Chemistry and Technology of Epoxy Resins PDF
Chemistry and Technology of Epoxy Resins PDF
Chemistry and Technology of Epoxy Resins PDF
of Epoxy Resins
Chemistry and Technology
of Epoxy Resins
Edited by
BRYAN ELLIS
Department of Engineering Materials
University of Sheffield
m
SPRINGER-SCIENCE+BUSINESS MEDIA, B.V.
First edition 1993
© Springer Science+Business Media Dordrecht 1993
Originally published by Chapman & Hali in 1993
Softcover reprint ofthe hardcover lst edition 1993
A catalogue record for this book is available from the British Library.
Library of Congress Cataloging-in-Publication Data available.
Preface
Epoxy resins have been commercially available for about 45 years and now
have many major industrial applications, especially where technical
advantages warrant their somewhat higher costs. The chemistry of these
resins is fascinating and has attracted study by many very able scientists. The
technological applications of the epoxy resins are very demanding and there
are many new developments each year.
The aims of the present book are to present in a compact form both
theoretical and practical information that will assist in the study, research
and innovations in the field of epoxy resin science and technology. The
literature on epoxy resins is so vast that it is not possible to be encyclopaedic
and that is not the function of the present text. It is the editor's hope that the
selection of topics discussed will provide an up-to-date survey. There is some
overlap in the chapters but this is minimal and so each chapter is essentially
self contained.
As with all chemicals there are toxicological and other hazards. These are
not dealt with in this text since a little knowledge can be dangerous, but
material supplied can provide information regarding any safety precautions
that may be necessary. However, often these precautions are not onerous
and epoxy resins, or more specifically the hardeners, can be handled readily.
It is hoped that this text will provide an up-to-date outline of the science
and technology of epoxy resins and stimulate further research into unsolved
problems and assist further technological developments.
Bryan Ellis
Acknowledgements
B.E.
Note
For ease of reference a general index and a separate index of curving agents and hardeners
are provided.
Contents
Index 327
Contributors
The term 'epoxy resin' is applied to both the prepolymers and to the cured
/0"
resins; the former contain reactive epoxy groups, R-CH-CH2' hence their
name. In the cured resins all of the reactive groups may have reacted, so that
although they no longer contain epoxy groups the cured resins are still called
epoxy resins. The relative size of the market for epoxy resins is indicated in
Table 1.1 from which it can be seen that they are important industrial
polymers. Since they are more highly priced than other resins, they will only
find application when they have technical advantages. Many of the
applications involve high added value products.
Although the first products that would now be called epoxy resins were
synthesized as early as 1891 (see Dearborn et al., 1953; Lee and Neville,
1967a) it was not until the independent work of Pierre Castan in Switzerland
and Sylvan Greenlee in the United States that commercial epoxy resins were
marketed in the 1940s, although similar resins had been patented in the
1930s. The earliest epoxy resins marketed were the reaction products of
bisphenol A and epichlorohydrin and this is still the major route for the
manufacture of most of the resins marketed today, although there are many
other types ofresin available (section 1.3).
Pierre Castan investigated potential resins which could be readily
moulded at low pressures for the replacement of vulcanite as a denture base
material. The BPA-epoxy resins could be cured by reaction of epoxy groups
with phthalic anhydride without the evolution of low molecular species and
hence did not require high moulding pressures. An alternative acrylic resin
is now used for denture base and the patents were licensed to CIBA in 1942.
Epoxy adhesives and casting resins were marketed in the USA in 1946.
Greenlee working for Devoe and Raynolds produced resins which were
similar to those of Castan but with a somewhat higher molecular weight with
the objective of developing superior surface coatings. The epoxy coatings
developed by Greenlee offered improved adhesion, hardness, inertness and
thermal resistance compared with alkyd or phenolic resins. Following the
Table 1.1 Epoxy resin sales in the USA, 1990-1991
1990 1991
Consumption in Japan in 1991 was 166 tonnes of epoxy resin and 386 tonnes of phenolic resin. t
• Relative to low density polyethylene = 1.0, Birley and Scott (1982).
t Modern Plastics International, 1992.
Me~ Cl
/ ° , I
HO-@-{\Q! OH + 2CH 2-CH-CH 2
Me Epichlorohydrin
Bisphenol A
(BPA) Alkali
MOH
Catalyst
BPA
° ~M~_ OH M~__ °
CH'2~CH-CH2
1 0~{-&-0-CH2-tH-CH2 0--@-{----&-0-cH2-cH~CH2 t
Me Me
II
4 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
0 340 0 170
1 624 1 312
2 908 2 454
10 3180 10 1590
first patent application (1948) Greenlee obtained about 40 patents for epoxy
resins. The major application of epoxy resins is still for surface coatings
which consumes about 50% of all epoxy resins produced. The relative use
pattern of epoxy resins is indicated in Table 1.3 and applications are
discussed in more detail in chapter 9.
Innovation in epoxy resin technology has involved the synthesis of epoxy
resins with specific characteristics and some of the more important are
discussed in section 1.3. Equally important has been the development of
hardeners which depend on their reactivity with epoxy groups. The more
important hardener systems are treated in detail in chapter 2. Much of the
chemistry of epoxy resins depends on the reactivity of the epoxy groups
which will be briefly outlined in the next section (1.2).
The cure of epoxy resins involves the formation of a rigid three-
dimensional network by reaction with hardeners which have more than two
·reactive functional groups, that is, functionality is f > 2. Often f ~ 4 for
common hardeners for BPA resins which often have an effective
functionality of two, but may be higher when the cure temperature is high
enough for the secondary hydroxy groups to react.
The cure of epoxy resins is complicated and it is useful to visualize the
process in several stages, which are illustrated in Figure 1.1, although except
Table 1.3 USA applications of epoxy resins (Modern Plastics International, 1992)
[1990] [1991]
1000tonnes % 1000tonnes %
Protective coatings 89 49 84 51
Electrical applications 25 14 22 13
Reinforced resins 14 7.5 13 8
Bonding and adhesives 13 7.5 12 7.25
Flooring 12 6.5 11 6.25
Tooling and casting 13 7.5 12 7.25
Other 15 8.3 12 7.25
Total 181 100 166 100
INTRODUCTION TO EPOXY RESINS 5
I
Viscous liquid
: rubbery .
Visco-elastic I visco-elastic
liquid solid glassy viscoelastic
Rheological
Newtonian fluid T = rj'Y solid
properties shear stress r
~~
rate of shear
power law fluid
r = myn
IJ i- const.
m and n ~ f (T" to)
-----------------------------T---------------------·
.... I
I
difunctional
~
Molecular epoxy ~ : formation
branched lof highly
structures pre polymer
crosslinked
.J-( molecules : incipient
network
hardener oligomers I 3D network
f. 4 I
I
average molecular I increasing
weak and can be easily disrupted. To produce a structural material, cure has
to continue until most of the sample is connected into the three-dimensional
n~lwork so that the sol fraction becomes small and for many cured products
it has to be essentially zero. Cure and gelation are discussed in more detail
in ~hapter 3.
As cure proceeds there are major changes in the properties of the epoxy
resins. Initially the resin-hardener mixture is fluid and finally an elastic solid
is produced. The glass transition temperature of the curing resin increases as
cure proceeds and these changes can be represented in a time-temperature
trahsition (TTT) diagram introduced by Gillham (1986). Figure 1.2 is a
simplified version which illustrates the dominant effect of the onset of vitri-
fication as Tg increases to the cure temperature Te. For cure temperatures
well above Tg , the rate of reaction between the epoxy and hardener reactive
groups is chemically kinetically controlled. When ll.T = Te - Tg becomes
small the curing reactions become diffusion controlled, and will eventually
becpme very slow and finally stop. For products for which it is necessary to
ensure complete reaction of all epoxy groups it is a normal practice to post-
cure the resins at an elevated temperature. For successful application of
Cure
temperature
Te
Figure 1.2 Simplified time-temperature-transition diagram (TIT). Tgo is the glass transition
temperature of the mixture of epoxy prepolymer-hardener-additive. For cure temperature
Tc < Tgo, the mixture is glass (1) and reaction of epoxy groups is inhibited. In glass (2) the
epoxy resin vitrifies before gelation. For glass (3) the glass transition temperature increases with
increased cross-linking of the network. Tg~ is the limiting glass transition temperature as the
concentration of epoxy group E -> O.
INTRODUCTION TO EPOXY RESINS 7
This method is general for the synthesis of epoxy compounds but ethylene
oxide is now manufactured by direct oxidation of ethylene with air or oxygen
and a silver catalyst. Some of the early history of the synthesis and chemistry
of epoxy compounds has been discussed by Malinovskii (1965) with especial
reference to early Russian work. The synthesis of epoxy rings has been
discussed in detail by Gritter (1967) and Lewars (1984) and epoxy resins by
Tanaka (1988).
There are many methods for the synthesis of epoxy rings (Rosowsky,
1964), which are classified in Table 1.4. Although not the only ones, the
most important routes for the manufacture of epoxy resins are reaction of a
halohydrin with hydroxyl compounds, and the oxidation of unsaturated
compounds with a peracid. The first method is similar to the original
synthesis of ethylene oxide by Wurtz and may be illustrated by the reaction
of epichlorohydrin with hydroxyl compounds, such as phenols or aliphatic
alcohols.
/ "° CI
I
R-OH + CHz-CH-CH 2
Catalyst
------->
OHCl
I I
RO-CHz-CH-CH2
Alkali
/0"
R-0-CHz-CH-CH2 + MCI + H 20
~ °
'c'-'c"- +
..-, °
CH 3-C?-
'OH
8 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
1. Oxidation of alkenes
2. From halohydrin
Hypohalous addition to alkene and then cyclodehydrohalogenation
X OH 0
HOX I I Alk I' .... / \ "
:C=C~ ~ :C-C~ M~~) "C-C.... + MX
@-CHO +
Cl
I
0
II
CH 2-C-OEt ---
Na 0 0 0
II /\
@ - CH-CH-C-OEt
X-@-o Bf
III
C-CH2
o Sr 0- Sr
MeO.... 1 I
o
Me-~-tH2 MeO.... / \
MeO-)
"C-CH2 - "C-CH 2 + Sr-
Me Me
e. Grignard reagent
OX
R3-~-t-Rl
R2
Compiled from: Gritter (1967); Lewars (1984); Malinovskii (1965) and Rosowsky (1964) who
give extensive lists of references to the very many routes available for the synthesis of epoxy
rings. Advances in Heterocyclic Chemistry should be consulted for more recent references.
INTRODUCTION TO EPOXY RESINS 9
This is Prileschaiev's reaction which was published in 1912 and has been
extensively discussed in several reviews (Rosowsky, 1964; Tanaka, 1988). A
possible mechanism for this reaction with epoxidation by perbenzoic acid is
I
CH
\I +
CH
I
I o
CH
I ........0
CH- + f-@
I H-O
oII
Ar-C1 1r +'/C=L,
,-,/
o'0/c=o
the formation of which has to be avoided to improve the yield of the epoxy
compound. The manufacture of epoxy resins from unsaturated compounds
by epoxidation with peracids is discussed in section 1.3.3.
The geometrical structure of the epoxy ring is planar, with bond angles
and lengths (Figure 1.3) determined from electron diffraction and
microwave spectroscopic measurements which have been discussed by
Lwowski (1984), Lewars (1984) and Peters (1967). The differences in bond
angle from dimethyl ether are considerable as can be seen from Figure 1.3,
and there must be considerable ring strain due to angular distortion from the
tetrahedral carbon angle of 109°. The dipole moment of simple ethers is 1.1
10 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
o o
(a) (b)
1.41 1\ 141pm
c c
c
""
""
H
e'
H
(c) (d)
F4
Figure 1.3 The structure of the epoxy ring. (a) Bond lengths and angles for ethylene oxide.
(b) C-O bond lengths and angles for dimethyl ether (the positions of the hydrogen atoms not
shown). (c) 'Bent' bonds in the strained epoxy ring. (d) Projection view of ethylene oxides.
to 1.3 D and that for ethylene oxide is 1.82 and 1.91 D in benzene solution
and the vapour phase respectively. The ionization potential of the oxygen
2p.n lone pair in ethylene oxide is 10.6 to 10.8 eV which is rather higher than
that of dimethyl (10.0 eV) and diethyl (9.5 eV) ethers which Peters (1967)
has compared with the ionization energies of other simple oxygen
compounds.
The electronic structure of three-membered rings poses difficult problems
o C
/" /" 3
since with a c e or C C bond angle of about 60° the 'normal' sp hybridiz-
ation with linear bonds between the ring atoms is impossible. The bonding in
cyclopropane has been discussed extensively and Halton (1991) in an
interesting review of ring strain in cyclic molecules considered the latest
evidence. The bonding in cyclopropane is abnormal with the interbond
angle compressed to about 60° which is required for ring formation with the
nuclei 'moving ahead' of the bonding electron density with the formation of
a 'bent' or so-called 'banana' bond as illustrated in Figure 1.3.
The geometry of the epoxy ring is similar to that of cyclopropane but
because of the electronegativity of the heteroatom the internal ring bond
angles and lengths are not equal (Figure 1.3). Parker and Isaacs (1959)
discussed the various structures that have been proposed for ethylene oxide.
It has been suggested that the carbon atoms are trigonally hybridized, that is
Sp2, and that one such orbital from each carbon atom overlaps with an
oxygen atomic orbital to form a molecular orbital which occupies the centre
of the ring (Figure 1.3). It is possible that the presence of the 'central' ring
INTRODUCTION TO EPOXY RESINS 11
orbital accounts for conjugation of the epoxy ring with other delocalized
electrons, which is shown by bathochromic shifts in UV (electronic) spectra
(Lewars, 1984) and NMR ring currents (Gritter, 1967). Of course such
conjugation does not prove that the electrons in the un substituted
compounds are de localized , and there has been dispute regarding the
possibility of ring currents in these compounds (Gritter, 1967). Although the
strain energies of cyclopropane and the epoxy ring are very similar, 27.43
and 27.28 kcaVmole respectively (Gritter, 1967), it may be that the bonding
is very different. For instance, from the NMR data compiled by Lwowski
(1984) the chemical shifts and coupling constants for the epoxy ring are
different from those for cyclopropane.
The many industrial applications of epoxy resins require the formation of
three-dimensional networks by reaction with suitable polyfunctiQnal
hardeners, which are discussed in detail in chapter 2. Many of these curing
reactions depend on the reactivity of the epoxy ring, which is very much
more reactive than the 'normal' non-cyclic ethers, R-OR', where Rand R'
are alkyl or aryl groups. In normal ethers the oxygen link is resistant to
attack by alkalis, ammonia or amines. Epoxy resins will react with some
aliphatic amines at room temperature; these amines may be used as curing
agents at ambient temperatures (see chapter 2). This increased reactivity of
cyclic ethers is due to the ring strain.
The chemistry of the epoxy ring has been reviewed comprehensively by
Parker and Isaacs (1959) and a more recent discussion is that of Lewars
(1984). The literature of heterocyclic chemistry (Katritsky and Weeds, 1966;
Katritsky and Jones, 1979; Belen'kii, 1988) including that of the epoxy ring
has been listed periodically; initially references to epoxy resins were listed
but not recently. However, these annotations are a useful source of
reference to information on the reactions of the epoxy group. The reactivity
of epoxy compounds is summarized in Table 1.5.
