by
adhesive. The tension is thus contained with the adhesive in shear with
torsion and cleavage taken by the solid casting. Wooden furniture is
dowelled, rigidity being obtained by the use of adhesives which, in
addition to possessing strength and durability are also gap-filling.
Polyvinyl acetate emulsions are formulated to meet these conditions
but urea-formaldehyde adhesives, two-part, can also be used.

Glass Reinforced Plastics


Glass reinforced plastics have entered the ship building industry as a
major method of constructing war ships. It has tended to be more
expensive a method than that from steel plate, but has advantages
where non-magnetic properties are desirable and also for life-boats
where double-hulls together with resilience and lightweight construc-
tion are at a premium. Part of the expense of manufacture with GRP
arises from the amount of hand laying-up followed by brush work with
the resin when minor seatings are required attached to the main or the
inner hull. An alternative method of introducing brackets, or cross
members is to pre-fabricate the required GRP shape and then bond in
place with a suitable adhesive.
Toughened acrylics have been used for this purpose and the technol-
ogy is discussed by Bowditch and Stannard (1981). A major advantage
of adhesion bonding in manufacturing processes involving GRP is the
ease with which smooth surfaced unions of GRP-to-GRP or GRP-to-
metal can be made. Since substrate failure can be achieved with a
toughened acrylic and GRP, the failure strength of GRP to itself or
to aluminium are virtually identical at a shearing force of about
10MN.m-2 .
278 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Helicopters
Boeing Vertol's new Model 360 fuselage shell uses composite materi-
als such as a sandwich of Kevlar diagonal weave skins with a Nomex
honeycomb core and carbon fibre reinforced plastics for reinforcing
edges and window cut-outs. Tbe panels are bonded and the bonds are
stressed to carry the full fuselage loads. Metallic fasteners are also used
to facilitate assembly and jigging.

Helicopter Blades
Helicopter bl ades are complex structures which are highly stressed and
of limited fatigue life. Tbey are also expensive and if their critical
shaping involved manufacture from bulk metal they would be even
more expensive. Instead the shape of the Westland Lynx helicopter
blade is contoured to the stress pattern by bonding stainless steel in
steps with a nylon-modified epoxy. Helicopter blades are, in fact,
almost wholly dependent on adhesives. Tbey consist of an extruded
aluminium spar which carries the aerodynamically shaped blade, con-
toured with bonded metal sheets towards the root of the blade as stress
increases, and made rigid in compression by fitting the cavity with
honeycomb stiffening. Tbis last may be of resin impregnated paper or
aluminium foil but in either case the cell edges are bonded to their
covering surfaces by a suitable filleting adhesive for it is the controlled
fillet that ensures the rigidity and integrity of the honeycomb structure.
Details of a field repairable-expendable blade are given by Falcone
and Miller (1977). Its main features are a single piece extruded spar, a
nomex core and GRP skin. Tbe trailing ege is of aluminium. Repair
kits are based on the insertion of surfaced plugs secured by epoxy.

Hovercraft
Tbe main structural component of the SR.N 4 (British Hovercraft
Corporation) according to Powis (1973) is a large sandwich panel 40 m
long, 23 m wide and about 1 m thick. Tbe top and bottom skins of this
sandwich panel are made of 2·5 mX 1· 25 mX 37 mm sandwich panels
which are bolted to an egg-box structure made from aluminium shear
web with bonded stiffeners. Tbe skin panels are of 6·4 mm and
3·2 mm cell aluminium honeycomb made from foil 0'05-0'076 mm
thick. Tbe higher density core is used in the more highly stressed areas.
The superstructure of the craft is composed of two vertical beams
running fore and aft, which are fabricated from this aluminium sheet
with bonded stiffeners. Tbe roof is one unsupported span between
APPLICATIONS 279

these longitudinal beams and is made from thin aluminium sheet with
bonded corrugated stifIeners supported by bonded I beams.

Lamp Posts
Aluminium alloy bonded lamp posts were put into service in the early
1960's. One of these triangular tapering structures (about 4 m long) is
pictured, in an article by Garnish (1982), being carried by one man on
his shoulder. The post is fabricated from three long but narrowing
pieces of metal bent to give the 60° corner along one edge and
provided with a lip at the other. The edge from the 60° bend fits into the
lip of a second piece, and this with the third, to give a tapering post
with an equilateral triangle of decreasing size as the cross section
ascends the post. The edges are bonded into the lips with a two-part,
room temperature curing epoxy which resists the only possible opening
stress by shear. Bolger (1977) gives details of a very similar construc-
tion adopted by the Crouse Hinds Corporation (of the USA) in their
'Tri-Round' lamp post. This tapers from a 12 in triangular base to a
6 in diameter top. Three shaped aluminium alloy extrudates are joined
by three tapering pieces of aluminium sheet by slotting the sheet into
slots in the extrudates and bonding them with a two-part polyamine
cured epoxy.

Magnets
Armatures for magnets or the rotors of dynamos and electric motors
are laminated from ferrous alloy sheets for reasons associated with the
form of the magnetic field and the contaiI'Jllent of eddy currents. The
massive magnet cores required in atomic energy devices such as that of
the CERN organization, involve adhesives as the main structural
component. In the CERN magnets, Araldite epoxies were used to
bond the laminated stack of thin ferrous sheets into the solid core
(Garnish, 1982). Epoxies were preferred becausethey are reasonably
radiation resistant. In less massive constructions polychloroprene con-
tact adhesives have been used for bonding the lamellae of rotors.

PABST
The PABST programme is an imaginatively conceived and typically
American attempt to advance significantly the use of bonded structures
in aircraft. The acronym is derived from Primary Adhesively Bonded
Structures Technology and the programme is based on the develop-
ment of a section, 42 ft long, of a particular transport aircraft
280 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

designed for short runways. As already detailed, adhesives have been


used for main stress-bearing members of aircraft bodies from the
wartime Mosquito (in which synthetic adhesives were used in wood
structures), the De Havilland 'Comet' and the more recent Fokker
'Friendship' both of which made or make extensive use of Ciba-
Geigy's Redux 775. A very detailed assessment of the results of the
programme is given by Thrall (1980). These may be summarized as
follows:

(i) Following tests on various panels and sections, the 42 ft long


section of the YC-15 aircraft fuselage was built and subjected
to a pressure-cyc1e test to demonstrate fail-safe features of
large flaws in the bonded fuselage skins. An arrestment and
subsequent 90-degree turning of a longitudinal crack as it
contacted the bonded frame shear tee occurred. 'This feature
does not exist in a riveted structure and therefore allows high
panel stresses with assurance of crack containment.'
(ii) Fewer fatigue cracks occur when compared to similar tests of
riveted structures. After 68 000 pressure cyc1es on the bonded
structure, 9 developed, all associated with fastener holes. (The
press ure cyc1es were between 1 and 4 cyc1es per hour plus
environmental change. A faster rate is not recommended).
(iii) The adhesive used. in the final trials was a toughened epoxy
(FM 73, Cyanamid Corp.) applied over a priming layer
(BR 127). FM 73 is arecent introduction replacing FM 123-2,
a nitrile epoxy, also used in the programme of testing.
(iv) Clad aluminium alloys exhibited a lower environmental resis-
tance at the bond line and it was conc1uded that 'No c1adding
should be allowed in the faying surface of a bond joint'
(Shannon and Thrall, 1977).

Rollers
Ciba-Geigy report the use of their adhesives (Garnish, 1982) for the
assembly of rollers for motor mowers. The roller consists of a mild
steel cylinder, with slotted surface to provide grip on the grass as the
drive traversing the mower is provided by this cylinder. The end plates
carrying the bearings are adhesively bonded to the cylinder by fitting
the cylinder over flanges on the plates. The adhesive withstands shear
during driving, direct tension with temperature changes, impact resis-
tance when mishandled on concrete roadways and paths and accom-
APPLICATIONS 281

modates the machining tolerances which require some gap-filling prop-


erty.

Segmental Constrnction
During the last 20 years or so, there has been increasing interest in the
technique of segmental construction using pre-cast concrete units made
under factory-controlled conditions in a casting yard which may be
remote from the construction site. The sections are then erected and
post-tensioned on the site. To achieve a good fit, each segment is
match-cast against the end of its neighbour, using a release agent to aid
separation. This technique has been widely used for tunnels, stadia,
lighthouse foundations and other civil engineering constructions. Al-
though bridges have already featured in this survey of applications, it is
convenient to refer again to them because of the high stresses involved
and their importance. The bridge over the Seine at Choisy-Ie-Roi in
Paris, completed in 1963, was the first example of this type of adhesive
construction. Of particular interest in the UK is the Byker Viaduct,
S-shaped in plan, 800 m long and forming part of the Newcastle-on-
Tyne Metro. This viaduct was the first bridge to be built in the UK
using cantilever construction with precast segments joined by epoxy
adhesives (Seymour, 1980). 253 segments were involved, varying in
size but generally 2·25 m deep, 3·3 m long and a top width of 8·6 m.
The heaviest weighed 46 tonnes.