The reactions that are most important for the synthesis and cure of epoxy
resins involve either electrophilic attack on the oxygen atom or nucleophilic
attack on one of the ring carbon atoms. For the unsymmetrically substituted
/0"
epoxy compound R-CH-CHz, which occurs in most epoxy resins, several
factors determine ring opening reactions, such as, the nature of the reagent
or catalyst which may be either electrophilic or nucleophilic, the influence of
the substituent and the relative steric hindrance at the two carbon atoms.
With the general reagent HR', two possible products of ring opening may be
produced: 0 OH
/ " I
R-CH-CHz + HR' ~ R-CH-CHz-R'
Normal
OR t
I
~ R-CH-CHz-OH
Abnormal
12 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
OH OH
0.2% H,SOJ100·C) I I
R-CH-CH 2
C
Me-c~
ROO R
I II II I
I "
CI-CH-C-O-CH 2 -CH 2-C-O-CHCI
CI-CH-C-OH
R\ R2 =H: ammonia
R 1= H; R2 == alkyl, aryl: primary amine
R\ R 2 =alkyl, aryl: secondary amine
2. Electrophilic additions
a. Alkyl halides
~
-200"C
X==Br; R==Et
X=I; R=Me, Et, Pr
b. Isocyanates
INTRODUCTION TO EPOXY RESINS 13
c. Oxides of sulphur
S03
1,4 dioxane
3. Reduction
OH
I
/0, LIAIH
~~'
©
CH-CH, 100%
OH OH
I I
~-'"'
LiBH
~'" +
74% 26%
4. Oxidation
OH 0
HNO, ) I II
CI-CHz-CH-C-OH
l00'C
Compiled from: Gritter (1967); Lewars (1984); Malinovskii (1965) and Rosowsky (1964).
This is only a small selection of the very many reactions of epoxy groups that have been
reported. Base-catalysed rearrangements of epoxides are discussed by Yandovskii and Ershov
(1972), retention of configuration by Akhrem et al. (1968) and ring expansion by Grobov et al.
(1966).
oII OH 0
I II
+ HO-C-R' -> R-CH-CH2-O-C-R'
OH
I
+ HSR' -> R-CH-CHz-SR'
14 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
°
c-c
,\"
/".....
x-
~ [,I
C-C
0
0 .. -
" ]
,/:......
0-
~ -C-C-
I I
I
~ etc.
I
)(0- X
Acid catalysed addition involves proton attack on the ring oxygen atom,
, I~" \H+
"C C, ~ "C C, ~
'/~\+" x- [, ~
. ot
"c c,:
j H9
~ -C-C-
I
I
I
)(0- X
but unfortunately this process is not so efficient for higher olefines. The
methods that have been used for the synthesis of epoxy rings have been
discussed by Lwowski (1984) and Lewars (1984), the former in a general
review of the synthesis of small and large heterocyclic rings (Table 1.4). A
comprehensive review of the synthesis of epoxy compounds is that of
Tanaka (1988). The most important for the synthesis of epoxy resins are
INTRODUCTION TO EPOXY RESINS 15
/
°
"
R-OH + CHz-CH-CH2
aI I Catalyst
~ R-0-CHz-CH-CH2
OHO
I i I
MOH
Stoichiome~
/0"
R-0-CH 2-CH-CH2 + MCI
Cl CI OH
I I I
CH2=CH-CHO CI-CHz-CH-CHO -- CI-CH2-CH-CH2
16 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
I
(catalytic
concentration)
OHP
/0" NaOH I I
R-0-CHr CH-CH2 + NaCI ( Stoichiometric RO-CH2-CH-CH2
concentrations
acid gas (Faith et at., 1965b). The reaction conditions will depend on the
design of the production unit (Materials and Technology, 1972). The purity
of the product is high, >95% p,p'-isomer; the other isomers formed are o,p'
and 0,0'. For resin manufacture the p,p' isomer content should be at least
98%. The light yellow colour of some epoxy resins may be due to trace
impurities in the bisphenol A, such as iron, arsenic and highly coloured
organic compounds. Other names for bisphenol A are 4,4' -isopropylidene
diphenol and diphenylolpropane (DPP).
When a large excess of epichlorohydrin is reacted with bisphenol A with a
stoichiometric amount of sodium hydroxide at about 65°C the resin
produced contains about 50% diglycidyl ether of bisphenol A, DGEBA
INTRODUCTION TO EPOXY RESINS 17
Table 1.6 Epoxy resins derived from epichlorohydrin
1. Phenols
a. Many difunctional phenols have been investigated. BPA is the most important, others
are listed in Table 1.7
b. Monofunctional phenols: modifiers for epoxy resins
c. BP A resins: esterified with fatty acids
d. Halogen-substituted phenols
2. Alcohols
a. Multifunctional
/0.., /0"
l,4-butanediol~ CH2-CH-CH2-0-(CH2)4-0-CH2-CH-CH2
/0"
/0" yH2-O-CH,-CH-CH,
glycerol~ CH,-CH-CH,-O-yH /0"
CH,-O-CH,-CH-CH,
b. Monofunctional: Resin modifier
/0"
Butanol~ CH 3 -(CH,h-0-CH,-CH-CH 2
@t
o ° il
C-O-CH2-CH-CH~
rr-O-CH2-C~/.cH2
°
'\
° °
b. Long-chain acids ~ epoxy resin esters
c. Acrylic acid
BPA resins + CH,=<;"H ~ epoxy acrylates
C=O
I
OH
Acrylic acid
5. Nitrogen compounds °
a. Amines
aniline ~ @-N/ - CH,-CH~CH,
"\
CH2-CH-CH~
° -
'\
O-CH,-CH-CH, P"
p-amino phenol ~ ~ - -
/0, ~
CH2-CH-CH2-N-CH2-CH-CH~
/0,
18 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
b. Cyanuric acid
6. Miscellaneous
Many other compounds have been reacted with epichlorohydrin. This table is a guide to
some of the more important. Other chlorohydrins and halohydrins may also be used,
MeO
\ / \
e.g. methylepichlorohydrin CI-CH 2-C-CH 2
Me~
HO--@-{---&-OH
Me
1. Bisphenol A (BPA)
Me
HO-@-{%OH
Me
2. Bisphenol F
3. Tetrakis phenylolethane
4. Resorcinol
HO%OH
5. Methylolated phenol
CH 0H
2 e dCH,OH
HO-@-{~~H Me
"oj§>-I~o" HO -@- C~
0 {~OH
CF1
Br Br
.~.
~~
o
I
INTRODUCTION TO EPOXY RESINS 21
the production of higher molecular weight resins the chain extension process
is used (section 1.3.2.4). Batzer and Zahir (1975, 1977) have discussed in
detail the molecular weight distributions of epoxy resins prepared by
different processes (Ravindrath and Gandhi, 1979).
Although the commercially produced resins have a distribution of chain
lengths it is possible to obtain the DGEBA by molecular distillation and it
can also be crystallized. Pure DGEBA is a solid which melts at 43°C. Also in
liquid commercial resins that have been stored for prolonged periods some
crystallization of DGEBA will occur and the resin appears cloudy. Outlines
of the production process for the manufacture of a low molecular weight
resin are given by McAdams and Gannon (1986) and also Savla and Skeist
(1977c). Preparative methods are given by Sandler and Karo (1977).
Not all of the species present in commercial resins are diepoxides; side
reactions can occur when epichlorohydrin is reacted with bisphenol A.
These can become more important when higher molecular weight resins are
produced. The important side reactions are
i. Hydrolysis of epoxy groups
/0" OH\OH
NaOH 1 1
-CH 2-CH-CH2 ----> -CH2-CH-CH 2
H 20
a-Glycol formation
1
~0-CH2-CH-CH2
CI
+ CH
°
/ "
2-CH-CH2CI
1
~ ~0-CH2-CH-CH2
CI
1
OH
1
°
1
CH 2
1
/
CH=OH
CI CH 2 CI
1
~ 0-CH2-CH-CH 2
1
° 1
CH 2
1
CH...o
1 "
CH 2
~OH+
22 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
1.3.2.3 The 'Taffy' process. The charge is such that the epichlorohydrinl
bisphenol A ratio will yield a resin with the required value of the degree of
polymerization 1 :::; n :::; 4, so that the upper limit for the average molecular
weight produced is about 1500. A stoichiometric amount of caustic soda in
aqueous solution is added with stirring and the reaction temperature raised
to 45-50°C. As the molecular weight increases the reaction temperature is
raised to 90--95°C for about 80 min with maybe increased pressure and more
vigorous agitation. At the end of the reaction period the product is in a water
resin emulsion plus an alkaline brine. The epoxy resin is recovered by
separating the phases, washed with water to remove inorganic salts and the
INTRODUCTION TO EPOXY RESINS 23
water removed by drying at temperatures of up to 130°C and under vacuum.
For purification of the resin dissolution in an organic solvent may be
advantageous and removal of water may be assisted by the use of methyl
isobutyl ketone. However, it is then essential that the level of solvent
remaining in the resin is minimized. The recovery of the resin in this process
is a major disadvantage especially because of the large amounts of brine that
have to be removed. These problems are not encountered in the fusion or
advancement process (section 1.3.2.4).
In the 'Taffy' process integral values of n, the degree of polymerization,
are usually produced with n values of (0),1,2,3 whereas in the advancement
process n is even numbered. A typical product with a weight per epoxide
(wpe) ca. 500 and n = 2 has a softening point of about 70°C and the practical
limitation for this process is n = 3.7 and a softening point of 95-100°C.
These resins will not crystallize as will DGEBA as mentioned previously and
the determination of their softening point is discussed in section 1.4.
During the latter stages of this process the temperature is increased up to
about 95°C and hydrolysis of the epoxy groups occurs so that there is a
relative high a-glycol value of ca. 0.5 eq/kg. Although the formation of
a-glycol will reduce the molecular weight attained, their presence may be
advantageous since they catalyse the reaction between amine hardeners and
epoxy groups and hence accelerate cure.
1.3.2.4 Fusion process. This is also known as the chain extension process
for reasons which will become clear. The starting materials for the fusion
process are bisphenol A plus a liquid epoxy resin which is essentially
difunctional in epoxy groups, produced by the processes used for the
manufacture of DGEBA. Thus
/0,
CHo-CH-CHo-O
-@-~e-@-
C
/0"
O-CHo-CH-CHo
- - I --
Me
+
Me
Ho--@-F--@-OH
Me
24 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
The reaction is carried out at temperatures between 180 and 200°C for about
30 minutes with a nitrogen 'blanket' to minimize oxidative degradative
reactions. The reaction is very sensitive to type and concentration of catalyst
which must be very strictly controlled to ensure production reproducibility.
Thus there has been considerable development of specialized catalysts to
minimize chain branching and gelation by promoting reaction between
phenolic hydroxyl with epoxy groups and limiting the side reactions,
especially between epoxy-epoxy and epoxy-alcoholic hydroxyl groups
(McAdams and Gannon, 1986; Savla and Skeist, 1977c). It can be appreci-
ated that the chain extension process is a step-wise polymerization and the
relationship between average molecular weight and extent of reaction will
depend on the degree of stoichiometric equivalence in the reacting mixture,
i.e. the epoxy/-OH ratio and the extent of side reactions which effectively
remove either epoxy or -OH groups and hence alter the effective
epoxy/-OH ratio. The important parameters were investigated by Batzer
and Zahir (1975, 1977).
Mixed isomcrs
Bisphcllol F
ti LJ OH
~CH'l2JrI,--@J
n
with n depending on reaction conditions and P/F ratio. Not all the
substitution is in the ortho position but the reactivity of the ortho position is
higher than that of the para because of the 'activating' effect of the phenolic
group.
The reaction conditions used to manufacture epoxy-novo lacs is similar to
those used with BPA resins and the idealized structure of the product is
There are considerable 'variants on a theme' with this type of resin; use of
excess epichlorohydrin minimizes the reaction of phenolic hydroxyl with the
epoxy groups attached to the novolac resin and limits the amount of
branching that can occur. Also it is essential that all phenolic hydroxyl
groups have reacted because their presence would adversely affect the
storage life of the resin and also volatiles would be formed during cure. The
epoxy novo lacs have improved thermal and chemical resistance compared
to the BPA resins.
26 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
0
..... / \ .,.-
"C-C, + H2 O
- HO OH
,I I"
"C-C, Glycol
"C-C,
0
..... / \ ,/
+ HOR
- OH
'C-C::'OR
" , Ether
"C-C,
0
...... / \
./
+ CH3-C'OH
O
'i
- OH
,I "
0
II
"C -C::·O-C-CH3 Ester
These reactions can be minimized by the use of peracetic acid in which the
sulphuric acid used in its synthesis is neutralized. With this reagent natural
oils can be epoxidized but it is not satisfactory for the production of
compounds with more reactive epoxy groups, even when reaction
temperatures are kept low and reaction times short. However, the hydroxyl
groups formed are reactive and can be employed in the cure of these resins.
The epoxidation of unsaturated compounds with a peracid is used in the
manufacture of cycloaliphatic epoxy resins. Not only are these resins free of
hydrolysable chloride and inorganic salts (ash) they do not contain aromatic
compounds and hence are more stable to UV exposure than the bisphenol A
derived epoxy resins. The presence of aromatic rings in BPA resins increases
the UV absorption of the resins and also degradative processes occur by the
formation of conjugated structures (Atherton et al., 1982; Bellenger et al.,
1986).
The starting materials for the production of cyclo-aliphatic resins may be
synthesized by Diels-Alder addition of unsaturated compounds. This can be
illustrated by the dimerization of butadiene to yield vinyl cyclohexene which
can then be epoxidized with a peracid.
28 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
"CH 2
CH Diels-Alder
I +
CH
''cH 2
/"
~xidized
~~~acid
with
Peracetic
acid
j Peracetic
acid
cH~'cH-CH,r o@-r@-o-cH,-b::-CH,t@-r@-O-CH'-cH~'cH'
T Me I Me
and in general there are different species present with different values of n,
the number of structural units. A first approximation in characterizing this
resin is to determine the concentration of epoxy groups present since this is
needed to calculate the amount of hardener required. A similar
consideration will apply to the other types of epoxy resin that are
commercially available.
However, the composition of epoxy resins is more complicated than that
represented by the general structure and characterization involves much
more than only determination of the epoxy content. Thus, the most
important parameters that manufacturers of uncured epoxy resins provide
as sales specifications include as well as epoxy content, hydrolysable
chlorine, specific gravity, colour and for liquid resins their viscosity at
ambient temperature. For solid resins the softening point and a solution
viscosity are often specified. Some resin manufacturers give detailed
instructions for some analytical procedures; for instance Ciba Geigy (1988)
give test methods for the determination of epoxy content and readily
hydrolysable and total chlorine contents. A detailed treatment of the
analysis of uncured resins is given by Urbanski (1977) with experimental
details of methods for the qualitative identification of epoxy resins and
procedures for their quantitative analysis. The characterization of uncured
resins involves determination of their detailed structure. Some of the
methods of analysis applied to uncured resins will now be considered.
a Parr bomb where the material is oxidized and then the chloride ion present
is determined as silver chloride (Jahn and Goetzky, 1988d). An alternative
procedure has been proposed by Urbanski (1977) in which the chlorine is
determined hydrolytically by heating for 2 hours with 0.5 N potassium
hydroxide in an ethylene glycol-dioxane solution, followed by potentio-
metric titration with silver nitrate in a dioxane-acetic acid solution. The
organically bound chlorine is obtained by correcting for any ionic chloride
ions that may be present, which may be determined by a standard pro-
cedure. In commercial resins, the inorganic chloride content should be very
low, but of course resin manufacturers have to monitor its presence to
ensure all of the salt formed by the dehydrochlorination reaction has been
effectively washed from the resin.