Ski Constructions
Semerdjiev (1970) gives details of the seven layers used in the con-
struction of a make of adhesively bonded ski. The materials of the
layers are diverse comprising the following: phenolic-bonded fibre
sheets, spring steel, aluminium and wood particle board (to provide the
damping and dynamic properties necessary for control). Bonding to-
gether seven layers is stated to be by a modified phenolic adhesive. It
may be assumed that a nitrile-phenolic adhesive would be the material
of choice though the surface preparation of the components, particu-
lady of the outer skins of phenolic/fibre, could be exacting.

Telephone Kiosks
Keimel (1966) reported a design exercise on fabricating telephone
kiosks with aluminium L-sections bonded together instead of welding
or riveting. This design was not actually carried forward into produc-
tion because although comparison of costs initially showed highly in
282 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

favour of bonding as against welding, the competition led to such


improvements in welding aluminium on a production line that the final
advantage lay marginally with welding. Rivet construction was costed
at 7 times bonded construction, the main cost savings arising from the
elimination of perforating, shorter time of assembly and almost total
elimination of finishing operation.

Yachts
The Nelson dass yacht Teneriffe, designed by Uffa Fox and built by
Pochin's (Manchester) Ltd, used adhesively laminated wood for keel,
stern, ribs, bilge strings and gunwale. The planking was also bonded to
the ribs with a resorcinol-formaldehyde wood adhesive (Aerodux 185,
Ciba-Geigy).
References

Adams, R. D. (1980) Adhesion 4, Allen, K. W~, Ed., Appl. Sei. Publ., London,
p.87.
Adams, R. D. (1981) Developments in Adhesives-2, Kinloeh, A. J., Ed.,
Appl. Sei. Publ., London.
Adams, R. D. and Coppendale, J. (1976) J. Meeh. Eng. Sei. 18, 149.
Adams, R. D. and Coppendale, J. (1977) Adhesion 1, Allen, K. W., Ed.,
Appl. Sei. Publ., London, p. 1.
Adams, R. D. and Coppendale, J. (1979) J. Adhesion 10,49.
Adams, R. D. and Peppiatt, N. A. (1973) J. Strain Anal. 8, 134.
Adams, R. D. and Peppiatt, N. A. (1974) J. Strain Anal. 9, 185.
Adams, R. D. and Peppiatt, N. A. (1977) J. Adhesion 9(1), 1.
Adams, R. D. and Peppiatt, N. A. (1977a) Fibre reinforeed materials, design
and engineering applieations, Paper 6, lust. Civil Engrs.
Adams, R. D., Chambers, S. H., DeI Strother, P. J. A. and Peppiatt, N. A.
(1973) J. Strain Anal. 8, 52.
Adams, R. D., Coppendale, J. and Peppiatt, N. A. (1978a) Adhesion 2, Allen,
K. W., Ed., Appl. Sei. Publ., London, p. 105.
Adams, R. D., Coppendale, J. and Peppiatt, N. A. (1978b) J. Strain Anal. 13,
1.
Adams, R. D., Peppiatt, N. A. and Coppendale, J. (1978e) Predietion of
strength of joints between eomposite materials, Symposium: Jointing in Fibre
Reinforced Materials, Imperial College, London, IPC Seienee and Teehnol-
ogy Press Ltd, London, p. 64.
Allan, G. G. and Neogi, A. N. (1971) J. Adhesion 3(1), 13.
Allen, K. W. and Alsalim, H. S. (1977) J. Adhesion 8(3), 183.
Allen, K. W. and Shanahan, M. E. R. (1975) J. Adhesion 7(3), 161.
Allen, K. W. and Shanahan, M. E. R. (1976) J. Adhesion 8(1), 43.
Allen, K. W., Alsalim, H. S. and Wake, W. C. (1974) J. Adhesion 6(1/2),153.
Allman, D. J. (1977) Quarterly J. Meehanies Appl. Maths. 30, 415.
Althof, W. (1966) Metall 20(10), 1042.
Althof, W. (1977) SAMPE Symp. Proe. 22, 784.
Althof, W. (1981) The influenee of moisture on adhesive bonded joints,
Adhesion 5, Allen, K. W., Ed., Appl. Sei. Publ., London, p. 15.
283
284 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Althof, W. (1982) Int. Syrnp. on Adhesive Joints, Kansas City, Proc. Org.
Coatings Appl. Polym. Sci., 184 Nat. Meeting, Amer. Chern. Soc., p. 267.
Alwar, R S. and Nagaraja, Y. R (1976a) J. Adhesion 7, 279.
Alwar, R S. and Nagaraja, Y. R (1976b) J. Adhesion 8, 79.
Anon. (1973) Plastics & Rubber Weekly, 16th March.
Anon. (1974) Plastics & Rubber Weekly, 18th October.
Aponyi, T. J. (1977) Handbook of Adhesives, Skeist, 1., Ed., 2nd ed., van
Nostrand-Reinhold, New York, Chapter 38.
Argyris, J. H. (1962) Unpublished test results quoted (W. C. Wake), Research
15, 183-90.
Barker, R M. and Hatt, F. (1973) Amer Inst. Aero. Astro. J. 11, 1650.
Barth, B. P. (1977) Phenolic resin adhesives, Handbook of Adhesives, Skeist,
I., Ed., van Nostrand-Reinhold, New York, p. 382.
Bascorn, W. D., Jones, R L. and Timmons, C. O. (1976) Mixed-mode fracture
of structural adhesives, Adhesion Science and Technology, Lee, L-H., Ed.,
Part B, Plenum, New York, p. 501.
Bates, R (1971) Communication frorn Evode, Ltd, Stafford, England.
Benham, P. P. and McCammond, D. (1971) Plastics Polymers, 39(140), 130.
Benson, N. K. (1969) Infiuence of stress distribution on joint strength,
Adhesion-Fundamentals and Practice, UK Ministry Technology, MacLa-
ren, London, p. 191.
Benthem, J. P. and Minderhoud, P. (1972), Int. J. Solids Structures 8, 1027.
Bijlmer, P. F. A. (1978) Adhesion 2, Allen, K. w., Ed., Appl. Sei. Publ.,
London, p. 45.
Black, J. M. and Blomquist, R F. (1962) Adhesives Age 5(2), 30; 5(3), 33.
Bolger, J. C. (1973) Structural adhesives for metal bonding, Treatise on Adhe-
sion and Adhesives, Vol. III, Patrick, R L., Ed., Dekker, New York,
Chapter 1.
Bolger, J. C. (1977) Durability of adhesive bonded structures, Bodnar, M. J.,
Ed., Appl. Polym. Symp. No. 3, Wiley, New York, p. 369.
Bossler, R c., Franzblau, M. C. and Rutherford, J. L. (1968) J. Sci. Inst.: 1.
Phys. E. 21, 829.
Botsco, R (1968) Matls. Eva!. 26, 21.
Bowden, P. B. (1973) The yield hchaviour of glassy polymers, The Physics oj
Glassy Polymer, Haward, RN., Ed., Appl. Sei. Publ., London, p. 279.
Bowden, P. B. and Jukes, J. A. (1972) J. Mat!. Sci. 7, 52.
Bowditch, M. Rand Stannard, K. J. (1981) Bonding GRP with acrylic
adhesives, Adhesion 5, Allen, K. W., Ed., Appl. Sei. Publ. London, p. 93.
de Bruyne, N. A. (1939) Flight Suppl. Aircraft Engineer, 29 Dec.
de Bruyne, N. A. (1951) Adhesion and Adhesives, de Bruyne, N. A., and
Houwink, R, Ed., Elsevier, Arnsterdarn, p. 91.
Bryant, R W. and Dukes, W. A. (1965) Brit. J. Appl. Phys., 16, 101.
Buck, B. I. and Hockney, M. G. D. (1969) Adhesion: Fundamentals and
Practice, UK Ministry Technology, MacLaren, London, p. 124.
Burgman, H. A., Freeman, J. H., Frost, L. W., Bower, G. M., Traynor, E. J.
and Ruffing, C. R (1968) J. Appl. Polym. Sei. 12, 805.
Butt, R I. and Cotter, J. L. (1976) J. Adhesion 8(1), 11.
Cassidy, P. E. anel Yager, B. J. (1972) Reviews in Polymer Technology I, Skeist,
I., Ed., Dekker. New York, p. 1.
REFERENCES 285

Chugg, W. A. and Gray, V. R. (1965) Conference on Plastics in Building,


Plastics Institute, London.
Ciba (ARL), Ltd (1969) How to design for Araldite bonding, Technical notes,
Duxford.
Coker, E. C. (1912) Proc. Royal Soc. 86(A587), 291.
Coleman, B., and Knox, A. G. (1957) Text. Res. J. 27, 393.
Comyn, J. (1981) Developments in Adhesives-2, Kinloch, A. J., Ed., Appl.
Sei. Publ., London.
Cornell, R. W. (1953) J. Appl. Mech., Trans. ASME 75, 355.
Cotter, J. L. (1977) The durability of structural adhesives, Developments in
Adhesives-l, Wake, W. C. Ed., Appl. Sei. Publ., London, Chapter 1.
Counsell, P. J. C. (1979) Communication from Evode, Ltd, Stafford.
Crocombe, A. D. and Adams, R. D. (1981a) J. Adhesion 12(2), 127.
Crocombe, A. D. and Adams, R. D. (1981b) J. Adhesion 13(2), 141.
Crocombe, A. D. and Adams, R. D. (1982) J. Adhesion 13(3/4), 241.
Dannenberg, H. and May, C. A. (1969) Epoxide adhesives, Treatise on
Adhesion and Adhesives, Vol. 11, Patrick, R. L., Ed., Dekker, New York,
Chapter 2.
Demarkles, L. R. (1955) Investigation of the use of a rubber analog in the
study of stress distribution in riveted and cemented joints, Tech. Note 3413,
Nat. Advisory Cttee Aeronautics, Washington, D.C.
Diggwa, A. D. S. (1974) Polymer 15, 101.
Eichhorn, F., Hahn, O. and Fuchs, K. (1973) Adhäsion 17(12), 417.
Eichhorn, F., Hahn, O. and Stepanski, H. (1979) Schweissen Schneiden 31(1),
23.
ESDU 79016 (1979) Inelastic shear stresses and strains in the adhesives
bonding lap joints loaded in tension or shear, Engineering Sciences Data
Unit.
Eyring, H. (1936) J. ehem. Phys. 4, 283.
Fa1cone, A. S. and Miller, J. E. (1977) Durability of adhesive bonded struc-
tures, Bodnar, M. J., Ed., Appl. Polym. Symp. No. 3, Wiley, New York, p.
247.
Fields, D. (1973) Adhesives Age 16(9), 41.
Filon, L. N. G. (1902) Phi!. Trans. Royal Soc., Series A. 198, 147.
Foulkes, H., Shields, J. T. and Wake, W. C. (1970) J. Adhesion 2(4), 254.
Franzblau, M. C. and Rutherford, J. L. (1967), Study of micromechanical
properties of adhesive bonded joints, 1st Quarterly Progress Report, Con-
tract No. DAAA 21-67-C-0500, Picatinny Arsenal, Picatinny, USA.
Garnish, E. W. (1977) Advances in epoxy adhesive technology, Developments
in Adhesives-l, Wake, W. C., Ed., Appl. Sei. Publ., London, Chapter 1.
Garnish, E. W. (1982) Some applications of structural adhesives, Adhesion 6,
Allen, K. W., Ed., Appl. Sei. Publ. London, p. 173.
Gent, A. N. and Meinecke, E. A. (1970) Poly. Engng. Sei. 10,48.
Goland, M. and Reissner, E. (1944) J. Appl. Mech., Trans. ASME 66, A17.
Grant, P. (1976) Strength and stress analysis of bonded joints, British Aircraft
Corp. Ltd, Rep. 50R(P) 109.
Grant, P. (1978) Analysis of adhesive stresses in bonded joints, Symposium:
Jointing in Fibre Reinforced Plastics, Imperial College, London, I.P.c. Sci-
ence and Technology Press Ltd, p. 41.
286 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Gray, V. R. (1967) Adhesives in the timber trade, Aspects of Adhesion-3,


Alner, D. J., Ed., Univ. Lond. Press, London, p. 71.
Hahn, K. F. (1960) Photo stress investigation of bonded lap joints, Part II:
Analysis of experimental data, McDonnell-Douglas Co. Research Report
SM 4000-1, Long Beach, California.
Hahn, K. F. (1961) Adhesives Age 4(12), 34.
Hamilton, S. B. (1958) Building and civil engineering construction, A History
of Technology, Vol. IV Singer, c., Holmyard, E. J., Hall, A. R. and
Williams, T. 1, Eds, aarendon, Oxford, Chapter 15.
Harris, J. A. and Adams, R. D. (1982) Int. Symp. on Adhesive Joints, Kansas
City, Proc. Org. Coatings Appl. Polym. Sei., 184 Nat. Meeting, Amer. Chem.
Soc., p. 243.
Harris, J. A. and Adams, R. D. (1983) Adhesion 8, Allen, K. w., Ed., Appl.
Sei. Publ., London, p. l.
Harrison, N. L. and Harrison, W. J. (1972) J. Adhesion 3, 195.
Hart-Srnith, L. J. (1972) Design and analysis of adhesive bonded joints,
McDonnell-Douglas Co. Report No. 6059A, Long Beach, California.
Hart-Srnith, L. J. (1973a) Adhesive-bonded single lap joints, CR-112236,
NASA, Langley Res. Centre.
Hart-Srnith, L. J. (1973b) Adhesive-bonded scarf and stepped-Iap joints,
CR-112237, NASA, Langley Res. Centre.
Hart-Srnith, L. J. (1973c) Adhesive-bonded double-Iap joints, CR-112235,
NASA, Langley Res. Centre.
Hart-Srnith, L. J. (1978a) J. Eng. Mat. Technol., Trans. ASME 100, 16.
Hart-Srnith, L. J. (1978b) Adhesive-bonded joints for composites-
phenomeno-logical considerations, McDonnell-Douglas Co. Paper 6707,
Technology Confs. Assoc. Conf. on Advanced Composites Technology, EI
Segundo, California.
Hart-Srnith, L. J. (1980a) Further developments in the design and analysis of
adhesive-bonded structural joints, McDonnell-Douglas Co. Paper 6922,
ASTM Symp. on Joining of Composites, Minneapolis, Minnesota.
Hart-Srnith, L. J. (1980b) Structural details of adhesive-bonded joints for
pressurized aircraft fuselages, McDonnell-Douglas Co. Report MOC-J8858,
Long Beach, California.
Hart-Smith, L. J. (1981) Developments in Adhesives-2, Kinloch, A. J., Ed.,
Appl. Sci. Publ., London, p. l.
Hergenrother, P. M., and Levine, H. H. (1970). J. Appl. Polym. Sci. 14, 1037.
L'Herrnite, R. and Bresson, J. (1971) Synthetic Resins in Construction, Rilem
Coll., :Edition Peyrolles, Paris.
Hewlett, P. C. and Shaw, J. D. N. (1977) Structural adhesives used in eivil
engineering, Developments in Adhesives-l, Wake, W. c., Ed., Appl. Sei.
Publ., London, p. 25.
Hockney, M. G. D. (1970) Effect of outdoor exposure on stressed and
unstressed adhesive bonded metal-to-metal joints, Trial 1, Part 1, Two year
summary, Royal Aircraft Establishment Technical Report 70081, Farn-
borough, UK.
Hockney, M. G. D. (1972) Trial 2, Two year summary, Royal Aircraft
Establishment Technical Report 72100, Farnborough UK.
REFERENCES 287

Hockney, M. G. D. (1973) Trial 1, Part 2, Four year summary, Royal Aircraft


Establishment Technical Report 73016; and Trial 2, Part 2, Four year
summary, Royal Aircraft Establishment Technical Memo, Mat 175, Farn-
borough, UK.
Hole, L. G., and Abbott, S. G. (1973) The chemical stability of polyurethanes
in artificial leather, Artificial Leathers: Their Manufaeture Properties and
Uses, Hole, L. G., Ed., Shoe and Allied Trades Research Association,
Kettering, UK, p. 195.
Huggett, L. P. (1971) The toxicity of solvent vapours, Report IP 106, Shoe and
Allied Trades Research Association, Kettering, UK.
Humpidge, R. T. and Taylor, B. J. (1967) J. Sei. Inst. 44, 457.
Inglis, C. E. (1923) Proc. Royal Soe. AI03, 598.
Jennings, C. W. (1972) J. Adhesion 4(1), 25.
Kaelble, D. H. (1959) Trans. Soe. Rheology 3, 16l.
Kae1ble, D. H. (1960) Trans. Soe. Rheology 4, 45.
Keimei, F. A. (1966) Structural adhesives bonding, Bodnar, M. J., Ed., Appl.
Polym. Symp. No. 3, Wiley, New York, p. 27.
Kinloch, A. J. (1980) Metal Science 14, 305.
Kinloch, A. J. and Shaw, S. J. (1981) Fracture mechanics approach to joint
failure, Developments in Adhesives-2, Kinloch, A. J., Ed., Appl. Sei. Publ.
London, p. 83.
Kinloch, A. J., Dukes, W. A. and Gledhill, R. A. (1976) Durability of adhesive
joints, Adhesion Seienee and Technology, Part B, Lee, L-H., Ed., Plenum
Press, New York, p. 597.
Kizer, J. A. and Grosko, J. J. (1972) Development of the weidbond joining
process for aircraft structures, Processing for Adhesive Bonded Structures,
Bodnar, M. J., Ed., Wiley, New York, p. 353.
Knight, R. A. G. (1959) Wood 24(8), 330.
Krieger, R. B. (1975) Stiffness characteristics of structural adhesives for stress
analysis in hostile environment, American Cyanamid Company, Havre de
Grace, Maryland.
Krieger, R. B. (1980) Fatigue testing of structural adhesives, Adhesion 4,
Allen, K. W., Ed., Appl. Sci. Publ., London, p. 55.
Krueger, G. P. (1962) Mat. Res. Standards 2(6), 479.
Kuenzi, E. W. and Stevens, G. H. (1963) Determinationof mechanical
properties of adhesives for use in the design of bonded joints, US Forest
Products Service Research Note FPL-Oll, US Dept of Agric., Madison,
Wisconsin.
Kukovyakin, V. M. and Skoryi, I. A. (1972) Russ. Engng J. 52,40.
Kutscha, D. (1964) Mechanics of adhesive bonded lap type joints-survey and
review, US Forest Products Lab. Report MLTDR-64-298, US Dept of
Agric., Madison, Wisconsin.
Kutscha, D. and Hofer, K. E. (1969) Feasibility of joining advanced compo-
site flight vehic1e structures, AD 690616, ITT Research Inst.
Laidlaw, R. A. and Paxton, B. H. (1974) The effect of moisture content and
wood preservatives on the assembly gluing of timber, Building Research
Establishment Current Paper CP54j74, Princes Risborough Laboratory,
Aylesbury, Bucks.
288 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Lees, W. A. (1980) Modified epoxides-practical aspects of 'toughening',