The organically bound chlorine will be in the forms known as (i)
hydrolysable and (ii) inactive. The presence of the former is due to
incomplete dehydrochlorination and the latter to either reaction of
epichlorohydrin with the secondary alcohol or abnormal addition of the
phenolic hydroxyl group as discussed previously (section 1.3). The
hydrolysable chlorine, also called saponifiable chlorine, can be determined
by reaction with excess caustic soda and then back titration with standard
hydrochloric acid (Jahn and Goetzky, 1988d). The inactive chlorine can
be calculated by difference from the total organically bound and the
hydrolysable chlorine.
where m and n are power law parameters and both are temperature-
dependent. When the power law exponent n = 1, the fluid is Newtonian
with m = YJ, the viscosity. For polymer fluids n is less than unity and the fluid
is pseudoplastic or shear thinning. For epoxy resins such as Epon 828 the
value of n at ambient temperatures is approximately unity (Chen and Ellis,
1992), but with higher molecular weight resins it is necessary to confirm that
they are Newtonian. The most suitable instrument is a cone and plate
viscometer but use of a Hoeppler viscometer with timing of the rate of
descent of the ball for two different angles of tilt may give sufficient data to
determine m and n (Barnes, 1980).
The temperature dependence of viscosity of many simple liquids is
Arrhenius, YJ = YJe exp [EYJIRT]' so that a plot of In YJ vs liT is linear.
However, in some cases it may be satisfactory to represent the temperature
dependence on a log-linear plot, so that:
log 1] = A + BT
References
Akhrem, AA., Moiseenkov, AM. and Dobryin, V.N. (1968) Russ. Chem. Rev. (Eng. Trans.)
37,448--462.
Atherton, N.M., Banks, L.G. and Ellis, B. (1982) J. Appl. Polym. Sci. 27,2015-2023.
Barnes, H.A. (1980) In Rheometry: Industrial Applications, ed. Walters, K. Research Studies
Press, John Wiley and Sons, Chichester.
Batzer, H. (1964) Chem. and Ind. 5, 179.
Batzer, H. and Zahir, S.A. (1975) J. Appl. Polym. Sci. 19,585-617.
Batzer, H. and Zahir, S.A. (1977) J. Appl. Polym. Sci. 21, 1843-1857.
Belen'kii, L.I. (1988) Adv. in Heterocyclic Chemistry 44,269-396. Synthesis of epoxy groups
p. 304. Reactions of epoxy groups pp. 303-330.
Bellenger, V., Verdu, J., Francillette, J. and Hoarau, P. (1986) Polymer Commun. 27,
279-28l.
Birley, A.W. and Scott, M.J. (1982) Plastics Materials, Properties and Applications, Leonard
Hill, Glasgow.
Braun, D. (1986) Simple Methods for Identification of Plastics 2nd edn. C. Hanser Verlag,
Miichen, pp. 27, 84 and PIT chart.
Chen, X.M. and Ellis, B. (1992) Unpublished measurements.
Ciba-Geigy (1988) Araldite for Surface Coatings, Laboratory Manual, Ciba-Geigy Plastics,
Cambridge.
Dannenberg, H. (1963) Soc. Polym. Eng. Trans. (Jan. 1963) p. 78.
Dearborn, E.C., Fuoss, R.M., Mackenzie, A.K. and Shepherd, R.G. (1953) Ind. Eng. Chem.
45,2715.
Dobinson, B., Hofmann, W. and Stark, B.P. (1969) The Determination of Epoxide Groups,
Pergamon Press, Oxford.
Ellis, B. (1972) Amorphous Materials Ch. 38, eds. Douglas, R.W. and Ellis, B. Wiley-
Interscience, London. pp. 375-388.
Faith, W.L., Keyes, D.B. and Clark, R.L. (1965a) Industrial Chemistry 3rd edn. John Wiley,
New York. p. 404. (1965b) p. 152.
Frost, AA. and Pearson, R.G. (1961) Kinetics and Mechanism 2nd edn. Wiley, New York,
pp.288-307.
Greenspan, F.P. and Gall, R.J. (1953) Ind. Eng. Chem. 45,2722-2726.
Gillham, J.K. (1986) Encylopedia of Polymer Science and Engineering, 2nd edn. John Wiley,
New York, pp. 519-524.
Gritter, R.J. (1967) The Chemistry of the Ether Linkage, Ch. 9, section II, ed. Patai, S.
Interscience Publishers, pp. 381-390.
Grobov, L.N., Sineokov, A.P. and Etlis, V.S. (1966) Russ. Chem. Rev. (Eng. Trans.) 35, 67l.
Hadad, D.K. (1988a) In Epoxy Resins, Chemistry and Technology, 2nd edn, Ed. May, C.A.
Marcel Dekker Inc, New York, pp. 1111-1124. (1988b) pp. 1096-111l.
Halton, B. (1991) Adv. in Ring Strain in Organic Chemistry 1,1-17.
Ives, G.C., Mead, J.A. and Riley, M.M. (1971) Handbook of Plastics Test Methods, IIiffe
Books, London.
Jahn, H. & Goetzky, P. (1988a) In Epoxy Resins, Chemistry and Technology, 2nd edn, ed.
May, C.A. Marcel Dekker Inc, New York, pp. 1050-1056. (1988b) p. 1067. (1988c) p. 1069.
(1988d) pp. 1070-1072.
Katritsky, A.R. and Jones, P.M. (1979) Adv. in Heterocyclic Chemistry 25, 303-391, see
327-329.
Katritsky, AR. and Weeds, S.M. (1966) Adv. in Heterocyclic Chemistry 7, 225-299, see
232-233.
36 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Kilner, E. and Samuel, D .M. (1960) Applied Organic Chemistry, Macdonald and Evans, p. 54.
Kopf, P.W. (1988) Enc. Polym. Sci. Eng. Vol. 11. Wiley-Interscience, New York, pp. 45-95.
Lee, H. and Neville, K. (1967 a) Handbook of Epoxy Resins, McGraw-Hili, New York, sections
1-2 to 1-4. (1967b) sections 3-6 to 3-8 (1967c) sections 4-13 to 4-15.
Lever, A.E. and Rhys, J.A. (1968) The Properties and Testing of Plastics Materials, Temple
Press Books, Feltham, U.K. pp. 166-169.
Lewars, E.G. (1984) Comprehensive Heterocyclic Chemistry, ed. Katritsky, A.R and Rees,
C.W. Vol. 7, section 5.0.5.4, ed. Lwowski, W. Pergamon Press, Oxford, pp. 114-118.
Lwowski, W. (1984) Comprehensive Heterocyclic Chemistry, eds Katritsky, A.R and Rees,
C.W. Vol. 7, Part 5, ed. Lwowski, W. Pergamon Press, Oxford, pp. 1-16.
McAdams, L.V. and Gannon, J.A. (1986) Ency. Polym. Sci. Eng. Vol. 6, Wiley-Interscience,
New York, pp. 222-382.
Malinovskii, M.F. (1965) Epoxides and their Derivatives, Section II. Israel Programme for
Scientific Translations, Jerusalem, pp. 38-101.
Materials and Technology (1972) Vol. IV. Petroleum and Organic Chemicals Longman,
London,pp.322-324.
Mestan, S.A. and Morris, C.E.M. (1984) Rev. Macromol. Chern. Phys. C24(l), 117-172.
Modern Plastics International (1982) Jan. p. 48.
Parker, R.E. and Isaacs, N.S. (1959) Chern. Rev. 59,737-799.
Peters, D. (1967) The Chemistry of the Ether Linkage, Ch. 1, ed. Patai, S. Wiley-Interscience,
New York (see especially p. 14).
Price, c.c. (1967) The Chemistry of the Ether Linkage, Ch. 11, ed. Patai, S. Wiley-
Interscience, New York, pp. 501-503.
Rabek, J.F. (1980a) Experimental Methods in Polymer Chemistry Wiley, Chichester, UK.
Ch. 15, pp. 221-254; Ch. 20, pp. 299-331. (1980b) Ch. 23,24,25. pp. 368-442. (1980c)
Ch. 9, pp. 123-141.
Ravindrath, K. and Gandhi, K.S. (1974) f. Appl. Polym. Sci. 24, 1115-1123.
Reinking, N.H., Barabeo, A.E. and Hale, W.F. (1963) f. Appl. Polym. Sci. 7,2135-2144.
Rosowsky, A. (1964) Heterocyclic Compounds with Three- and Four-membered Rings, Part I,
Ch. 1, ed. Weissberger, A. Wiley-Interscience, New York, Section II, pp. 31-181.
Sandler, S.R. and Karo, W. (1977) Polymer Synthesis, Vol. II, Ch. 3. Academic Press, New
York, pp. 75-113.
Savla, M. and Skeist, I. (1977a) In Polymerization Processes. ed. Schildkrecht, C.E. Wiley-
Interscience, New York, Ch. 16, p. 608 and Table 16.8. (1977b) p. 605. (1977c) pp. 589-591.
Schoff, c.K. (1988a) Ency. Polym. Sci. Eng. 2nd edn, Vol. 14. Wiley-Interscience, New York,
pp. 454-541. (1988b) pp. 64-71.
Stenmark, G.A. and Weiss, F.T. (1956) Anal. Chern. 28, 1784.
Streitweiser, A. (1956) Chern. Rev. 56,571-752, see pp. 682-690.
Swern, D. (1970) Organic Peroxides, Vol. 1, Wiley-Interscience, London, pp. 355-533.
Tanaka, Y. (1988) Epoxy Resins, Chemistry and Technology, 2nd edn, Ch. 2. ed. May, C.A.
Marcel Dekker, New York, pp. 9-283.
Tess, RW. (1988) In Epoxy Resins, Chemistry and Technology, 2nd edn, ed. May, C.A.
Marcel Dekker, New York, pp. 735-742.
Union Carbide (1989) Cycloaliphatic Epoxide Systems Union Carbide Chemicals and Plastics
Technology Corporation.
Urbanski, J. (1977) Handbook of Analysis of Synthetic Polymers and Plastics Ch. 10. ed.
Urbanski, J. Ellis Horwood, pp. 295-318.
Yandorskii, V.N. and Ershov, B.A. (1972) Russian Chern. Rev. (Eng. Trans.) 41, 403-410.
2 Curing agents for epoxy resins*
W.R. ASHCROFT
2.1 Introduction
Amines - aliphatics 45 10
- cycloaliphatics 30 7
- aromatics 15 3
- dicyandiamide 10 2
Polyamides 75 16
Polyamidoamines 30 7
Phenol- and amino-formaldehyde resins 75 16
Carboxylic acid functional polyesters 100 22
Anhydrides 55 12
Polysulphides and polymercaptans 15 3
Catalysts 10 2
Total 460 100
* Anchor Chemical (UK) Ltd-An Air Products Company. © Air Products and Chemicals,
Inc. 1992.
38 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
CHrCH(OH)-CHzO*
/
-> R1N
\
+ + ->
CHz-CH(OH)-CHzO*
(b)
CHrCH(OH)-CHzO* *OCHrCH(OH)-CH z
/ \
*OCHz-CH( OH)-CHz-N N-CHrCH( OH)-CHzO*
\ 1 /1
*OCHz-CH(OH)-CHrN
f ~
N-CH r CH(OH)-CH 20*
\ /
CHrCH(OH)-CHzO* *OCHrCH(OH)-CHz
(c)
*OCHrHC\-~Hz + HOR -> *OCHz-HC-CHz + R1NHz ->
o
\ I
ROH-O
*OCHrHC-CHrNHR 1 + HOR
I
OH
Scheme 2.1 Mechanism of cure of primary amines. -CH20* indicates that only the reactive
epoxy group is shown.
acids and amine carbamates. These can revert to amine and CO2 which in the
presence of water (humidity or excess) generates carbonic acid at the curing
surface which can then react irreversibly with amines to form the
undesirable amine bicarbonates.
The low molecular weights of the ethyleneamines mean the usage rates
are low and therefore they have to be mixed in to the epoxy resin with a high
degree of accuracy - apart from the need to avoid excess additions which
would exudelbloom, additions under stoichiometric amounts lead to
embrittlement. Various modification techniques have been developed,
however, to overcome most of the disadvantages and it is the derived
products, rather than the unmodified ethyleneamines, which are now used
extensively in epoxy applications. Most involve increasing the molecular
weight of the amine to reduce the volatility and most lead to significant
changes in reactivity and cured mechanical and chemical resistance
properties.
¢C
I
R R'
C0 2 H
¢C
I
R R'
CONH(CH 2 CH 2NH),H
C0 2 H CONH(CH 2CH 2NH)xH
R-CH=CH-CH=CH-CONH(CH2CH2NH)xH
polyaminoamide
42 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
R-CH=CH-CH=CH-C~
N
J
/N(CH 2CH 2NH)x_IH
polyaminoimidazoline
Generally tall oil fatty acids are used as the mono-basic acid component
and the resulting mixtures of polyaminoamide and polyaminoimidazoline
vary in pot-life/reactivity depending upon the imidazoline content. High
imidazoline content mixtures are long working life, low reactivity curing
agents which in the presence of water, particularly substrate surface
humidity, ring open back to the more reactive polyaminoamides. In effect,
they function as semi-latent curing agents and are widely used on damp
concrete for their excellent wetting characteristics and adhesion in concrete
coating, bonding new to old concrete and other adhesives applications.
Their low viscosity and excellent corrosion resistance also makes them
suitable for electrical potting and filament winding applications.
Mono-basic acids other than fatty acids have also been used by Wilson
(1970) in the preparation of pure imidazolines for use in fibre-reinforced
epoxy lamination applications where low reactivity, high strength and good
chemical resistance are important. Additionally, imidazolines, prepared
from mono-basic acids, have been further reacted with polycarboxylic acids
to form solid salts, and used by Veba Chemie (1977) for their unique ability
to provide matt finishes in powder coatings applications.
With liquid epoxy resin ad ducts prepared using an excess of amine, fairly
viscous materials of low volatility and higher mixing ratios result. Highet
mix ratios allow small errors to be made in weighing that would be critical in
the use of the parent amines. Once again due to the presence of in-built
hydroxyl groups, they are faster curing and thus ideal for laminating and
adhesives applications. Their ability to give excellent solvent resistance also
leads to usage with higher functionality resins for optimum solvent-resistant
coatings.
With solid epoxy resin, adducts can be prepared in two ways. In-situ
adducts are produced simply by addition of amine to a solution of the resin to
yield an adduct solution directly. This type is used widely for its ease of
manufacture, low cost, good film flexibility and excellent chemical
resistance, in solvent based coating applications. In-situ adducts do,
however, contain low levels of free amine that have a tendency to exude and
therefore mixtures with solid epoxy resin solutions often need to be aged
prior to application. Isolated adducts are prepared by addition of excess
amine to the solid epoxy resin solution followed by removal of unreacted
amine and reaction solvent by distillation. This yields solid isolated ad ducts
which need to be redissolved and used in the same way as the in-situ adduct
solutions. They do nevertheless exhibit a number of advantages associated
with the lack of any free amine induding low odour, and are better suited for
application at lower temperatures and higher humidity where bloom,
carbonation, loss of gloss and intercoat adhesion criteria are critical.
chemical resistance and are used for a wide variety of civil engineering,
coating, adhesive and laminating applications. Products of this type
generally contain 10-20% of unreacted phenols which aid compatibility for
coal tar extended epoxy coatings applications. They are causing legislative
concern and when phenol rather than an alkyl phenol is used in the
manufacture, 'skull and cross-bones' labelling is required for the resulting
products, even though the free phenol is associated with the amine in salt
form.
HO-Ar + H2N(CH2CH2NH)xH --+ Ar-O- + H3N+(CH2CH2NH)xH
+ PhOCH2-HC-CH2
(j
H2NCH2CH2NHCH2CH2NHCH2CH2CN
CURING AGENTS 45
The cyanoethylated amines are less reactive than the parent amines, low
in viscosity and give physical properties generally inferior to those of the
parent amines. They are not lower in irritation potential though and care
must be taken in operation to avoid the reformation of acrylonitrile from the
retro-Michael reaction which takes place at around 120°C. Their ability to
wet minerals and glass fibres well and low reactivity makes them particularly
suitable for laminating and filament winding applications.
HMD [124-09-4] which is more widely known for its use as an intermediate
for nylon-6,6 manufacture, is derived by tail to tail dimerisation of
acrylonitrile and subsequent catalytic reduction of the intermediate
adiponitrile. This synthetic route leads to a number of coproducts including
MPMD [15520-10-2] which is the hydrogenated head to tail dimerisation
product. TMD which was developed for use as a urethane intermediate is
46 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
x Molecular
weight
2.6 230
5.6 400
33.1 2000
x+y+z
5.3 403
CURING AGENTS 47
carbonation tendency compared to other types of aliphatic amines. This is
due to the slightly hindered secondary aliphatic amine functionality. When
cured they do show good film flexibilities and thermal cycling properties due
to the long chain distances between cross-link sites. The flexible aliphatic
nature does, however, limit their solvent resistance and heat resistance. This
combined with an inconveniently slow cure rate at ambient temperature
means that they are commonly used in combination with other amines and/
or accelerators, principally for flooring, adhesive, electrical potting and
encapsulation applications.
"'NCH'l§jCH,NH,
MetaXyleneDiAmine (MXDA)
N-AminoEthylPiperazine (AEP)
48 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Spiro-acetal diamines
n = 1--4
(a)
(*OCHz-CHOH-CH2)zN-C-N(CN)CHz-CHOH-CH2-0*
II
*OCHz-CH-CH2 + H2N-C-NH-CN ->
t;J
\ I II CH2-CH(0-)-CHz-0*
o NH
(b)
*OCHz-HC-CH2 + [R3N] -> *OCHz-HC-CH 2----
'cf I
(0-CHz-CH-CH20*)n
I
0----
(c)
R'z-N-C=N-R' R' z-N-C=N-R'
I I
R'-N-CN R'z-N-C=O
I
HO oI
I
*OCHz-CH-CH2- *OCHz-CH-CH 2-
R' = OCH2-CHOH-CHz-
Scheme 2.2 The mechanism of cure of dicyandiamide.
The reactive nitrile group readily trimerises to form the cyclic guanamine
or melamine [108-78-1]. General Mills Inc. (1965) found that reaction also
occurs with other nitriles such as those derived from fatty acids as in the
production of fatty guanamines. Amines add to the nitrile group to form
biguanides and Gempeler and Zuppinger (1971) discovered that aromatic
CURING AGENTS 51
Methylene-di(cyclohexylamine) or
H2 N - Q - C H 2D - N H 2 bi sParaAminoCyclohexylMethane (P ACM)
1,2-Cyc1ohexanediamine or
1,2-DiAminoCycloHexane (I ,2-DACH)
Of the many possibilities evaluated with epoxy resin only two have
achieved commercial success: 3,3'-dimethyJmethyJene-di(cyclohexyJamine)
[6864-37-5]; and the unhindered analogue PACM or methylene-di(cyclo-
hexylamine) [1761-71-3] as it is also known. Two grades of the latter are
available: PACM 48 which is a solid produced under conditions which
maximise the thermodynamically stable trans,trans [6693-29-3] isomer
content; PACM 20 which is a liquid produced with a low trans,trans and high
cis,trans [6693-30-7] cis,cis [6693-31-8] isomer content.
Both the 3,3'-dimethylmethylene-di(cyclohexylamine), which is also a
liquid, and PACM 20 are low in viscosity and light in colour. Like aliphatic
polyamines they are skin irritants but have lower vapour pressure which
reduces the hazard due to inhalation of vapour. When used to cure liquid
epoxy resins at temperatures in excess of 100°C they give excellent heat
52 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
resistance and high mechanical strengths and both are used in filament
winding, laminating, tooling and small-medium sized casting applications.
The 3,3' -dimethyl substituted material which is relatively hindered and thus
lower in reactivity being chosen for longer working life and the PACM 20
due to its higher conformational fluidity chosen for its superior fracture
toughness. PACM 20 has been found better suited to epoxy usage than
PACM 48 due to the ease of mixing with liquid resins and is also especially
suitable for modification as an adduct for ambient cure applications where
the flexibility and unhindered nature can be used to advantage.
1,2-DACH [694-83-7] is not produced by catalytic hydrogenation of an
aromatic precursor but is available commercially as the principal coproduct
from the catalytic reduction of adiponitrile (see section 2.2.1.7). This unique
1,2-diamine has a higher degree of reactivity even compared to the PACM
20 and this coupled with the high volatility and the limited flexibility due to
the short distance between cross-link sites makes 1,2-cyclohexanediamine
better suited for use in modified form in ambient cure epoxy applications.
1,8-Menthane diamine
IsoPhoroneDiAmine
(IPDA)
IPDA which, like its stable-mate TMD (section 2.2.1.7), is also a product
of a multi-step synthesis starting from acetone, via isophorone. The low
viscosity, light colour, excellent heat resistance and cured mechanical
strength mean that in unmodified form IPDA is particularly useful for
small-medium sized casting, filament winding, electrical and general
laminating applications. The differential reactivity resulting from the
presence of primary aliphatic and cycloaliphatic amine functionality content
allows selective adduction with epoxy resin and selective formylation/
Mannich base formation. This results in low viscosity, activated derivatives
suitable for room temperature curing high performance coatings and civil
engineering applications.
CURING AGENTS 53
2.2.2.3 Accelerated cycloaliphatic polyamines, adducts and Mannich
bases. The lower reactivity of the cycloaliphatic primary amine function is
caused in part by steric hindrance from the ring( s) and though an undoubted
advantage for heat-cure applications it does mean that acceleration is
necessary to achieve reasonable room temperature cure rates. As cyclo-
aliphatic polyamines are corrosive, irritant to the skin and have a strong
tendency to carbonate/water spot due to their high basicities they are also
best modified for ease and safety of handling. Although only a limited
number of the modification techniques used with aliphatic polyamines are
applicable to cycloaliphatic polyamines a range of curing agents with widely
varying properties can none the less be produced. Usage rate, viscosity,
reactivity, cured mechanical properties and chemical resistance are all
dependent on the formulating technique used.
The simplest modification approach involves blending the cycloaliphatic
polyamine with an accelerator and a plasticising diluent. Accelerators which
can be used include aliphatic polyamines, simple tertiary amines and organic
acids. The most used plasticising diluent particularly in combination with the
most used cycloaliphatic polyamine IPDA is benzyl alcohol as this latter
reduces brittleness, provides a weak accelerating effect and helps the
hindered amine get through the B-stage. Applications for this low viscosity
curing agent type include highly filled screed flooring and coatings
applications.
Ciba Ltd (1970) found a further useful method of modifying cycloaliphatic
amines involving partial pre-reaction with epoxy resin coupled with addition
of plasticisers such as benzyl alcohol to reduce the brittleness and lower the
viscosity. Accelerators may be added but are not essential as the hydroxyl
group formed in the adduction process has an accelerating effect, as does
benzyl alcohol. IPDA, 1,2-DACH and PACM 20 are all used to produce this
type of curing agent; the IPDA-based products performing as industry
standard; the DACH-based materials are used for optimum chemical
resistance; the PACM 20 ad ducts providing the best overall carbonation and
water spotting resistance. Applications for these high gloss, blush-resistant,
colour-stable curing agents include decorative and chemical spill resistant
high solids coatings and flooring applications, as well as laminate/gel coat
applications.
Rutgerswerke (1969) reported cycloaliphatic polyamines can also be
reacted with formaldehyde and phenols to form Mannich bases and again
plasticising diluents such as benzyl alcohol are best added to lower viscosity
and reduce brittleness. The Mannich bases are darker in colour than the
adducts and blends and are not colour stable, but do however cure better in
the presence of humidity and at lower temperatures. IPDA is effectively the
only cycloaliphatic polyamine type used, as the other polyamine types give
products which are too viscous to be of any value. Applications for the
Mannich bases include flooring, old concrete priming, new to old concrete
bonding and coal tar modified epoxy coatings.
54 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Meta-PhenyleneDiamine (MPD)
Of all the aromatic amines, MPD [1477-55-0} provides the tightest cross-
link density and exhibits the best solvent resistance. Philipson (1959) found
MPD, when blended with other solid aromatic amines, formed low viscosity
eutectics useful in the manufacture of chemically-resistant filament wound
pipe. DDM or MDA as it is also known, prepared by condensation of aniline
with formaldehyde exists in a number of forms ranging from the pure 4,4'-
diamino-diphenylmethane solid [101-77-9} with a melting point of around
90°C to low softening-point mixtures of methylene bridged polyphenyl
polyamines which are easier to melt/mix with liquid epoxy resin, as reported
by Seeger and Fauser (1954). The low polarity of the pure DDM curing
agent which imparts excellent electrical insulation properties combined with
high mechanical property retention even under conditions of high humidity,
make this the most suitable candidate for electrical casting, laminating and
adhesives applications. Although the most widely used of the aromatic
amines, DDM is now classified as a potential human carcinogen and for
many applications is being replaced by other types of curing agents.
CURING AGENTS 55
4,4'.DDS [80-08-0] has the principle advantage over other aromatic
amines in providing the highest heat resistance. It has become the standard
curing agent, for use in combination with the BFrMEA catalyst (see section
2.5.1), with specialised epoxy resins such as the tetra-glycidyl ether of
DDM, in high temperature tooling and high performance military and
aerospace laminating applications. 3,3'·DDS [599-61-1]' although giving
reduced heat resistance compared to the 4,4'- analogue, has been adopted
for its enhanced honeycomb peel strength in aerospace laminating
applications.
~NH2
Et~Et
DiEthylTolueneDiAmines
(DETDA)
NH2
Teo"a·alkyl DDM
R R
~ ~
(b)
R3N+ -CHz-CH(O-)-CH 2-O* ---> R3N+ -CHz-CH-CHz-O*
I
+ HOR' OH + -OR'
R'O- + H2C-CH-CHz-O* ---> R'O-CHz-CH(O-)-CH2-0*
'd
DiazaBicycloUndecene (OBU)
58 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
When used as a unique curing agent they react with epoxy resin at the 3-N
position to form a 1: 1 molar adduction product (Scheme 2.4a). They then
react at the 1-N position with a second molecule of epoxy resin (Scheme
2.4b) to form a 2 : 1 adduct which contains a highly reactive alkoxide ion
which initiates rapid anionic polymerisation of epoxy resin. For 1-(N)-
substituted imidazoles, 1 : 1 molar adducts with highly reactive alkoxide ions
CURING AGENTS 59
C=C
(a) C=C I \
I
\ + 'O-CH -HC-CH 'O-CH 2-HC-CH -N N
HN ,N 2 \ / 2 I 2, ~
'CR o HO CR
(b)
+ 'O-CH -HC-CH -
2 \ / 2
o
Scheme 2.4 Mechanism of cure of imidazoles_
R"
R = Me, R'= H: R"= NHCONMe2: 2,4-to1uene di-isocyanate derivative
the availability of a wide range of relatively cheap raw materials. These can
be combined together easily to give the necessary properties of correct
softening point and reactivity for powder coatings applications.
The mechanism of cure involves two stages: addition of the carboxyl
group to the epoxy functionality (Scheme 2.5a) and esterification with the
secondary hydroxyls on the epoxy backbone (Scheme 2.5b). The water
released in this latter condensation reaction then volatilises from the film
during the cure process.
The poor resistance to weathering of the bisphenol A, di-glycidyl ether
epoxies has precluded widespread use in external topcoat systems, but their
excellent corrosion resistance and adhesion make them ideal reaction
partners for carboxyl functional polyesters in a number of appliance and
metal-finishing powder coatings applications. Epoxy resins with high
aliphatic contents such as Tri-GlycidylIsoCyanurate (TGIC) on the other
hand provide improved weatherability and are used with carboxyl functional
polyesters for architectural powder coatings.
R = -CH 2-CH-CH 2
'd
TGIC
(b)
(
CO
Co
P +NR3 -
(COWR3
Scheme 2.6 (a) Mechanism of cure of uncatalysed anhydrides. (b) Mechanism of cure of Lewis
base catalysed anhydrides.
CURING AGENTS 63
o > (X
©reo I
eo> 0=>
co co co
PA THPA HHPA
Me
(XCD
.-
:
,
7
MTtIPA
\
/
co
0
(XCD'0
Me/.
MHHPA
/
CO ac
I
Me/.l'
MTHPA
P
CD
co
\
orNMA
preferred, and used with cycloaliphatic epoxy resins for optimum light
stability.
Methyl Endomethylene Tetra-HydroPhthalic Anhydride (METHPA or
Nadic Methyl Anhydride (NMA) [25134-21-8] as it is also known), prepared
by Diels-Alder reaction between methylcyclopentadiene and maleic
anhydride is used in components where electrical property retention at high
temperatures is required. NMA is the anhydride of choice here due to the
high rigidity of the fused cycloaliphatic ring backbone.
CI
C'W)
CI~CO
CI
HET DDSA TMA
/o ocIQr
0 co'0/
'oe CO
PMDA BTDA
PMDA which is the more compact cross-linking agent of the two gives the
better chemical resistance and the best resistance of all anhydride curing
agents. It is however a high melting solid and very reactive towards epoxy
CURING AGENTS 65
resin at high temperatures. It is used therefore in combination with other
anhydrides such as maleic anhydride to make it easier to dissolve and more
convenient to handle for a variety of electrical casting applications. Maleic
anhydride by itself is not a particularly useful curing agent as it produces
brittle cured epoxies. BTDA, also a high melting point solid, is somewhat
easier to handle due to its lower reactivity and is mostly used, alone or in
combination with other anhydrides, for high temperature-stability
electronic moulding powder and adhesive applications.
2.3.1.5 Linear poly-anhydrides. Black et al. (1964) have reported the use
of linear aliphatic poly-anhydrides derived by dehydration polymerisation
of adipic, azelaic and sebacic acids which contain terminal carboxylic acid
groups at either end of long chain aliphatic polymers. They are used
essentially as flexibilising modifiers for other anhydrides where
improvements in thermal shock resistance are required.
The novolac resins, which are the reaction products from formaldehyde and
excess phenol under acidic catalysis, when co-cured with high molecular
weight solid bis-A epoxy resins result in coatings with excellent adhesion,
film strength, flexibility and chemical resistance.
They are especially useful in powder coatings applications for corrosion
resistant pipeireinforcing bars (rebars) and with brominated epoxy resins for
FR3 electrical laminate production. The cure mechanism (Scheme 2.7)
involves poly-addition to epoxy resin and is activated by acids such as
p-toluenesulphonic acid.