Adhesion and Adhesives: Seienee, Teehnology and Applieations, Plastics and
Rubber Inst. Intern. Conf. Durham, Paper 17.
Lees, W. A. (1981) J. Adhesion 12(3), 233.
Lewis, A. F. and Ramsey, W. B. (1966) Adhesives Age 9, 20.
Lewis, A. F., Kinmonth, R. A. and Kreahling, R. P. (1972) J. Adhesion 3(3),
249.
Lindsey, G. H. (1966) Hydrostatic tensile fracture of a polyurethane elas-
tomer, Report ARL 66-0029, Wright-Patterson Air Force Base, Dayton,
Ohio.
Lipatov, Yu. S., Veselovski, R. A. and Znachkov, Yu. K. (1979) J. Adhesion
10(2), 157.
Lloyd, E. A. and Brown, A. F. (1978) Adhesion 2, Allen, K. w., Ed., Appl.
Sei. Pub!., London, p. 133.
Lubkin, J. L. (1957) J. Appl. Meeh. 79, 255.
Lubkin, J. L. and Reissner, E. (1956) Trans ASME 78, 1213.
McAbee, E., Tanner, W. C. and Levi, D. W. (1970) J. Adhesion 2(2), 106.
McCarvill, W. T., and Bell, J. P. (1974) J. Adhesion 6(3), 185.
Macdonald, M. D. and Calder, A. J. J. (1982) Int. J. Adhesion and Adhesives
2, 119.
McIntyre, R. T. (1972) Processing for adhesives bonded structures, Bodnar, M.
J., Ed., Appl. Polym. Symp. 19, Wiley, New York, p. 309.
McLaren, A. S. and MacInnes, 1. (1958) Brit. J. Appl. Phys. 9, 72.
Maisey, J. and Wake, W. C. (1970) Unpublished results, Rubber and Plastics
Research Association.
Martin, J. T. (1967) Adhesion and Adhesives 11, 2nd ed., Houwink, R. and
Salomon, G., Eds, Elsevier, Amsterdam, p. 87.
Matthews, F. L., Kilty, P. F. and Godwin, E. W. (1982) Composites 13, 29.
Matting, A. and Draugelates, U. (1968) Adhäsion 12(1), 5; 12(3), 110; 12(3),
132; 12(4), 161.
Matting, A. and Ulmer, K. (1963) Kaut. Gummi, 16(4), 213; 16(5), 280;
16(6), 334; 16(7), 387.
Matz, D. J., Guldemond, W. G. and Cooper, S. L. (1972) J. Polym. Sei.,
Polym. Phys. 10, 1917.
Mayfield, B., Bates, M. W. and Snell, C. (1978) Civil Engng (London), March,
p.63.
Mayo, A. P. (1975) An investigation of the collapse of a swimming pool roof
constructed with plywood box beams, Building Research Establishment
Current Paper CP 44/75, Princes Risborough Laboratory, Aylesbury, Bucks.
Merriman, H. R. (1965) Aircraft Engineering 37(5), 155.
Miner, M. A. (1945) J. Appl. Meeh. 12(3), 159.
Minford, J. D. (1982) Int. J. Adhesion Adhesives 2(1), 25.
Moloney, A. C., Brewis, D. M., Comyn, J. and Cope, B. C. (1981) The effect
of carriers on the environmental stability of adhesive joints, Adhesion 5,
Allen, K. W., Ed., Appl. Sei. Publ., London, p. 133.
Mostovoy, S. and Ripling, E. J. (1966) 1. Appl. Polym. Sei. 10, 1351.
Mostovoy, S. and Ripling, E. J. (1969) J. Appl. Polym. Sei. 13, 1083.
Muller, G. (1959) Der Verformungs und Bruchvorgang an Metallklebever-
REFERENCES 289

bindungen verschiedener Werkstoffe bei ein-und mehrachsiger statischer


Belastung, Dissertation, Technical University, Berlin.
Mylonas, C. (1954) Proc. Soc. Experimental Stress Analysis 12, 129.
Nadai, A. (1931) Plasticity: A mechanics oi the plastic state oi matter.
McGraw-Hill, New York.
Octocka, E. P. and Kwei, T. K. (1968) Macromolecules 1,244.
Panek, J. R. (1977) Polysulfide sealants and adhesives, Handbook oi Adhe-
sives, 2nd ed., Skeist, l., Ed., van Nostrand-Reinhold, New York, Chapter
22.
Patrick, R. L., Brown, J. A., Verhoeven, L. E., Ripling, E. J. and Mostovoy, S.
(1969) J. Adhesion 1(2), 136.
Peppiatt, N. A. (1974) Stress analysis of adhesive joints, Thesis, University of
Bristol.
Phillips, D. C., Scott, J. M. and Jones, M. (1978) J. Mat. Sei. 13, 31l.
Pick, M. and Wronski, A. A. (1976) The effect of triaxial loading on the
yielding and fracture properties of epoxy resins, Conf. on deformation, yield
and fracture of polymers, Cambridge, Plastics Rubber lnst., London.
Pickett, G. (1964) Trans. ASME 66, A-176.
Powis, C. N. (1973) Adhesive Age 16(9), 44.
Pye, C. J. and Adams, R. D. (1981) NDT International 14, 111.
Raghava, R. S., Cadell, R. M. and Yeh, G. S. Y. (1973) J. Mat. Sci. 8, 225.
Raithby, K. D. (1980) External strengthening of concrete bridges with bonded
steel plates, Transport and Road Research Laboratory Supplementary Re-
port 612, Crowthorne, UK. See also Int. J. Adhesion Adhesives (1982),2(2),
115.
Rayner, C. A. A. (1965) Adhesion and Adhesives I, 2nd ed., Houwink, R. and
Salomon, G., Eds, Elsevier, Amsterdam, p. 186.
Reinhart, T. J. and Hidde, R. (1966) Structural adhesives bonding, Bodnar, M.
J., Ed., Appl. Polym. Sci. Symp. No. 33 Wiley, New York, p. 299.
Renton, w. J. (1976) Experimental Mechanics 33,409.
Renton, W. J. and Vinson, J. R. (1975) J. Adhesion 7(3), 175.
Robertson, R. E. (1963) J. Appl. Polym. Sci. 7, 443.
Roff, W. J. and Scott, J. R. (1971) Fibres Films Plastics and Rubbers, Butter-
worths, London, p. 307.
Romanko, J. (1979) AGARD Lecture Series 102, 4-l.
Rotem, A. and Hashin, Z. (1975) J. Composite Matls 9, 19l.
Sazhin, A. M. (1964) Russ. Engng J. 44, 45.
Schliekelmann, R. J. (1979) AGARD Lecture Series 102, 8-l.
Schrader, M. E. and Cardamone, J. A. (1978) J. Adhesion 9(4), 305.
Schwartz, H. S. (1977) Durability of bonded structures, Bodnar, M. J., Ed.,
Appl. Polym. Symp. 32, Wiley, New York, p. 65.
Segal, E., and Rose, J. L. (1980) Research Techniques in Nondestructive
Testing, IV, Academic Press, London, p. 275.
Semerdjiev, S., (1970) Metal-to-Metal Adhesive Bonding, Business Books,
London.
Seymour, M. (1980) Concrete, June, 39.
Shanahan, M. E. R. (1974) The Creep Behaviour oi Structural Joints, Thesis,
The City University, London.
290 STRUCTURAL ADHESIVE JOINTS IN ENGINEERING