The resole resins are the reaction products from excess formaldehyde and
phenol under basic catalysis. When co-cured with high molecular weight
solid bis-A epoxy resins in a poly-etherification reaction (Scheme 2.8)
because of the secondary hydroxyls on the epoxy backbone they yield even
higher cross-link densities and higher chemical resistance than novolac
resins. As such they are well-suited to drum and pail coating applications
where the high stoving temperatures easily drive off the water produced in
the condensation cure reaction.
NHCH 20H
U-F resin M-F resin
Rl = n-C4H9 , iso-C4H9 , C2Hs, CH3; R2 = another M-F resin
The mechanism of cure with epoxy resin involves: etherification with loss
of water or alcohol through the many secondary hydroxyls on the backbone
of the high molecular weight solid bisphenol A epoxy resins (Scheme 2.9a)
and, addition of N-methylol groups to the epoxy functionality (Scheme
2. 9b ). M-F amino-resins will also catalyse homopolymerisation of the epoxy
resin.
(a)
CH 20H + HO-CH(CH 20*h CH2-O-CH(CH 20'h + H20
***-N\ 1
---+ ***-N\ 1
CH20R + HO-CH(CH 20*)2 CH 2-O-CH(CH20*h + HOR
(b)
CH 20H + H2C--CH-CH 20* CHz-0-H 2C-CH(OH)CH20*
°
~
/ \ I /
'**-N ***-N
\ \
Scheme 2.9 Mechanism of cure of amino-resins.
The resulting cured films offer a combination of the best features of epoxy
and amino resins namely high adhesion, chemical resistance, gloss
retention, colour and colour retention and, with correct choice of amino-
resin, very good flexibility for metal decorating applications. M-F resins are
used for their superior film hardness and gloss properties where high
performance demands outweigh cost considerations, notably varnish and
moulding powder applications. U-F resins offer cost performance benefits
for fast bake enamel, stoving primer and can and drum top coating
applications.
The thiol or mercaptan group (-SH) is able to react with epoxy resin in an
addition reaction. This requires catalysis at room temperature by amines
which promote production of reactive mercaptide ions (Scheme 2.10). With
°
Scheme 2.10 Mechanism of cure of thiols/mercaptans.
68 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
2.4.1 Polysulphides
H(S-CH2CHz-OCH20-CH2CH2-S)X<SCH2CH(SH)CHzS)yH
2.4.2 Polymercaptans
R-[O-{ CH 2CH2CH20-} nCH2-CH(OH)-CHz-SHh
n = 1 - 2; R = aliphatic hydrocarbon
2.6 Summary
It can be readily appreciated that the precise choice of curing agent type can
markedly affect the properties of an epoxy resin when cured. As it is unusual
for a single catalyst or curing agent to provide optimum characteristics in any
given application, the identification and selection of a satisfactory curing
agent is often, therefore, in itself a complex procedure. Dunn (1984) has
however successfully rationalised the technique for surface coatings
applications and Ashcroft (1991) has attempted civil engineering/
construction applications.
References
Although the cure of an epoxy resin involves reaction between epoxy and
hardener reactive groups, as discussed in chapter 2, the full characterization
of the cure process involves many other factors. During cure, a liquid or fluid
resin-hardener mixture is converted to a solid (chapter 1, section 1.1 and
Figure 1.1). Critical features are gelation and the onset of vitrification. The
latter effect occurs as the glass transition temperature, Tg , approaches the
cure temperature, Te. The determination and alternative definitions of Tg
will be considered later (section 3.2). However, it should be noted that the
glass transition temperature is a function of the extent of reaction, X e ,
X = Eo - E(I,)
e E
o
with Eo and E(t,) being the concentrations of epoxy groups present initially
and at cure time te and has been used as a measure of the extent of reaction
(Wisanrakkit and Gillham, 1990). When the difference between the cure
temperature and the glass transition temperature becomes 'small', that is,
I::..T == Te - Tg(Xe) is small, curing reactions become diffusion controlled
because molecular mobility is rapidly reduced as I::..T -'> O. Cure may
continue very slowly even when I::.. T becomes negative. Then the glass
transition temperature exceeds the cure temperature.
Initially when Te > Tg (0) the rates of the curing reactions are chemically
kinetically controlled, and the rate expressions which have been proposed
for the early stages of cure will be discussed in section 3.4. With the
formation of larger more highly branched molecules, a critical point is
reached, the gel point, and then there is the formation of an infinite three-
dimensional network. If Te is too low, vitrification may occur before gelation
and then further reaction may be inhibited. An elevated temperature may
be used to reduce the viscosity so that the hardener can be mixed into the
resin and then the curing reactions can be quenched by rapidly cooling the
mixture to a storage temperature below Tg(O) , the glass transition
temperature of the 'compounded' resin. However, it is necessary to ensure
that mixing is complete before gelation, since the viscosity rises rapidly at
the gel point and the resin will not flow after formation of an incipient
network. Thus, information on gelation and vitrification is required to
KINETICS OF CURE 73
In this section the critical conditions necessary for gelation will be outlined.
After gelation the sol fraction, Ws decreases to essentially zero and the gel
fraction increases so that Wg ~ 1, where Ws and Wg are the weight fractions of
sol and gel respectively. With increasing extent of reaction the structure of
the network changes with more and more chains being 'tied' into the
network. Also the length of the network chains decreases as the degree of
cross-linking increases. Many of the physical and mechanical properties of
the cured resin depend on the details of the network structure. Because of its
crucial importance for the final stages of cure the glass transition
temperature will be considered, after discussion of gelation. There are still
many theoretical papers each year on the nature of the glass transition but
here only the essential features will be outlined which are relevant to the
cure of epoxy resins.
For an initial analysis of the change in macroscopic properties of epoxy
resins during cure it is usual to assume that the curing resin is homogeneous.
That is, the property changes are uniform throughout the sample. However,
there is some evidence for uneven curing (Ghaemy et al., 1982), and there
have been a number of studies of the morphology of cured resins. However,
with BADGE cured with DDM it was not possible to identify any structure
larger than a few bisphenol A repeat units but with other hardeners which
were less soluble in BADGE some features could be identified (Rashid,
1978). It would appear that at least some of the morphological features
observed in the cure of resins are due to poor mixing and dissolution of the
hardener in the resin. In other cases there may be either phase separation or
the formation of micro gel during the curing reaction. Ideally, before
interpreting macroscopic property changes the morphology of the cured
KINETICS OF CURE 75
+ = 'f.NJ;
Jav 'f. N
I
\
\
\
\
\
\
\
\
\
\
\
\
\
(a)
(b)
Figure 3.1 The spatial configuration of highly-branched moledules. (a) Tetra functional
junction point. (b) Small portion of a three-dimensional network with tetra functional junction
points. --> indicates that the junction point is connected to other network chains - for a tetra
functional junction point there will be three such connections.
= af[a(f - )r -1 (3.1)
KINETICS OF CURE 77
rank
--------------~~~----------------o
Figure 3.2 Formation of branched molecules. (e), Reacted functional group; (0), reactive
functional group. For alternative representations and discussion of molecular trees and
networks see Gordon and Ross-Murphy (1975).
when [aU - 1)] < 1 the number of branches in the nth rank decreases as n
increases, Bn ~ 0 as n ~ 00. However, when [aU - 1)] > 1 then Bn > 0
for all n. Thus, there is finite probability that an infinitely large molecule will
be formed, i.e.
1
a =-- (3.2)
C f-l
where a c is the critical branching coefficient required for incipient network
formation.
It should be noted that this prediction of the critical condition for gelation
excludes the possibility of ring formation and hence the observed gel point
may occur at a higher extent of reaction.
The relationship between a and the extent of reaction for a simple case
may be obtained by the following argument. The calculation of a is
equivalent to the probability that a branched unit is part of the structure
~HE---EH~
where HE and EH are reacted epoxy and hardener groups. This structure
must be present when a chosen reacted hardener group is part of a link
joining two branched units. This chosen hardener group can only be part of
such a structure if it has reacted and is also attached to an epoxy group which
has also reacted. Thus
a = PH,PE (3.3)
where PH = probability that a hardener group has reacted
(3.4)
78 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
where [Hlo and [Elo are initial concentrations of hardener and epoxy groups
and [HlI and [Ell are the concentrations at time t.
It should be noted that equal reactivity of hardener functional groups must
apply for this definition of PH to be exact. This is often not the case as will be
discussed in section 3.4 (Durand and Bruneau, 1983). When the initial
concentrations of hardener and epoxy groups are exactly equal and only
reactions between such groups occur PH = PE = P and a = l. The use of
equation 3.1 to calculate Bn , the number of branches in Rank n is only valid
when all reactions are intermolecular. Thus, it becomes more and more
inaccurate after gelation. However, it does illustrate the small degree of
branching for low extents of reaction. Also, the number-average molecular
weight only increases slowly with extent of reaction and is finite at the gel
point (Figure 3.3). The number of average molecular weight can be readily
calculated since it is the total mass divided by the number of molecules
present.
M = Mass (3.6)
n NI
where No is the number of molecules present initially and Nb is the number of
bonds formed, since for each bond formed the number of molecules present
decreases by one. The general case is treated by Macosko and Miller (1976)
who give relations for calculation of number, weight and Z average
molecular weights as a function of the extent of reaction. They have given a
general consistent treatment of non-linear polymerizations which is now
frequently cited in papers dealing with the cure of epoxy resins. Similarities
with the classical treatments of Flory and Stockmayer are pointed out so that
the relationship of their approach to that of others can be appreciated. In
Figure 3.3 the weight average molecular weight is also plotted and it can be
seen that it becomes infinite for P > Pc at the gel point. For P < Pc the
weight fraction of sol is unity, Ws = 1. For extents of reaction greater thanpc
the weight fraction of sol, w" decreases monotomically and the weight
fraction of gel, wg , increases correspondingly (ws + Wg = 1), as shown in
Figure 3.3. Network formation, in terms ofP(X3) and P(X4) is also shown in
Figure 3.3, the former has maximum for P - 0.85, and P(X4) approaches
unity only when P ~ 1. Thus, for the network to be fully 'cross-linked' cure
has to be essentially complete.
A general treatment of gelation was given by Case (1957) and re-
examined by Tanaka and Kakiuchi (1965) who assessed its applicability to
the cure of epoxy resins with acid anhydrides. Other work on the application
KINETICS OF CURE 79
5-------------------------
ws
~--.,1.0
Average I Weight
molecular 4 I 0,8 fraction
weight / and
(arbitrary I probability
units) / P(X n )
3 I
I 0,6
I
I
I
P(Xj__ "
~ 'f', 0.4
/" I',
/ / \
I I \
I I \
"/ IIp (X.) \ 0,2
/, / \
/' '')c./ \
, ,.~ ,.
;' ........
.... -
°0~----OL,'----0~,2----0~,-3---0~.4----~0,-5--JWO,~6~~~-0~,;~--OL,8---~-~0~,-9--~1.8
Extent of reaction - p
Figure 3.3 Effect of the extent of reaction on molecular properties for a stoichiometric mixture
of branching theory to the cure of epoxy resins has been discussed by Tanaka
and Mika (1973) who annotate research published by Japanese workers_ A
more general review is that of Dusek (1986) who shows that it is necessary
to consider the kinetics of the epoxy reactions to model the curing process
when there are substitution effects. The statistical theory yields only
approximate results due to neglect of stochastic correlations when a
substitution effect is important. The importance of the gelation and network
formation has led to the development of several different mathematical
models. Gordon has published a large number of papers using graph theory
to model the structure of networks, and a good starting point is the paper
by Gordon and Ross-Murphy (1975). The sol-gel transition is a critical
phenomenon and its relationship to other critical processes is outlined by
Stanley (1986). Ziman (1979) considers macromolecular disorder and deals
with gelation as part of a general discussion of models of homogeneously
disordered systems. Scaling concepts have been used to predict the sol-gel
80 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
(a)
tan I)
Tg Temperature
(b)
9~'----- --~--
I
I
I
----'* -----
I
midpoint
log E' 8 I I
I I
I I
I I
I
7 I
I
Tg m Temperature
(c)
\ .
\ Intercept
9 --11----
log E' 8
Temperature
determine the extent of chemical reaction but changes in the physical and
mechanical properties of the resin with cure treatment, time and
temperature (te, Te) are also important. The methods that are used to
monitor cure as functions of Te and te may be classified as follows.
1. Direct assay of the concentration of reaction groups present, usually
the epoxy concentration.
2. Indirect estimation of the extent of chemical reaction.
3. Measurement of the changes in physical and mechanical properties.
For a full interpretation of the changes occurring during cure it is necessary
to monitor both the concentration of reactive groups and other properties
(Verchere et al., 1990; Adolf et al., 1990).
To obtain data amenable to a kinetic analysis of the rates of the competing
consecutive reactions that occur during cure, the concentration of the
reactive groups must be assayed as a function of cure time, to at a specific
isothermal cure temperature Te. In some cases non-isothermal data have
been analyzed but they are not so reliable as isothermal data. However, it
should be noted that collection of isothermal rate data may involve con-
siderable experimental difficulty. For instance it is not possible for samples
to reach the cure temperature instantaneously, heat has to be supplied to the
sample to raise its temperature, i.e. Tambient ~ Te. Also, the resin is curing so
that there may be temperature gradient across the sample and subsequently
the heat of reaction may cause its centre to become much higher than the
nominal cure temperature Te. This may cause sampling problems since the
extent of reaction may be inhomogeneous.
overtone region and hence the spectra are less complicated so that problems
due to the presence of overlapping bands are avoided. Also, the sample path
length is longer, 1-10 mm, and hence sampling is more representative and it
is possible to use a glass cell to confine a liquid sample. Goddu and Delker
(1958) studied the overtone region and Dannenberg (1963) extended their
work to establish calibrations which enable accurate assay of epoxy and
amine group concentrations during cure. Recently (Ellis et al., 1992) the
method has been extended by the design of transparent soda-lime-silica
glass cells so that the extent of cure could be studied from initiation until the
epoxy absorption becomes too low to measure.
x = 1- f1Ht
f1HT
where f1Ht and f1H T are the amounts of exothermic heat evolved in the time
internal t = 0 to tc and the total heat evolved. The method involved
integrating the areas under exothermic peaks which requires the speci-
fication of a procedure for 'drawing' a baseline, which has been discussed by
Barton (1985) and Prime (1981) cites a case where a dynamic DSC scan does
not return to the baseline. However, Wisanrakkit and Gillham (1990)
established a good correlation between the extent of reaction and the glass
transition temperature of the curing resin samples.
Barton (1985) discussed, with reference to most of the early work, non-
isothermal reaction kinetics when the temperature of cure is increased at a
heating rate, f3 = dT/dt. Many mathematical approximations (Flynn and
KINETICS OF CURE 87
Wall, 1966) have been applied to analyse the data, and some have been
re-examined by Keenan (1987) who evaluated the errors involved in
estimating activation energies for the curing of two epoxy resin adhesives.
A recent suggestion for the transformation of dynamic DSC measurements
into isothermal parameters for the cure of thermosetting resins is that of
Khabenko and Dolmatov (1990).
8.0
log E*
7.5
7.5
log E"
~------=" 6.5
5.5
6.5
4.5
log E'
0.8
~,
7.5
I \
I \
! \
I
I \
0.6 I \
tan B I \
(arbitrary I \
I \ 7.0
units) I
\
Log E'
tan S / \
\ (arbitrary
0.4
/iJ' \ tan 8 units)
I I \
I I \
I I \
! I
\
\
I \ 6.5
I \
,
0.2 I \
I \
I \
J
---
\\\
I
J
30
---- ~/
Figure 3.5 Dynamic mechanical properties as a function of cure time. (a) Log E* and log E"
vs te' The gel time, t e .ge !> can be determined from the change in slope of log E" vs te' (b) Log E'
and tan c5 vs te' The shoulder at lower cure times on the tan c5 vs tc curve is due to gelation.