Shannon, R. W. and Thrall, E. W. (1977) Durability of adhesive bonded


struetures, Bodnar, M. J., Ed., Appl. Polym. Symp. 32., Wiley, New York, p.
131.
Shen, H. K. and Rutherford, J. L. (1972) Mat. Sei. Eng. 9, 323.
Shields, J. (1967) Unpublished results, Scientifie Instrument Research Associ-
ation.
Shields, J. (1976a) Adhesives Handbook, 2nd ed., Newnes-Butterworths,
London.
Shields, J. (1976b) Adhesives Handbook, 2nd ed., Newnes-Butterworths,
London, p. 295.
Shields, J. (1976e) Adhesives Handbook, 2nd ed., Newnes-Butterworths, Lon-
don, p. 294.
Shields, J. (1976d) Adhesives Handbook, 2nd ed., Newnes-Butterworths, Lon-
don, p. 283.
Shields, J. (1977) Evaluation of an ultra-violet-euring struetural adhesive,
Adhesion 1, Allen, K. W., Ed., Appl. Sei. Publ., London, Chapter 11.
Sims, M. G. (1972) Bibliography on Low and Cryogenic Temperature Properties
of Adhesives, 1965-72, Library Bibliography No. 332, Royal Aireraft
Establishment, Farnborough, UK.
Sneddon, I. N. (1961) The distribution of stress in adhesive joints, Adhesion,
Eley, D. D., Ed., Clarendon, Oxford.
Soden, A. L. and Wake, W. C. (1951) Trans. Inst. Rubber Ind. 27, 223.
Späth, W. (1965) Konstruktion 17(5), 170.
Stone, M. H. (1980) Strengths of metal-earbon fibre eomposite adhesives
joints at low temperatures, Adhesion and Adhesives: Seience, Technology
and Applications, Plastics and Rubber Inst. Conf. Durharn, Paper 10.
Sultan, J. N. and MeGarry, F. J. (1973) Polym. Eng. Sei. 13, 29.
Terekhova, L. P. and Skoryi, I. A. (1973) (Russian) Strength of Materials 4,
1271.
Thrall, E. W. (1980) The primary adhesively bonded strueture teehnology
(PABST) prograrn, Adhesion 4, Allen, K. W., Ed., Appl. Sei. Publ., London,
p. 1.
Thamm, F. (1976) J. Adhesion 7, 301.
Titow, W. V. (1978) Solvent welding of plastics, Adhesion 2, Allen, K. W.,
Ed., Appl. Sei. Publ., London, p. 181.
Turner, S. (1973) Creep in glassy polymers, The Physics of Glassy Polymers,
Haward, R. N., Ed., Appl. Sei. Publ., London, p. 223.
Vavilov, V. P. and Taylor, R. (1982) Research Techniques in Non-destructive
Testing 5, Aeademie Press, London, 239.
Vick, C. B. (1973) Forest Products J. 23(11), 33.
Volkersen, O. (1938) Luftfahrtforschung 15, 41.
Volkersen, O. (1965) Construction Metallique 4, 3.
Wah, T. (1976) Int. J. Mech. Sei. 18, 223.
Wake, W. C. (1970) Unpublished Report to Direetorate of Materials (Aviation
Group) UK Ministry of Teehnology, quoted by permission.
Wake, W. C. (1982) Adhesion and the Formulation of Adhesives, 2nd ed.,
Appl. Sei. Publ., London.
Wang, S. S., MandelI, J. F., Christensen, T. H. and MeGarry, F. J. (1976)
REFERENCES 291

Analysis of lap shear adhesive joints with and without short edge cracks,
Mass. lnst. Tech. Res. Report R76-2.
Wang, T. T., Ryan, F. W. and Schonhorn, H. (1972) J. Appl. Polym. Sei. 16,
1901.
Wargotz, B. (1969) J. Adhesion 1(4), 282.
Webber, J. P. H. (1981) J. Adhesion 12, 257.
Westland Helicopters, Ltd (1974) Materials Laboratory Research Paper No.
469, Yeovil, England.
Westland Helicopters, Ltd (1975) Addendum to Research Paper No. 469,
Yeovil, England.
Whitney, W. and Andrews, R. D. (1967) J. Polym. Sei. C16, 298l.
Wooley, G. R. and Carver, D. R. (1971) J. Aireraft 8, 817.
Wright, M. D. (1978) Composites 9, 259.
Wright, M. D. (1980) Composites 11, 46.
Zienkiewicz, O. C. (1971) The Finite Element Method in Engineering Seienee,
McGraw-Hill, New York.
Appendix

Standard American and UK Specifications


for Adhesion Tests

We present here for reference a list of standard American and British


specifications relating to bonded joints to be found in ASTM and BSI
publications. It should be appreciated that there are other American
and British (as weH as German, French, Japanese and so on) organiza-
tions which have specified tests to be carried out on adhesive joints. To
cover aH of these would be impossible and is, fortunately, unnecessary
since there are sufficient tests listed here to cover almost all even-
tualities. The sketches accompanying each section provide the reader
with a quick identification.
For the ASTM specifications, abbreviations used are:
STM = Standard test method for .. .
RP = Recommended practice for .. .
SRP = Standard recommended practice for ...
SP = Standard practice for ...
For the materials: M = metal, P = plastic, W = wood, G = glass, RP =
reinforced plastics

293
Parameterlschematic Title Materials ASTM BS

Definitions Standard definitions of terms relating to adhesives All D 907-82

Static tension lap shear STM strength properties of adhesives in shear by M D 1002-72
strength tension loading (metal-to-metal) (1978)
Bond strength in longitudinal shear MP 5350:C5
(1976)
STM strength properties of adhesives in shear by M D2295-72
tension loading at elevated temperatures (metal- (1978)
N to-metal)
ID
.j:>. STM strength properties of adhesives in shear by M D2557-72
tension loading in the temperature range of (1978)
-267-8 to -5SOC (-450 to -67°P)
SRP determining the strength of adhesively-bonded P D 3163-73
rigid plastic lap-shear joints in shear by tension (1979)
loading
SRP determining the strength of adhesively bonded P D3164-73
plastic lap-shear sandwich joints in shear by ten- (1979)
sion loading

Static tension lap-shear SP measuring strength and shear modulus of non- MW D3983-81
strength and modulus rigid adhesives by the thick adherend tensile lap
(thick adherend) specimen

1----
- L___ L1
Static compression lap STM strength properties of adhesive bonds in shear W D 905-49
shear strength by compression Ioading (1981)

-+8-
Static tension double- STM strength properties of double-Iap shear adhe- M D 3528-76
lap shear strength sive joints by tension loading (1981)
Bond strength in longitudinal shear MP 5350:C5
1 -..... (1976)
:::
....-.

Static tension lap shear STM strength properties of adhesives in two-ply W D2339-70
strength wood construction in shear by tension loading (1976)
IV
10 !s==J - + STM strength properties of adhesives in shear by M D 3165-73
Vl ---E:sl
tension loading of laminated assemblies (1979)

Static shear STM shear strength and shear modulus of structu- M E229-70
ral adhesives (napkin ring) (1981)

{[}
Tensile butt STM tensile properties of adhesive bonds (:n; test) M,W D 897-78

STM tensile strength of adhesives by means of bar M D 2095-72


.-L_ l- and rod specimens (1978)
Determination of bond strength in direct tension M 5350:C3
(1978)
Determination of bond strength in direct tension in foam plastic, 5350:C6
sandwich panels honeycomb (1981)
Appendix-contd.

Parameterlschematic Title Materials ASTM BS

Peel STM peel or stripping strength of adhesive bonds M, P, W others D 903-49


0
(180 peel) (1978)
0
180 pe el test for a flexible-to-rigid assembly M,P 5350:C11
(1979)
STM cross-lap specimens for tensile properties of G,M,W D 1344-78
N 00 adhesives
\Cl
0- STM climbing drum pe el test for adhesives M, honeycomb D 1781-76
(1981)
© Climbing drum peel test M, honeycomb 5350:C13
t (1980)
STM peel resistance of adhesives (T-peel test) M or any flexi- D 1876-72
ble material (1978)
0
180 T-peel test for flexible-to-flexible assembly M or any flexi- 5350: C12
~ ble material (1979)
+ SP determining durability of adhesive joints stres- All D 2918-71
sed in peel (1981)
STM floating roller pe el resistance of adhesives M D3167-76
(1981)
Floating roller pe el test M or any flexi- 5350:C9
ble material (1978)
0
90 peel test for a flexible-to-rigid assembly Any 5350:ClO
~ (1979)
~
90° peel test for a rigid-to-rigid assembly M 5350: C14
(1979)
STM adhesive bonded surface durability of alumi- M D 3762-79
-,,~ nium (wedge test)
t STM strength properties of adhesives in cleavage P D 3807-79
peel by tension loading (engineering plastics to
engineering plastics)

Cleavage SP fracture strength in c1eavage of adhesives in M D 3433-75


bonded joints (1980)
Strength properties of adhesives in c1eavage peel by Various engin- D 3807-79
:1 ==:J tension loading (engineering plastics-to- eering plastics
:[ engineering plastics) materials
STM c1eavage strength of metaHo-metal adhesive M D 1062-78
bonds
N
:s
~
Creep RP conducting creep tests of metal-to-metal adhe- M D 1780-72
sives (1978)
STM creep properties of adhesives in shear by M D 2294-69
tension loading (metal-to-metal) (1980)
Creep and resistance to sustained application of M,P 5350:C7
force (1976)
STM creep properties of adhesives in shear by M D 2293-69
~: ;1"; :!;; ~ compression loading (metal-to-metal) (1980)

Fatigue STM fatigue properties of adhesives in shear by M D 3166-73


..... ' tension loading (metall metal) (1979)
--.
Appendix-contd.

Parameterlschematic Title Materials ASTM BS

Impact STM impact strength of adhesive bonds M,W D 950-78

IV
'Ci
00
E:J
Environment STM resistance of adhesive bonds to chemical re- All D89~66
agents (1979)
STM effect of moisture and temperature on adhe- All D 1151-72
sive bonds (1979)
STM resistance of adhesives to cyclic laboratory All D 1183-70
aging conditions (1981)
SP atmospheric exposure of adhesive bonded joints All D 1828-70
and structures (1981)
Author Index

Numbers in italic type indicate those pages on which reJerences are given in Juli.

Abbott, S. G., 174,287 Bijlmer, P. F. A., 138,284


Adams, R. D., 18, 24, 26, 28, 29, 30, Black, J. M., 174,284
31,34,35,36,37,38,39,40,41, Blomquist, R. F., 174,284
43,44,45,46,47,48,49,57,58, Bolger , J. c., 183, 279,284
60,61,62,63,64,70,71,72,74, Bossler, R. c., 93, 284
78,80,81,82,91,92,95,96,97, Botsco, R., 140,284
98,99,100,101,105,107,115,116, Bowden, P. B., 58, 163,284
117,118,123,124,126,127,129, Bowditch, M. R., 277,284
130,132,136,141,283,285,289 Bower, G. M., 258, 259,284
Allan, G. G., 179,283 Bresson, J., 274,286
Allen, K. W., 164,228,230,238,239, Brewis, D. M., 251,288
240,241,283 Brown, A. F., 140,288
Allman, D. J., 28, 46, 49, 50,283 Brown, J. A., 168,289
Alsalim, H. S., 228, 230,283 de Bruyne, N. A., 3, 93,284
Althof, W., 132,250,251,257,259, Bryant, R. W., 93,284
260,283,284 Buck, B. 1., 195, 196,258,284
Alwar, R. S., 92, 95,284 Burgman, H. A., 258, 259,284
Andrews, R. D., 161, 163,291 Butt, R. 1.,213,254,255,284
Aponyi, T. J., 195,284
Argyris, J. H., 8, 9,284
Cadell, R. M., 59, 102,289
Calder, A. J. J., 275,288
Barker, R. M., 70,284 Cardamone, J. A., 234,289
Barth, B. P., 170,284 Carver, D. R., 37, 38,291
Bates, M. W., 273,288 Chambers, S. H., 38, 72,283
Bates, R., 161,284 Chugg, W. A., 199,285
Bell, J. P., 93,288 Coker, E. C., 36,285
Benham, P. P., 166,284 Coleman, B., 164,285
Benson, N. K., 24, 28,284 Comyn, J., 250, 251,285,288
Benthem, J. P., 95, 98,284 Cooper,S. L., 164,288
299
300 AUTHORINDEX

Cope, B. c., 251,288 Grosko, J. J., 216,287


CoppendaIe, J., 18,37,44,58,60,61, Guldemond, W. G., 164,288
62,63,64,74,91,95,96,97,98,
99, 100, 101, 115, 116, 123, 124,
126,127,283 Hahn, K. F., 28, 239,286
Cornell, R. W., 23, 24,285 Hahn, 0.,217,253,254,285
Cotter, J. L., 11, 184,213,223,230, Hamilton, S. B., 2,286
234,242,244,254,255,261,263, Harris, J. A., 37, 63,118,132,286
267,269,284,285 Harrison, N. L., 37, 95,286
Counsell, P. J. c., 161,285 Harrison, W. J., 37, 95,286
Crocombe, A. D., 37, 46, 47, 48, 49, Hart-Smith, L. J., 52, 53, 54, 55, 56,
129, 130,285 57,58,69,103,286
Christensen, T. H., 38,290 Hashin, Z., 77,289
Hatt, F., 70,284
Hergenrother, P. M., 195,286
Dannenberg, H., 183, 184,285 L'Hermite, R., 274,286
DeI Strother, P. J. A., 38, 72,283 Hewlett, P. c., 2,286
Demarkles, L. R., 29, 31, 33, 73,285 Hidde, R., 258,289
Draugelates, V., 156, 157, 158,159, Hockney, M. G. D., 186, 195, 196,
160,243,244,246,248,249,288 234,242,244,258,261,263,266,
Dukes, W. A., 93, 168,234,284,287 267,268,269,270,284,286,287
Hofer, K. E., 25,287
Hole, L. G., 174,287
Eichhorn, F., 217, 253, 254,285 Huggett, L. P., 222,287
Eyring, H., 164, 165,285 Humpidge, R. T., 93,287

IngIis, C. E., 36, 39,287


Falcone, A. S., 278,285
Fields, D., 216,285
Filon, L. N. G., 95,285 Jennings, C. W., 124,287
Foulkes, H., 93, 163, 193,211,257, Jukes, J. A., 58,284
285
Franzblau, M. c., 93,121,284,285
Freeman, J. H., 258, 259,284 Kaelble, D. H., 129,287
Frost, L. W., 258, 259,284 Keimei, F. A., 111,281,287
Fuchs, K., 253, 254,285 Kilty, P. F., 74,288
Kinloch, A. J., 167, 168, 172,234,
287
Garnish, E. W., 181, 183,201,279, Kinmonth, R. A., 238, 248,288
280,285 Kizer, J. A., 216, 287
Gent, A. N., 95,285 Knight, R. A. G., 3, 212,287
Gledhill, R. A., 168, 234,287 Knox, A. G., 164,285
Godwin, E. W., 74,288 Kreahling, R. P., 238, 248,288
Goland, M., 23, 24, 25, 26, 27,28, Krieger, R. B., 120,287
29,30,37,46,51,52,54,120,285 Krueger, G. P., 161,200,287
Grant, P., 52,285 Kuenzi, E. W., 93, 94, 95, 99, 123,
Gray, V. R., 199,285,286 160,287
AUTHORINDEX 301

Kukovyakin, V. M., 92,287 Octocka, E. P., 159,289


Kutscha, D., 25, 29,287
Kwei, T. K., 159,289
Panek, J. R., 183,289
Patrick, R. L., 168,289
Laidlaw, R. A., 198,287 Paxton, B. H., 198,287
Lees, W. A., 187,288 Peppiatt, N. A., 24, 26, 28, 29,30,
Levi, D. W., 242,288 31, 34, 35, 36, 37, 38, 39, 40, 41,
Levine, H. H., 195,286 ~,M,~,%,~,~,OO,M,~,
Lewis, A. P., 124,238,248,288 ~,64,n,~,W,~,~,~,~,
Lindsey, G. H., 95,288 96, 97, 98, 99, 100, 107, 115,
Lipatov, Yu. S., 148,288 116, 126, 283, 289
Lloyd, E. A., 140,288 Pick, M., 58,289
Lubkin, J. L., 24, 69, 90, 92,288 Pickett, G., 95,289
Powis, C. N., 278,289
McAbee, E., 242,288 Pye, C. J., 141,289
McCammond, D., 166,284
McCarvill, W. T., 93,288
Macdonald, M. D., 275,288 Raghava, R. S., 59,102,289
McGarry, P. J., 38, 58,290 Raithby, K. D., 274, 275,289
MacInnes, I., 27,288 Ramsey, W. B., 124,288
McIntyre, R. T., 231,288 Rayner, C. A. A., 189, 289
McLaren, A. S., 27,288 Reinhart, T. J., 258,289
Maisey, J., 195,288 Reissner, E., 23, 24, 25, 26, 27, 28,
MandeH, J. P., 38,290 29,30,37,46,51,52,54,90,92,
Martin, J. T., 225,288 120,285, 288
Matthews, P. L., 74,288 Renton, W. J., 28,119,120,289
Matting, A., 156, 157, 158, 159, 160, Ripling, E. J., 168,288,289
243,244,245,246,248,249,288 Robertson, R. E., 164,289
Matz, D. J., 164,288 Roff, W. J., 170,289
May, C. A., 183, 184,285 Romanko, J., 132,289
Mayfield, B., 273,288 Rose, J. L., 142,289
Mayo, A. P., 201,288 Rotem, A., 77, 289
Meinecke, E. A., 95,285 Ruffing, C. R., 258, 259,284
Merriman, H. R., 211,288 Rutherford, J. L., 93,121,153,160,
Miller, J. E., 278,285 166,284, 285, 290
Minderhoud, P., 95, 98, 284 Ryan, P. W., 135, 136,291
Miner, M. A., 248,288
Minford, J. D., 261, 265,288
Moloney, A. c., 251,288 Sazhin, A. M., 32, 33,289
Mostovoy, S., 168,288,289 Schliekelmann, R. J., 142,289
MuHer, G., 121,288 Schonhorn, H., 135, 136,291
Mylonas, c., 37,289 Schrader, M. E., 234,289
Schwartz, H. S., 225,289
Scott, J. R., 170,289
Nadai, A., 128,289 Segal, E., 142,289
Nagaraja, Y. R., 92, 95,284 Semerdjiev, S., 2n, 281,289
Neogi, A. N., 179,283 Seymour, M., 281,289
302 AUTHORINDEX