The shear viscosity of a resin during cure may increase due to either
gelation or the onset of vitrification, that is, I:!T = Tc - Tg(tel becomes
small. Thus, the whole curve process cannot be monitored by measurement
of a shear viscosity, it is necessary to measure a complex modulus and its
KINETICS OF CURE 89
components (e.g. G*, G' and G"). Such measurements are difficult for a
sample which is liquid initially and eventually solid (Figure 3.5). An early
solution to this problem was the torsional braid technique introduced and
developed by Gillham and coworkers (see Aronhime and Gillham (1986) for
a review). An instrumented Torsional Brain Analyser is commercially
available and there are several other instruments available for the
measurement of the dynamic mechanical properties of polymers. The
technique is known as dynamic mechanical thermal analysis, DMT A, or
DMA (Flynn, 1989). The instruments are equipped with a computer and
suitable programmes which enable measurement of the changes in the
complex modulus and its components by factors of 103 .
Adolf and Martin (1990) have measured the changes in G* and its
components as functions frequency and extent of reaction and interpreted
their results using a percolation model. Harran and Landourd (1986) found
that the slope of a plot of log G" vs tc decreases at the gel point, and Chen and
Ellis (1992b) have confirmed that d (log E")/dtc decreases at the gel point.
More recently Harran and coworkers (Serrano et aI., 1990) have applied
a percolation model to analyse their complex shear modulus data, as
mentioned previously.
3.4.1 Introduction
The determination of the kinetics of cure of epoxy resins involves more than
just measurement of the rates of reaction of epoxy groups with the hardener.
This is because it is necessary to locate the gelation time, tc(gel), and also the
rates of reaction become diffusion-controlled with the 'onset' of
vitrification, i.e. when I1T = Tc - Tg(t,) becomes small. It is possible to
inhibit reaction between hardener and epoxy groups by reducing the
temperature so that the resin/hardener vitrifies. However, for elucidation of
the mechanisms of the chemical reactions occurring during cure it is
necessary to study the initial rates of reaction and also assay the extent of
reaction between epoxy and reactive hardener groups. Usually the epoxy
group conversion is determined but assay of the concentration of hardener
reactive groups is also desirable, especially when their reactivity is unequal,
as for example primary and secondary amine hydrogen atoms. Two
approaches have been used:
1. Generalized empirical rate equations.
2. Rate equations derived from proposed chemical mechanisms.
Both of these approaches will be considered and also the effects of
vitrification when the rates of the curing reactions become diffusion-
controlled.
90 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
~ ~
where k~, is the rate constant for the uncatalysed reaction, but the reaction is
often catalysed by hydroxyl or other groups and then the rate is faster, i.e.
kJ > k~
OH
/0" k, I
~CH-CH2 + H-N-H + [Catlo ~ ~CH-CH2-N-H + [Catlo
$ ~ .
(3.11)
When K is approximately a half then with use of K = 0.5 + ,'!,x and also
using the stoichiometric relation for reacted epoxy and active hydrogen
atoms Aj + A2/2 = Ao - xe/2, when A 20 = 0, equation 3.11 can be re-
arranged to
1 . dxe = (k j Co + k 11 x e) [1 + 2A2 .I"!.,x ] (3.12)
(Eo - xe) (Ao - xe/2) dt 2Aj + A2
A plot of the left hand side of equation 3.12, regarded as a reduced reaction
rate, versus xe will be initially linear when the conditions used to derive
equation 3.12 apply. This is because the term [2A2 .I"!.,K/(2Aj + A 2)] is
initially zero and small compared with unity up to maybe 50% consumption
of epoxy groups. Rorie et al. (1970) regarded their linear plots of reduced
reaction rate versus conversion to be satisfactory for the initial stages of
cure. Subsequently these plots have a maximum because the rate of con-
version decreases as the curing reactions become more and more diffusion-
controlled as I"!.,T = (Tc - Tg) becomes small.
92 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
dXe _ 1
dt - Eo
(dx e)
dt .
When the initial stoichiometry is exact and the concentration of active
amine hydrogen atoms and epoxy groups are exactly equal, Ao = EoI2.
Then using these substitutions equation 3.12 becomes:
(3.14)
where 1<1 = k1 Co Eo and 1<2 = 1(11 E6 Xe are constants and R = HolED the
ratio of the initial concentrations of active hydrogen atoms to epoxy groups.
Equation 3.14 can be derived for a catalysed reaction between active
hydrogen atoms and epoxy groups and also a hydroxyl group autocatalysed
reaction:
-Ill
E + H + Co --> A
E + H + [P-OH] ~ A
(3.16)
KINETICS OF CURE 93
which has the same form as equation 3.13 but with He = 1 and the constants
are different. Thus, the linearity of plots of the left hand sides of equations
3.13 or 3.16 vs Xe will not be sensitive to the value of I( if it is either 0.5 or 1.0.
Thus, the value of the relative rates of reaction of primary and secondary
amine hydrogen can only be found when the values of kl' k2 or 1(1 and 1(2 can
be determined separately with sufficient accuracy.
A generalisation of equation 3.16 which has been used and regarded as
'quasi-theoretically' justified is:
However, it is difficult to relate this equation to the reaction steps that occur
during an amine cure of epoxy resins. A simplified form of equation 3.17
with K] = 0 has also been used by Kamal and Sourour (1972,1973):
Macosko (1985) and Camargo et al. (1983) suggest that a general rate
expression which is useful for fitting conversion data is
which is the rate equation for an nth order reaction (Wisanrakkit et al. ,
1990).
Barton (1985) gives a list of the rate equations that have been fitted by
about 20 different kinetic studies with the extent of cure measured by DSC
methods. However, from the present discussion it can be appreciated that
the kinetics of cure may be complicated, even before the rates of reaction
become diffusion-controlled due to vitrification. The empirical equations
yield only limited information on the mechanisms of the chemical reactions
involved. However, they have utility by providing a rate function for the
analysis of kinetic data, such as that obtained from DSC measurements.
Thus they may be sufficient for specification of cure conditions for specific
resin-hardener combinations. Also, they may be useful for quality control
and the in situ monitoring of the cure process.
Mijovic and coworkers (Mijovic, 1990; Mijovic and Wang, 1989; Mijovic
et al., 1984) have been studying the cure of epoxy resins over a number of
94 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
years and recently (Mijovic and Lee, 1989) used equation 3.17 with the
'overall' reaction order, m + n = 2, and solved it using a fourth order
Rungge-Kutta numerical integration to obtain curves for the extent of
reaction versus time of cure for several isothermal cure temperatures. Their
system was DGEBA-cured with a mixture of 4,4 methylene-dianiline
(MDA) and m-phenylene diamine (m-PDA) and they applied branching
theory to model the increase in viscosity prior to the gel point.
Sheppard and Senturia (1986) as part of their programme of studies of the
effects of cure on the dielectric properties of epoxy resins have also
measured the extent of cure using DSC methods. They found that plots of
1 dXe
- vs X
(1 - Xe)2' dt e
(equation 3.16) were linear for the cure of DGEBA with DDS with cure
temperatures in the range 410-460 K. They also compared the time
required for 50% primary amine conversion estimated from the kinetic
model and obtained satisfactory agreement with experimental determina-
tions (see also chapter 7).
An even simpler rate equation is obtained when m = 0 in equation 3.18,
an expression used by Acitelli and coworkers many years ago (1971) and
recently Wisanrakkit and coworkers (1990) obtained a first order fit with
n = 1. For cure of DGEBA (Epon 828) with PACM-20 (see chapter 2) plots
of dXe/dt vs (1 - Xe) for cure at 61°C were linear. Keenan (1987) discusses
in detail a method for determining the parameters 1(2, m and n with Kl = 0
when a simpler form of equation 3.17 is applicable. His method involved
measurement of an exothermic peak and the procedure was satisfactory for
representation of the cure of FM123-5 (American Cyanamid Co. ), a nitrile
modified-dicyandiamide-cured epoxy adhesive supported on a nylon
carrier.
An interesting alternative approach is that of Pin sheng and Chune (1989)
who apply the non-equilibrium thermodynamic fluctuation theory proposed
by Hsich (1982). In this theory the change in a mechanical property is
measured and the rate equation derived after some simplification is:
G -G(t) = exp [- (tIT) fJ]
00
Goo - Go
where Goo, G(t) and Go are the final, time-dependent and initial values of
the physical property, f3 is a constant which depends on the width of the
relaxation spectrum and r is the relaxation time. Pingsheng and Chune
studied the cure of a commercial DGEBA resin E51 (Shanghai Resin
Factory) with imidazole as the hardener and silica filler. They were able to
model the change in the shear modulus as a function of both cure
temperature and time.
KINETICS OF CURE 95
3.4.2.3 Analysis of competitive consecutive reactions. The rate equation
for the consumption of epoxy groups, equation 3.10, contains terms for the
concentration of secondary amine and hence to specify the kinetics of cure,
additional rate equations are required. When k~ = k~ = 0 the additional
rate equations are
kil = - dA I . ~ (3.28)
a dt Y
and from equation 3.27
k22 = _ dA 3 . ~ (3.29)
a dt Z
96 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Other estimates of k12 can be obtained from three equations of the form
~
k',
1':
k",
KA,A,
A,-A, A,-A,
l~
t-<
A,-A,
Figure 3.6 Reactions of diamines. A]-A] unreacted primary diamine; A]-A2 unreacted
primary amine and a secondary amine; A3 reacted secondary amine, i.e. a tertiary amine. The
epoxy prepolymer unit is omitted except for A]-A2 . Verchere et al. (1990) measured the
concentrations of ArA2 plus A]-A3; A2-A3 and A3-A3 by size exclusion chromatography, and
used a kinetic scheme with k] = k'{ = k; = k] and k'2 = k z = k 2 .
determined the reactivities of the diamines and found that they increased in
the order:
4,4' DDS < 3,3' DDS < BAPP < BAPS < DDM
and the secondary amine reacts completely within 5 min at 90°C with about
115 of the epoxy groups to produce a secondary hydroxyl group. The
reaction then proceeds by tertiary amine catalysed reaction of both the
primary and secondary hydroxyl groups with equal probability. The gel
point occurs after about 3 h at 90°C. Fully cured resins have a glass transition
temperature of c. 70°C. This study confirms that in equation 3.10
kOH = kE = 0 compared with the rates of reaction of primary and secondary
amines with epoxy groups.
For the cure of BADGE with either TET A or diaminodiphenyl methane
Chen and Ellis (1992a) have extended the approach given by Frost and
Pearson (1961) so that both catalysis by impurities and also the product
hydroxyl groups are included in the rate equations. For TET A it was found
that the relative rates of reaction of secondary to primary amines increased
with cure temperature, at 22°C K = 0.25 and increased to K = 0.75 at 50°C.
Thus, the activation energy for the reaction of secondary amines was larger
than that for primary.
and of course the use of such a parameter depends on the definition and
experimental method used to determine Tg(O) , Tg(Tel and Tg(oo). Also, it is
essential that the limited value Tg(oo) is determined unambiguously. They
found that for cure of DGEBA type prepolymers, Ciba-Geigy CT200 and
CY207, with phthalic anhydride and also a mixture of phthalic and tetra-
hydrophthalic anhydrides that the overall kinetics is neither first nor half-
orders, but first order kinetics was satisfactory for 85-90% of the
converSIOn.
Steinmann (1987, 1989) has considered many of the previous reports on
anhydride cures and reports on her own work in which she studied the
reactions of the several epoxy resins with hexahydrophthalic anhydride
(HHPA) and tertiary amine catalysts.
The relationship between extent of cure and the glass transition temperature
is discussed in detail by Stutz et al. (1990). However, it may be difficult to
determine the extent of cure as the concentrations of reactive groups
approach zero. Even a DSC zero residual heat does not necessarily indicate
'full' cure (Stutz et al., 1990; Stutz, 1992). Also, Stutz (1992) recommends
control of the final stages of cure by measurement of the amount of solubles.
Although this parameter may require experimental experience to obtain
accurate measurements it can be used to calculate a degree of cure using
branching theory (Macosko and Miller, 1976; see section 3.2).
3.5.1 Introduction
Epoxy resins find applications in areas such as coatings, adhesives or matrix
materials for composites, as discussed in chapters 7, 8 and 9. One of the
properties required for these applications is that the material should be
essentially a rigid solid at the use temperature, Tuse, although in some cases
a degree of flexibility is required (see chapter 4). The use temperature in all
these applications is below the glass transition temperature of the cured
resin, Tuse < Tg• For higher temperature applications there is continued
development of epoxy-hardener systems with higher and higher glass
transition temperatures and hence higher Tuse. Of course it is important to
ensure that chemical degradation, due to oxidative, photochemical or other
processes, do not impair the properties of the cured resin. These
relationships are represented schematically in the TTT diagram, Figure 1.2.
In many applications of epoxy resins it is their glassy state elastic moduli
that are of paramount importance. Modulus maps for amorphous polymers
have been proposed by Gilbert et al. (1986) who confined their discussion to
thermoplastics, so-called 'linear' polymers, and considered poly(methyl
methacrylate) and poly(styrene) in detail. There is a need to represent the
structure-property relationships for thermosetting resins. The elastic
properties will be a function of the structure of the network and the
measurement temperature relative to the glass transition temperature of
the cured resin. Thus, an elastic property P = f(T,C) where T is the
measurement temperature and C is a cure parameter. Choice of measurable
quantities to define C presents several problems. Of course, C is related to
extent of conversion, X e , which is difficult to measure accurately when the
concentration of reactive groups becomes very low, as discussed previously
(see section 3.3.2). Also it is important to note that the relationship between
Tg and Xe is non-linear (Wisanrakkit and Gillham, 1990; Wang and Gillham,
KINETICS OF CURE 103
.... Tg CXl
\l1..{~\C! -
_4--
I
TgO ....
Gel
o
Cure Parameter, C
Figure 3.7 The CfP diagram: a modified version ofa TgIP diagram (Wang and Gillham, 1992)
C = Tg - Tgo
Tg", - Tgo
Tgo is the initial glass transition temperature of the uncured resin; Tg", is the glass transition
temperature of the fully-cured resin; Tg is the glass transition temperature of the partially-cured
resin.
where Tgo is the initial and Tgoo is the 'fully cured' glass transition
temperature. Such a definition requires that Tg is a measure of the structure
of the cured resin regardless of the precise cure treatment, that is, cure
temperature and cure time. With this definition C increases from zero to
unity as cure progresses.
An earlier review of the effects of cross-linking on the properties of
polymers is that of Nielsen (1969), which is still useful as a starting point for
the study of the relationships between properties and cure of thermosetting
systems. It should be noted that there has been more recent discussion of the
104 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
(3.42)
where VL and VT are the velocities ofthe longitudinal and transverse waves,
and Q is the density. Thus, the shear modulus can be calculated from
equation 3.42 and also using equation 3.39 with equation 3.41 Young's
modulus can be calculated.
Poisson's ratio may be estimated from its relationship with the
longitudinal wave velocity by
listed by Morel et al. (1989) for their calculations for epoxy resins. A more
complete listing is given by van Krevelen and Hoftyzer (1976). Use of this set
of molar contributions Ui underestimates the longitudinal wave velocity and
hence would lead to an underestimate of the elastic moduli.