Shanahan, M. E. R., 153, 155, 159, Ulmer, K., 245,288


164,238,239,240,241,283,289
Shannon, R. W., 280,290
Shaw, J. D. N., 2,286
Shaw, S. J., 167,287 Vavilov, V. P., 141,290
Shen, H. K., 153, 160, 166,290 Verhoeven, L. E., 168,289
Shields, J., 161, 181, 195,205,211, Veselovski, R. A., 148,288
213,221,222,258,290 Vinson, J. R., 28,289
Shields, J. T., 93, 163, 193,211,257, Volkersen, 0., 20, 21, 23, 28, 32, 33,
285 52,54,92,121,290
Sims, M. G., 211, 258,290
Skoryi, 1. A., 92,287, 290
Sneddon,1. N., 24, 25,290 Wah, T., 69,290
SneH, c., 273,288 Wake, W. c., 93,163,193,194,195,
Soden, A. L., 212,290 211,212,230,257,283,285,288,
Spath, W., 249,290 290
Stannard, K. J., 277,284 Wang,S. S.,38,290
Stepanski, H., 217,285 Wang, T. T., 135, 136,291
Stevens, G. H., 93, 94, 95, 99, 123, Wargotz, B., 159,291
160,287 Webber, J. P. H., 69,291
Stone, M. H., 202,290 Whitney, W., 161, 163,291
Sultan, J. N., 58,290 Wooley, G. R., 37, 38, 291
Wright, M. D., 70,291
Wronski, A. A., 58,289
Tanner, W. c., 242,288
Taylor, B. J., 93,287
Taylor, R., 141,290
Terekhova, L. P., 92,290 Yeh, G. S. Y., 59, 102,289
Thamm, F., 70,104,105,290
Thrall, E. W., 280,290
Titow, W. V., 205,290
Traynor, E. J., 258, 259,284 Zienkiewicz, O. c., 37, 46,291
Turner,S.,165,290 Znachkov, Yu. K., 148,288
Subject Index

Acoustic emission Aircraft-contd.


monitoring, 44 structures, 279
testing, 140-1 Aluminium, 223, 254, 257
Acoustic methods, 140 Amplitude effects, 246-9
Adherend Anaerobic adhesives, 10
bevelled, 69-72 Anchorages, 273-4
definition, 3 Anodizing treatments, 223
materials, 116 Applications, 271-82
metal influence on, 193 Assembly, 9,104,113-14
parameters, 17 ASTM D 897-78,125
preparation of, 12 ASTM D 903-49, 131
scarfed, 69-72 ASTM D 950-78,133
shape effect, 69-72 ASTM D 1002-64, 45
thin sheet, 116-21 ASTM D 1002-72, 116
Adhesion ASTM D 1780-72, 132
bonding, advantages and disadvan- ASTM D 1781-76, 131
tages of, 12 ASTM D 1876-72, 131
diffusion theory of, 178 ASTM D 2094-69, 125
testing, 293 ASTM D 2095-72,125
Adhesive ASTM D 2293-69, 132
definition, 3 ASTM D 2294-69, 132
failure, 42 ASTM D 2918-71,132
form, 186, 190, 200 ASTM D 2919-71,132
parameters, 17 ASTM D 3165-73, 118, 119
selection, 175-217 ASTM D 3166-73,132
strength, 134 ASTM D 3167-76, 129, 131
types, 176 ASTM D 3762-79, 131
Aerospace industry, 116 ASTM D 3933-80, 223
Air pockets, 104 ASTM D 3983, 120
Aircraft ASTM D 3983-81,119,125
fuselage, 280 ASTM E 229-70,126
industry, 8, 181,272 ASTM specifications, 293
303
304 SUBJECT INDEX

Axially loaded butt joint, 121-5 Copper, 203


Corner joints, 111
Corrosion, 11
inhibition, 232
Bending moment(s), 27 problems, 275
factor, 24
Cast analysis, 12
Bevelled adherend, 69-72 Coupling agents, 232-6
Boeing wedge test, 131 Crack(s)
Bonding, 112 cracking, and, 44, 106, 134, 275
pressure, 186, 188, 200 growth,249
temperature, 186, 188 rate, 131
time/temperature, 200 initiation, 249
Brass, 204, 230
length,168
Bridge propagation, 167-9
construction, 274-5 Cracking, 275
decking, 276 Crazing, 165, 171-2
Brittle fracture, 149, 163, 166-71
Creep, 8, 103, 164-6, 187,238-41
BSI publications, 293
rate, 240
Buckling, 111
Critical energy, 167
Butt joints, 5-6, 93-102, 112 Cross-linking, 195
adhesives, 10
C-scan,139
Cadmium, 204, 230 CTBN (carboxyHerminated buta-
Carriages, 276 diene-acrylonitrile), 67
Cars, 276 Curing, 145, 186,208
CFRP, 7, 74, 75, 77-85, 88, 89, agents, 184, 189,203
202, 220, 222, 272, 278 problems, 134
Chair frames, 277 Cyanoacrylate adhesives, 10
Chemical etching, 223-31 Cycles-to-failure, 243-56
Chromium, 230 Cyclic deformation, 157
trioxide, 224
Cleavage stresses, 102, 167, 185, 199
Climbing drum test, 130 Decomposition temperature, 149, 151
Closed-form analyses, 37, 57 Defect types, 134-7
Co-curing technique, 73 Deformation
Cohesive failure, 42, 100, 169 characteristics, 29, 120, 152, 185,
Cohesive strength, 134 187,200
Compatibility, 176, 177 rate, 247
Composite materials, 73-90 Degradation, 173-4
Compressive transverse stresses, 109 Delayed yielding, 164
Computer programs, 52, 251 Design
Concrete, 220 criterion, 11
Condensation, 145 features, 8-11
Construction roles, 102-14
materials, 6-8 De-smutting, 228, 229
technique, 8-11 Destructive testing, 116-34
Contact adhesives, 9-10 Deterioration resistance, 173-4
Continuum mechanics, 52 Differential shear, 20, 66
SUBJECT INDEX 305

Diffusion, 177, 178,250-2 Failure-contd.


bond,178 stress, 54
constant, 251 surfaces, 45, 163
rate, 250 without yielding, 166-71
theory of adhesion, 178 Fatigue, 103, 216
Dismantling, 8, 113 cracks, 280
Displacement by moisture, 252-4, strength,9
256 Fillets, 6
Distortions, 113 Finite-difference method, 32, 33
Double bonds, 145 Finite-element analysis, 18,37,45,
Double butt-strap joint, 118, 119 50,59,74,92,99,106,129
Double-lap joint, 16,27, 77-81, 118, Flexural stiffness, 275
119 Floating roller test, 129-31
shear joint, 5, 270 Fokker
Double-scarf joints, 87 Bond Tester (Mark II), 140
Doublers, 113 Contamination Tester, 138
DTD 910C, 224 Friendship, 272
DTD 915D, 225 Fracture
DTD 5577, 129 energy, 167-70
Ductile failure, 163 modes, 67
Ductility effects, 270 Frequency effects, 245-6
Durability, 201, 212 Furniture, 277
Dynamic modulus, 155, 157-9
Dynamie properties, 160
Dynamic tests, 132-4 Gap-filling properties, 208-10
Gas turbine blades, 148-9
Glass transition temperature, 149-50,
Edge shapes, 37 172, 177, 212, 238
Elastic limit, 124 Goland and Reissner
Elastic modulus, 17 analysis, 23-7
Elastie-plastic behaviour, 90 second approximation, 24-7
Elasto-plastie analysis, 51-69 Gold,204
End effects, 34-51 Grit blasting, 220-1
Endurance limit, 237, 248 Grooved joint ends, 105
Environment effect, 11,253,275 Grooving, 111
Epoxy-polyamide, 262 GFRP, 7, 220
ESDU/Grant method, 54, 56 GRP, 222, 234, 277, 278
ESDU programs, 52
Etch primers, 232
Etching. See Chemical etching Harmonie Bond Tester, 140
Exposure trails, 260-70 Helicopter( s), 278
blades, 278
High-temperature
Failure adhesives, 211-12, 258-9
criteria, 8, 63, 67,69 metal-metal adhesion, 193-6
loads, 63 Holographie interferometry, 141
modes, 75-7, 106, 164-5 Honeycomb structure, 136-7
pattern, 143,243 Hooke's law, 31
306 SUBJECT INDEX