Morel et al. (1989) also evaluated the applicability of equation 3.38 by
plotting the bulk modulus B vs 0 2 , the cohesive energy density. A straight
line plot was obtained but with a slope of 11.5 compared with the predicted
value of 8.04. They considered that their estimates of the solubility
parameters neglected the hydrogen bonding that is well known to occur in
epoxy resins. To estimate the degree of hydrogen bonding they measured
the equilibrium water absorption, Wm , at 100°C and 95% relative humidity.
There was a good linear fit between the experimental data and the water
absorption and elastic modulus, equation 3.44.
M = Mo + kWm (3.44)
where M is the modulus and the slope, k, depending on the contribution due
to hydrogen bonding, and Mo depending on the aromatic group
concentration. Good linear correlations were obtained for all three moduli,
B,EandG.
The moduli that are usually measured in the temperature range 50 to
150°C below the glass transition temperature of the resin are often con-
siderably lower than the limiting low temperature high frequency moduli
which have been discussed in this section. The stress-strain behaviour of
epoxy resins will be considered next.
8 306
367
''~"" 5
(jj
4 381
x
3
2 I--- 4 °/o---t
i = 3.3 x 10- 4 sec'
Strain (%)
Figure 3.8 Stress-strain curves for an epoxy resin at temperatures in the range 77 to 381 K.
Stress, kp/mm2; strains, displaced for clarity, see gauge length (from Pink and Campbell, 1974).
pertinent to note that when the structure of a cured epoxy resin is altered by
changing the concentration of hardener the Young's modulus has a
minimum when there is an exact stoichiometric mixture of epoxy groups and
reactive amine hydrogen atoms (Bell, 1970b; Selby and Miller, 1975; Kim
et al., 1978; Gupta et aI., 1985). It may be that the modulus is related to the
density of the resin at the measurement temperature, which also passes
through a maximum with cure time and hence extent of reaction (Bell et al. ,
1992). The maximum in the relative modulus is found at all cure tempera-
tures studied by Wang and Gillham (1992) with the relative modulus
increasing with cure temperature and the maximum occurring at lower
extents of reaction as the cure temperature increased (Figure 3.9).
From Figure 3.8 it can be seen that the curvature of the extensional stress-
strain curves is more pronounced at higher test temperatures. At room
temperature it has been found that for stresses lower that the yield stress the
stress-strain curve fits a power law relationship a = MEn with a value of
n = 0.93 for a BADGE type resin cured with DDM (Bell et al. 1992).
Williams (1979) suggests the use of the power law equation for the stress-
strain relationship for polymeric materials which do not obey Hooke's law
exactly with n = 0.9. When strains are low, M is approximately equal to
Young's for values of n not too different from unity.
108 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
theoretical gelation
end of glass
transnion
20
E = 3G = 3QRT (3.45)
r r M
e
for extension ratios of up to 1.2, where a is the applied stress, equal to the
applied force divided by the un strained area of the sample. Thus, from this
experimental value of Gr it is possible to calculate the equivalent molecular
weight of the network chains. This is a direct method of determining a
parameter which is a characteristic of the network structure. Other methods
involve branching theory and either assumptions or direct determination of
the relative reactivities of the reactive groups, such as the reactivity of
primary and secondary amine groups.
-40
- 80
-120
-160
theoretical gelation
-200~~~--~~r_--r_r_r---r---~--~--4
o 20 40 60 80 100 120 140 160 180 Tg
o 0.5 1 C*
Figure 3.10 The fJ relaxation temperature Tfi versus the glass transition temperature T g • The
resin-hardener system is the same as in Figure 3.9 (from Wang and Gillham, 1992). * C is the
cure parameter as defined in Figure 3.7.
References
Acitelli, M.A., Prime, R.B. and Sacher, E. (1971) Polymer 12, 335-343.
Adolf, D. and Martin, J.E. (1990) Macromolecules 23,3700-3704.
Adolf, D., Martin, J.E. and Wilcoxon, J.P. (1990) Macromolecules 23,527-531.
Amdouni, N., Sautereau, H., Gerard, J.-F. and Pascault, J.-P. (1990) Polymer 31,1245-1253.
See Fig. 2 and Table 2, p. 1348.
Argon, A.S. (1973) Phil. Mag. 28,839.
Aronhime, A.T. and Gillham, J.K. (1986) Adv. in Polym. Sci. 78,83-113.
Arridge, R.G.C. and Speake, J.H. (1972) Polymer 13, 443-454.
Banks, L. and Ellis, B. (1982) Polymer 23,1466-1472.
Barton, A.L.M. (1983) Handbook of Solubility Parameters and other Cohesion Parameters.
CRC Press, Boca Raton, USA.
Barton, J.M. (1985) Adv. in Polym. Sci. 72, 111-154.
Bell, J.P. (1970a) I. Polym. Sci. A2. 6,417-436.
Bell, J.P. (1970b) I. Appl. Polym. Sci. 32, 1901-1906.
Bell, J., Ellis, B. and Found, M.S. (1992) Unpublished experimental measurements.
Benson, S.W. (1960) The Foundations of Chemical Kinetics. McGraw-Hili, New York,
pp.42-46.
Bero, c.A. and Plazek, D.T. (1991) I. Polym. Sci., B. Polym. Phys. 29, 39-47.
Bidstrup, W.W. and Senturia, S.D. (1989) Polym. Eng. Sci. 29, 290-294.
Billmeyer, F.W. (1984) Textbook of Polymer Science 3rd edn, Wiley, New York, pp. 40-47.
Bondi, A. (1968) Physical Properties of Molecular Crystals, Liquids and Glasses, Section
13.3.1, Wiley, New York, pp. 397-400.
Bowden, P.B. (1973) In The Physics of Glassy Polymers Ch. 5, ed. Haward, R.N. Applied
Science Publishers, London, pp. 279-339.
Camargo, R.E., Gonzalez, V.M. and Macosko, C.W. (1983) Rubber Chemistry and
Technology 56, 774--778.
Carothers, W.H. (1936) Trans. Far. Soc. 32,39-53.
Case, L.C. (1957) I. Polym. Sci. 26,333-350.
Chen, X.M. and Ellis, B. (1992a) Paper submitted to I. Appl. Polym. Sci.
Chen, X.M. and Ellis, B. (1992b) Unpublished data, but see Chen, X.M. Ph.D thesis Sheffield
1988).
Charlesworth, J. (1980) I. Polym. Sci. Polym. Chern. Ed. 18,621-628.
Chiu, W.Y., Carratt, G.M. and Soong, D.S. (1983) see Figure 2 Macromolecules 16, 348-357.
Choy, I.C. and Plazek, D.J. (1986) I. Polym. Sci., B. Polym. Phys. 24, 1303-1320.
Crowson, R.J. and Arridge, R.G.C. (1979) Polymer 20,737-746.
Dannenberg, H. (1963) S.P.E. Trans. 3,78-88.
de Gennes, P.-G. (1979) Scaling Concepts in Polymer Physics, Ch. V, Cornell University Press,
Ithaca, pp. 128-162.
Dillman, S.H. and Sefferis, J.C. (1989) I. Macromol. Sci.-Chem. A26(1), 227-247.
Dobas, I., Eichler, J. and Klaban, J. (1975) Collection Czech. Chern. Comm. 40,2989-2997.
Durand, D. and Bruneau, C.-M. (1983) Polymer 24,287-591 and 592-595.
Dusek, K. (1985) Brit. Polym. I. 17,185-189.
Dusek, K. (1986) Adv. in Polym. Sci. 75, 1-59. See pp. 9-12 for discussion of the morphology
of cured epoxy resins.
Dusek, K., IIIavsky, M. and LuMk, S.J. (1975) I. Polym. Sci. Polym. Symp. 53,29.
114 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Elias, H.-G. (1976) Macromolecules Vol. 1 and 2. Plenum Press, New York, pp. 606-614.
Elliott, S.R. (1990) Physics of Amorphous Materials 2nd edn, Ch. 2, Longman Scientific and
Technical, Harlow, pp. 29-69.
Ellis, B., Bell, J., Chen, X.M. and Found, M.S. (1992) Paper in preparation.
Enns, J.B. and Gillham, J.K. (1983) 1. Appl. Polym. Sci. 2S, 2567-2591.
Entelis, S.G. and Tiger, R.P. (1976) Reaction Kinetics in the Liquid Phase, Ch. 1. Translated
from Russian, John Wiley, Chichester, pp. 1-25.
Eyring, H., Lin, S.H. and Lin, S.M. (1980) Basic Chemical Kinetics, Ch. 9, John Wiley, New
York, pp. 376-451.
Fava, R.A. (1968) Polymer 9,137-151.
Flory, P.J. (1941) 1. Amer. Chern. Soc. 63,8083-3100.
Flory, P.J. (1953) Principles of Polymer Chemistry, Ch. IX Cornell University Press, Ithaca,
pp. 347-398.
Flynn, J.H. (1989) Ene. Polym. Sci. and Eng. 2nd edn Suppl. pp. 715-723.
Flynn, J .H. and Wall, L.A. (1966) 1. Res. Natl. Bur. Stand. 70A, 487-523.
Frost, A.A. and Pearson, R.G. (1961) Kinetics and Mechanism 2nd edn, John Wiley, New
York, pp. 178-184.
Ghaemy, M., Billingham, N.C. and Calvert, P.D. (1982) 1. Polym. Sci. Polym. Lett. 20,
439-443.
Gilbert, D.G., Ashby, M.F. and Beaumont, P.W.R. (1986) 1. Mat. Sci. 21,3194-3210.
Goddu, R.F. and Delker, D.A. (1958) Anal. Chern. 30,2013-2016.
Gordon, M. and Ross-Murphy, S.B. (1975) Pure and Applied Chern. 43, 1-26.
Grillet, A.C., Galy, J., Pascault, J.P. and Bardin, 1. (1989) Polymer 30,2094-2103.
Grimmett, G. (1989) Percolation Ch. 1. Springer-Verlag, New York, pp. 1-24.
Guillett, J. (1985) Polymer Photophysics and Photochemistry see Ch. 3, pp. 52-70 and Ch. 6,
pp. 114-135 especially p. 132. Cambridge University Press, Cambridge.
Gupta, V.B., Drzal, L.T., Lee, c.y-c. and Rich, M.J. (1984-1985) 1. Macromol. Sci. B23,
435-466.
Hadad, D.K. (1988a) In Epoxy Resins, Chemistry and Technology 2nd edn. ed. May, C.A.
Marcel Dekker, New York, pp. 1111-1123.
Hadad, D.K. (1988b) In Epoxy Resins, Chemistry and Technology, 2nd edn. ed. May, C.A.
Marcel Dekker, New York, pp. 1126-1136.
Harran, D. and Laudourd, A. (1986) 1. Appl. Polym. Sci. 32,6043-6062.
Havlicek, 1. and Dusek, K. (1987) In Crosslinked Epoxies ed. Sedlacek, B. and Kahovec, J.
Walter de Gruyter, New York, pp. 417-424.
Horie, H., Hiura, H., Sawada, M., Mita, 1. and Kambe, H. (1970) 1. Polym. Sci. A1.S, 1357-
1372.
Hsich, H.-Y. (1982) 1. Appl. Polym. Sci. 27,3265-3277.
Huguenin, F.G.A.E. and Klein, M.T. (1985) Ind. Eng. Chern. Prod. Res. Der. 24, 166-171.
Ilavsky, M., Bogdanova, L.M. and Dusek, K. (1984) 1. Polym. Sci. Polym. Phys. ed. 22,265.
Jha, A. and Parker, J.M. (1989) Physics and Chemistry of Glasses 30, 220-228.
Kaelble, D.H. (1973) In Epoxy Resins, Chemistry and Technology, Ch. 5. eds. May, C.A. and
Tanaka, Y. Dekker, New York, pp. 327-37l.
Kaelble, D.H., Moacanin, J. and Gupta, A. (1988) Epoxy Resins, Chemistry and Technology
2nd edn, Ch. 6. ed. May, C.A. Dekker, New York, pp. 603-652.
Kamal, M.R. and Sourour, S. (1972) Society of Plastics Engineers IS, 93-98.
Kamal, M.R. and Sourour, S. (1973) Polym. Sci. and Eng. 13,59-64.
Kamon, T. and Furakawa, H. (1986) Adv. in Polym. Sci. SO, 173-202.
Keenan, M.R. (1987) 1. Appl. Polym. Sci. 33, 1725-1734.
Khabenko, A.V. and Dolmatov, S.A. (1990) 1. Thermal. Anal. 36,45-54.
Kim, S.L., Skibo, M.D., Manson, J.A., Hertzberg, R.W. and Janiszewski, J. (1978) Polym.
Eng. Sci. IS, 1093-1100.
Kong, E.S.W. (1986) Adv. in Polym. Sci. SO, 126-171.
Kovacs, A.J. (1958) 1. Polym. Sci. 30, 131-147.
Kovacs, A.J., Hutchinson, J.M. and Aklonis, J.J. (1977) The Structure of Non-Crystalline
Materials ed. Gaskell, P.H. Taylor and Francis, London, pp. 153-163.
Labana, S.S. (1986). Ene. Polym. Sci. and Eng. 2nd edn, Vol. 4, John Wiley, New York,
pp. 350-360.
KINETICS OF CURE 115
Larson, R.G. (1988) Constitutive Equations for Polymer Melts and Solutions, Ch. 4.
Butterworth, Boston, pp. 93-127.
Lee, A. and McKenna, G.B. (1988) Polymer 29,1812-1817.
Lee, A. and McKenna, G.B. (1990) Polymer 31,423-430.
Lee, H. and Neville, K. (1976) Handbook of Epoxy Resins Appendix 4-1. McGraw Hill, New
York, pp. 4-35-4-56.
Lunak, S., Vladyka, J. and Dusek, K. (1978) Polymer 19, 931-933.
Macosko, e.W. and Miller, D.R. (1976) Macromolecules 9,199-206.
Macosko, C.W. (1985) Brit. Polymer J. 17,239-245.
Martin, J.E., Adolf, D. and Wilcoxon, J.P. (1988) Phys. Rev. Lett. 61,2620-2623.
Malkin, A.Yu and Kulichikhin, S.G. (1991) Adv. in Polym. Sci. 101,217-257.
Matsuoku, S., Quan, X., Bair, H.E. and Boyle, D.S. (1989) Macromolecules 22, 4093-4098.
McKenna, G.B., Santore, M.M., Lee, A. and Durran, R.S. (1991) J. Non-Crystalline Solids
131-133,497-504.
Mertzel, E. and Koenig, J.L. (1986) Adv. in Polym. Sci. 75,73-112. For discussion ofNMR see
pp. 92-112 and FT-IR see pp. 73-92.
Mestan, S.A. and Morris, C.E.M. (1984) J. Macromol. Sci. Rev. in Macromol. Chem. and
Phys. C24, 117-172.
Mijovic, J. (1990) J. Appl. Polym. Sci. 40,845-866.
Mijovic, J., Kim, J. and Slaby, J. (1984) J. Appl. Polym. Sci. 29, 1449-1462.
Mijovic, J. and Lee, C.H. (1989) J. Appl. Polym. Sci. 38,2155-2170.
Mijovic, J. and Wang, H.T. (1989) J. Appl. Polym. Sci. 37,2661-2673.