Hot melt adhesives, 152, 159 Magnets, 279


Hot setting adhesives, 10 Manufacturing conditions, 186, 187,
Hovercraft, 278 200
Humidity effects, 252, 254, 255, 267 Material non-linearity, 34
Hydrogen bonding, 179 Mathematical analysis, 18-19
Hydrolysis effect, 173, 256 Maxwell model, 153
Mechanical properties, 121, 155
Mechanical test procedures, 115-42
Melting temperature, 149, 151-2
Immersion effect, 253 Metal-wood structures, 201-2
Impact Metals, 6-8,180-96,203,219,228,
strength, 185, 187, 199 230-1
test, 133 Mild steel, 149,228,275
In-service testing, 138 Milled joint ends, 105
Inspection, 104 Mixed adhesives, 148
Intermolecular forces, 144 Mixed constructions, 201-2
Internal stresses, 24 Modified epoxies, 263
Interpenetrating polymer networks, Mohr-Coulomb criterion, 162
148 Moisture
absorption, 250-2
content, 198, 212
Jigging, 9-11 diffusion, 250-2
Joint effects, 11, 249-56, 270
classification, 14 Motor car structures, 114
cQnfigurations, 2-6, 16
design, 2
efficiences, 85 Nadai correction, 128
Napkin ring test, 120, 125, 126
Nitrile-phenolic adhesives, 269
Noble metals, 204
KGR-l extensometer, 120 Non-destructive testing, 134-42
reviews, 142
Non-metallic materials, 204
Laminated assembly, 119 Non-structural adhesives, 175, 176
Laminated techniques, 73
Lamp posts, 279 Oil tolerance, 214
Lap joints, 19-34, 104-7 Optical holography, 141
Lap-shear Overloading, 238
joints, 5,169,238,239,247,248,
250,251,259
tests, 118 PABST programme, 52, 55, 56, 120,
Latex-formulated adhesives, 212 279-80
Lear Fan aircraft, 273 Paste adhesives, 191-3
Linear elastic analysis, 19-20, 62, 90 Peel
Linear elastic failure, 85 resistance, 186,270
Liquid adhesives, 191-3 stresses, 46-51, 55, 57, 102, 104,
Load-bearing joints, 3 109
Low temperature, 210-11 tests, 16, 129, 130, 131
SUBJECT INDEX 307

Phenolic-modified adhesives, 263 Semi-structural adhesives, 176,202-


Phosphoric anodizing treatment, 225 17
Photoelastic analysis, 28, 37 Service Irre, 8, 12, 237-70
Plane-strain, 167 as indicated by climatic exposure
Plane-stress, 167 trails, 260-70
Plastic(s) Setting rate, 208
deformation, 54, 61, 65, 66, 80, 89 Shear
materials, 205 displacements, 31
Poisson's ratio, 29, 94, 96, 102, 123, modulus, 115, 120, 126, 128, 160,
160,162 161,245
'Poker chip' test, 121, 122 strain, 128, 153
Polymer structures, 145-8 strength, 185, 199
Polymerie adhesives, 121, 143-74 stress, 12,22,23,28,29,40,42,
Polymerization, 145,208 46-51,63,79,81,84-8,93,
Polyvinyl formaI/phenolic adhesives, 96,97,117,119,126,128
269 tests, 125-8
Porosity, 134 Shot blasting, 220-1
Post-bonding testing, 138 Shrinkage stresses, 172
Post-curing, 190 Silcone rubber, 262
Pot Irre, 186, 190,200,214 Siloxanes, 233,234-6
Pre-bonding tests, 137-8 Single bonds, 145
Preservatives, 198 Single-Iap
Pressure-sensitive adhesives, 156 joint, 14,34-69,136
Primers, 186, 187,217,218,231-6, joint test piece, 117
263 joint tests, 118, 119, 129
Principal stresses, 39, 41, 46-51 shear joints, 247
Ski constructions, 281
Smutting, 228
Quality control, 134 Solvent
Quinton Interchange, 274 degreasing, 221
toxicity, 221
Spacers, 108
Radiography, 141 Specifications, 293
Reinforced plastics structures, 202 Spewfillet, 36, 38, 42, 44, 95-7, 99,100
Reissner's criterion, 30 Spring and dashpot model, 155
Relaxation time, 155 Square edge, 36
Resin concretes, 2 Stainless steel, 195, 196, 228,230,259
Retardation time, 155 Stepped adherends, 69-72
Riveting, 112, 216-17, 280 Stepped lap joints, 88
Rollers, 280 Stiffeners and stiffening effects,
112-13
Stiffness reduction effect, 105
Safe dilution volume (SDV), 222 Strain
Sand blasting, 220-1 behaviour, 165
Scarfed adherends, 69-72 energy, 24
Sealing and sealants, 1, 104 Strength
Segmental construction, 281 prediction, 16-17,34,62,67
Self diffusion, 177 profiles, 257
308 SUBJECT INDEX

Strength-contd. Thermal ageing, 259


properties, 143, 162,253-6 Thermal expansion coefficient, 172-3
requirements, 185, 186, 199 Thermal test methods, 141
retention properties, 262, 264 Thermoplastic materials, 175
Strength-temperature profiles, 194 Thick adherend test, 119-21
Stress(es), 14-114 Thin sheet adherends, 116-21
analysis, 38, 104 Thread locking, 10
concentrations, 45-51, 92, 95, 98, Threshold limit value (TLV), 221-2
102,106,216,239 Timber condition, 198-9
distribution, 13, 46, 55, 64, 66, 71, Time-to-failure, 241-3, 260-2, 269
72,75,76,82,83,93,97,98, Titanium and titanium alloys, 229-30
99, 120, 168,238 T-joints, 108-11, 135
effects, 12, 152, 238 'Top-hat' design, 112
gradients, 98 Torsion
intensity factor, 167 effects, 95, 97
prediction, 121 testing, 127
rupture,8,238,240 Torsional shear, 256-7
Stress-strain curve, 52, 59, 101,143, Torsional test pieces, 163
152, 160, 164 T-peel test, 130, 131
Structural adhesives, 3,11. 176, Transition temperature, 165
180-202, 205-8 Transverse cracking, 106
Structural materials, 144 Transverse direct stress, 35
Structural uses, 1 Transverse shear stresses, 33,35
Sub-assemblies, 8 Transverse strength, 74
Substrate interaction, 177-9 Transverse stresses, 109
Supported filmic adhesives (tapes), Tresca criterion, 162
190-1 Triethylenetetramine (TETA), 189
Surface Tri-Round lamp post, 279
inspection tests, 137 Tubular joints, 90-2, 107-8
preparation, 7,11,13,218-36,275
unbonds, 135 Ultrasonic(s) 138-40
wettability , 137 resonance testing, 140
Underground railways, 276
Unsaturation, 145
Tack,156 Unsupported films, 191
Tape adhesives, 190-1
Tapered joint ends, 105
Telephone kiosks, 281 van der Waals's forces 164, 178
Temperature Vapour degreasing, 230-1
effects, 148-74, 193-6, 245, 253, Versamids, 184
254,256-60,267,270 Vibration effects, 9
tolerance, 210-12 Viscoelasticity, 153-60
Tensile butt joint tests, 122 Voids, 135, 140, 166
Tensile shear strength, 136 Voigt model, 153,154
Tensile strength, 144 Volkersen's
Tensile stress, 77, 78, 99, 109, 185 second theory, 28
Tensile testing, 125, 155 shear lag analysis, 20-3, 52
Tension effects, 95 von Mises criterion, 162
SUBJECT INDEX 309

Water X-Y recorder, 139


exposure, 255
sensitivity, 212-14, 254 Yacht construction, 282
uptake determination, 251 Yield
Weather parameters, 261 criterion, 59, 100
Weld-bonding, 114,216 point, 116
Welding, 9, 112, 114 strength, 68
Width increase effect, 103 stress, 54, 162, 163
Wiping, 221 Young's modulus, 57,73,94,96,115,
Wähler curve, 244-5, 248 121, 123, 124, 143, 160-2,
Wood,2, 7,196-201,204,216, 167
219-20
Work to proportionality limit, 160
Zero-volume unbonds, 135
X-ray techniques, 141 Zinc,204

You might also like