Mita, I. and Horie, K. (1987)J. Macromol. Sci. Rev. Macromol. Chem. andPhys. C27, 91-169.
Moehlenpah, A.E., Ishai, O. and Di Benedetto, A. (1969)J. Appl. Polym. Sci. 13,1231-1245.
Morel, E., Bellenger, V., Bocquet, M. and Verdu, J. (1989)J. Mat. Sci. 24,69-75.
North, A. M. (1964) The Collision Theory of Chemical Reactions in Liquids. Methuen, London.
Nielsen, L.E. (1969) J. Macromol. Sci., Rev. Macromol. Chem. C3(1), 69-103.
Ochi, M., Iesako, H. and Shimbo, M. (1985) Polymer 26,457-461.
Oleinik, E.F. (1986) Adv. in Polym. Sci. SO, 50-99.
Peyser, P. and Bascom, W.D. (1977) J. Appl. Polym. Sci. 21,2359-2373.
Pingsheng, H.E. and Chune, Li (1989) J. Mat. Sci. 24,2951-2956.
Pink, E. and Campbell, J.D. (1974) Mat. Sci. Eng. 15, 187-194.
Pinner, S.H. (1956)1. Polym. Sci. 21, 153-157.
Plazek, D.J. and Choy, I.e. (1989) J. Polym. Sci., B. Polym. Phys. 27,307-324.
Plazek, D.J. and Frund, Z.N. (1990) J. Polym. Sci., B. Polym. Phys. 28,431-448.
Pogany, G.A. (1969) J. Mat. Sci. 4,405-409.
Pogany, G.A. (1970) Polymer 11,66-78.
Prime, R.B. (1981) In Thermal Characterization of Polymeric Materials Ch. 5. ed. Turi, E.
pp. 435-569 see p. 445ff.
Rabek, J .F. (1980a) Experimental Methods in Polymer Science John Wiley, Chichester, Ch. 32,
pp. 527-542. (1980b) Ch. 20, pp. 299-331. (1980c) Ch. 15, pp. 221-253. (1980d) Ch. 34,
pp. 549-581.
Rabinowich, E. (1937) Trans. Far. Soc. 33, 1225-1233.
Rashid, H. U. (1978) Ph.D Thesis, University of Sheffield.
Rekhson, S.M. (1986) In Glass: Science and Technology, Vol. 3, Ch. 1, ed. Uhlmann, D.R. and
Kreidl, N.J. Academic Press, Inc, Orlando, pp. 1-117.
Riccardi, C.C., Adahbo, H.E. and Williams, R.J.J. (1984)J. Appl. Polym. Sci. 29,2481-2492.
Riccardi, e.e. and Williams, R.J.J. (1986a) J. Appl. Polym. Sci. 32,3445-3456.
Riccardi, C.C. and Williams, R.J.J. (1986b) Polymer 27,913-920.
Rogers, M.G. (1972) J. Appl. Polym. Sci. 16,1953-1958.
Rozenberg, B.A. (1986) Adv. in Polym. Sci. 75, 113-165.
Sandford, W.M. and McCullough, R.L. (1990) J. Polym. Sci. (Pt. B) Polym. Phys. 2S,
973-1000.
Santore, M.M., Duran, R.S. and McKenna, G.B. (1991) Polymer 31,2377-2381.
Scherer, G.W. (1986) Relaxation in Glass and Composites, Ch. 12, John Wiley & Sons, New
York, pp. 162-174.
Selby, K. and Miller, L.E. (1975) J. Mat. Sci. 10,12-24.
Senturia, S.D. and Sheppard, N.F. (1986) Adv. in Polym. Sci. SO, 1-47.
116 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Serrano, D., Peyrelasse, J., Borred, C., Harran, D. and Monge, P. (1989) Macromol. Chem.
Macromol. Symp. 25,55--61.
Serrano, D., Peyrelasse, J., Borred, e., Harran, D. and Monge, P. (1990)J. Appl. Polym. Sci.
39, 679--693.
Sharma, D.K. and Soane, D.S. (1988) Macromolecules 21,700-710.
Shen, M.e. and Eisenberg (1966) Prog. Solid State Chem. 3, 407-481. See also Rubber
Chemistry and Technology 43,95-170.
Sheppard, N.F. and Senturia, S.D. (1986) Polym. Sci. and Eng. 26,354-357.
Smith, T.T. (1961) Polymer 2,95-108.
Sojka, S.A. and Moniz, W.B. (1976) J. Appl. Polym. Sci. 20, 1977-1982.
Sourour, S. and Kamal, M.R. (1976) Thermochimica Acta 14, 41-59.
Sperling, L.H. (1986) Introduction to Physical Polymer Science, Ch. 6, John Wiley, New York,
pp. 224-300.
Stanley, H.E. (1986) Enc. Polym. Sci. and Eng. 2nd edn, Vol 4, pp. 323-349.
Stauffer, D. (1985) Introduction to Percolation Theory Taylor and Francis, London (A 2nd edn
is scheduled for 1992).
Stauffer, D., Coniglio, A. and Adam, M. (1982) Adv. in Polym. Sci. 44, 103-158.
Steinmann, B. (1987) In Cross-Linked Epoxies eds Sedlacek, B. and Kahovec, J. Walter de
Gruyter & Co., Berlin, pp. 117-130.
Steinmann, B. (1989) J. Appl. Polym. Sci. 37, 1753-1776.
Stevens, G.C. and Richardson, M.J. (1983) Polymer 24,851-858.
Struik, L.e.E. (1978) Physical Aging in Amorphous and Other Materials. Elsevier,
Amsterdam.
Stutz, H., Illers, K.-H. and Mertes, J. (1990) J. Polym. Sci. (Pt. B) Polym. Phys. 28, 1483-
1498.
Stutz, H. (1992) Private communication.
Tanaka, Y. and Kakiushi, H. (1965) J. Polym. Sci. 3A, 3279-3299.
Tanaka, Y. and Mika, T.F. (1973) In Epoxy Resins, Chemistry and Technology Ch. 3, ed. May,
C.A. and Tanaka, Y. section IV, pp. 212-238.
Tobolsky, A.V. (1960) Properties and Structure of Polymer Wiley, New York, pp. 1-12.
Treloar, L.R.G. (1975) The Physics of Rubber Elasticity, 3rd edn, Oxford University Press,
London.
van Krevelen, D. W. and Hoftyzer, P.J. (1976) In Properties of Polymers, their estimation and
correlation with chemical structure, 2nd edn, Ch. 3, Elsevier, Amsterdam (see also van
Krevelen, D.W. (1990), 3rd edn).
Verchere, D., Sautereau, H., Pascauit, J.P., Riccardi, C.C., Moschier, S.M. and Williams,
R.J.J. (1990) J. Appl. Polym. Sci. 38,725-731.
Wang, X. and Gillham, 1.K. (1992) J. Coatings Tech. 64,37-45.
Ward, I.M. (1983) Mechanical Properties of Solid Polymers 2nd edn, Ch. 8. John Wiley,
Chichester, pp. 166--174.
Whorlow, R.W. (1980) Rheological Techniques Ellis Horwood Ltd, Chichester, pp. 123-191,
Appendix AI, p. 422.
Williams, G. and Watts, D.C. (1970) Trans. Far. Soc. 66,80-85.
Williams, J.G. (1979) J. Appl. Polym. Sci. 23,3433-3444.
Williams, J.G. (1980) Stress Analysis of Polymers 2nd edn. Ellis Horwood, Chichester,
pp. 142-143.
Wisanrakkit, G. and Gillham, J.K. (1990)J. Appl. Polym. Sci. 41,2885-2929.
Wisanrakkit, G., Gillham, J.K. and Enns, J.B. (1990) J. Appl. Polym. Sci. 41, 1895-1912.
Wolfe, S.V. and Tod, D.A. (1989) J. Macromol. Sci. Chem. A26(1), 249-272.
Yamini, S. and Young, R.J. (1980) J. Mat. Sci. 15,1814-1822.
Yu, W. and von Meerwall, D. (1990) Macromolecules 23,882-889.
Ziman, J.M. (1979) Models of Disorder Ch. 7, Cambridge University Press, Cambridge,
pp.246--285.
4 Additives and modifiers for epoxy resins*
S.J. SHAW
4.1 Introduction
4.2 Diluents
o
(§:r'CH-cn, 1\
4.3 Fillers
Apart from the resin and curative, fillers are possibly the most common
formulatory ingredient employed in the majority of epoxy formulations.
Literally hundreds of different filler types can and have been used to modify,
in one form or another, the properties and characteristics of epoxies, in
addition to reducing cost. Table 4.1 indicates in terms of advantages and
disadvantages, the various effects that filler incorporation can have.
Although fillers can be considered beneficial for many applications,
disadvantageous characteristics such as an increased density (and hence
weight) together with an increase in viscosity which is likely to influence the
processing behaviour of the formulation, obviously require serious
consideration.
Table 4.2 lists some of the more important particulate fillers which have
been employed in epoxy formulations. Although the list cannot be regarded
as exhaustive, it does indicate the range of characteristic modifications
which the various filler types shown can provide. It is of interest to briefly
consider some of these property modifications.
ADDITIVES AND MODIFIERS 121
o
@(CH-CH, 1\
CH-CH2
"I
Butadiene dioxide Divinylbenzene dioxide
(©-+ 3
Advantages Disadvantages
similar levels to that obtained with many metals. This can be advantageous
in electronic and structural adhesive bonding applications.
4.3.3 Shrinkage
When epoxies undergo polymerisation and cross-linking, shrinkage occurs
throughout the cure process which can be damaging in a number of
applications ranging from electronics to adhesive bonding. Although
generally regarded as superior to other thermosetting polymers such as
phenolics and polyesters, further reductions in shrinkage beyond those
found for most unformulated epoxies is frequently considered necessary.
This can often be achieved through the use of various fillers and flexibilising
ingredients. Filler incorporation reduces shrinkage by simple bulk
replacement of resin with an inert compound which does not participate in
the cross-linking process.
4.3.5 Viscosity
The incorporation of fillers into an epoxy invariably results in an increased
viscosity, the maximum loading of filler for most applications often being
restricted by the maximum permissible working viscosity. Generally fibrous
fillers exert greater viscosity-enhancing effects, on an equivalent weight
basis than particulates. With the latter filler, particle size usually exerts a
dominating effect with fillers of small particle size tending to increase
viscosity to a greater degree than corresponding fillers of greater particle
size. This can be attributed to the greater surface area of the former (Potter,
1970).
Some fillers, in particular various types of silica, are capable of exerting a
thixotropic effect which has been put to good use in structural adhesives and
formulations requiring anti-sag characteristics (Lewis, 1988).
ADDITIVES AND MODIFIERS 125
4.3.6 Toughness
The usefulness of epoxy resins for many applications is often limited by their
comparatively brittle nature and susceptibility to catastrophic failure.
Consequently numerous investigations have been conducted over many
years in an attempt to provide processes and procedures which would allow
significant improvements in toughness, achieved at minimum expense to
other important properties and parameters such as modulus and Tg •
Probably the most successful means by which this has been achieved is by a
process known as rubber-modification and the subject will be discussed in
more detail later (section 4.6).
A number of investigations have shown that the incorporation of
particulate fillers such as silica, glass micro spheres and alumina trihydrate
can increase the toughness of various epoxy formulations (Moloney et ai.,
1983, 1984, 1987; Spanoudakis and Young, 1984). Perhaps of major
advantage here is that this can be achieved with an improvement in modulus;
although rubber modification generally provides a more efficient means of
toughness enhancement, this is usually achieved at the expense of other
important properties such as modulus.
Variables such as particle size, particle size distribution, particle surface
chemistry and particulate volume fraction have been studied by various
researchers which have indicated the most important variables necessary for
optimum toughness enhancement.
A mechanism based upon the concept of 'crack pinning' (Figure 4.4) has
been proposed to explain the toughness improvements brought about by
particulate reinforcement (Lange, 1970; Evans, 1972; Green et ai., 1979).
- - - - - - - - - - - - - - crack breakaway
crack pinning
o CF~CF3
I I /0,
/~
CHo-CH-CHo-O-C
- - I
0 C-O-CHo-CH-CHo
I - -
CF3 CF 3
C"F2,,+i
Fluoroepoxy resin
o
1/
C;@:FOC'o
I - /
HO- y \
C
CF3 0
4.5.1 Plasticisers
Plasticisers of the type conventionally employed in vinyl type polymers such
as PVC have not achieved wide usage in epoxy technology. In addition to
being essentially incompatable with epoxies they also suffer from the
traditional disadvantage exhibited by inert plasticising materials in being
prone to separation from the base epoxy. Even in circumstances where these
disadvantages are not manifested, the extent of flexibilisation achieved is
generally slight in comparison to the more efficient reactive systems.
o
II
C-NH-(CHo)o -NH-(CH 2h-NH2
I -- -
(CHo) 7
I - 0
/CH II
CH 'CH -(CH2h -C -NH-(CH2h-NH-(CHo)2-NH2
1/ I ""
CH, /CH-CH 2 -CH=CH-(CHZ)4-CH3
C
I
( CHo)
I - 5
CH 3
Figure 4.6 Structure of a typical polyamide curing agentlflexibilising additive.
130 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Of all the approaches that have been considered and adopted in an attempt
to alleviate the brittle characteristics of epoxy resins, elastomeric
modification has possibly been the most successful, where quite dramatic
improvements in toughness have been achieved with, in many cases, only
modest reductions in other important properties. Due to the technological
importance with which elastomeric modification is now viewed, it will be of
interest and indeed importance to consider this subj ect in some detail. In this
account the various factors which have been shown to be of importance for
successful toughness enhancement, including factors such as elastomer
structure and chemistry, compatibility and morphology, will be discussed
together with an account of the main mechanism which has been proposed to
account for the effects of elastomeric modification.
(4.1)
where I1Hm and I1Sm are the molar enthalpy and entropy of mixing
respectively.
By considering a cohesive energy density approach to molecular
interaction, compatibility can, to a degree, be characterised by the solubility
parameter, 0, the value of which for a particular material determines the
tendency of its molecules to attract its own species in preference to those of
a dissimilar species. This approach can be expressed quantitatively in the
equation,
(4.2)
(4.3)
diameter) promoting crazing within the epoxy matrix and smaller particles
(approximately 0.01 J-lm) leading to shear deformation. Craze generation,
in their view, was seen to be the most efficient means of toughness
enhancement and they therefore proposed that the formation of relatively
large rubber particles was the key to maximum toughness. Crazing is now, of
course, almost universally recognised as not being a significant deformation
mechanism with epoxies. So in this respect alone this theory can be regarded
as flawed. However, in one major respect involving the formation of
particularly tough rubber-modified epoxies, the role of particle size and, in
particular, particle size distribution has been and indeed continues to be
emphasised. In the 1970s Riew et al.(1976) found that by incorporating
bisphenol A into a DGEBA/CTBN/piperidine system, substantial improve-
ments in toughness could be achieved over and above that obtained from
the non-bisphenol A formulation. They noted in particular that this highly
beneficial effect was accompanied by a bimodal particle size morphology,
which they regarded as responsible for the enhanced mechanical
performance. Since this early work, similar observations have been found by
other workers employing the same approach and indeed the bimodal
particle size concept has been utilised in commercial rubber-modified epoxy
formulations, where the incorporation of both low and high molecular
weight CTBN rubbers have been found to produce the required morphology
together with impressive fracture performance data. Although such
evidence would appear to strongly support the particle size protagonists, it is
important to recognise that the parameters of interest, volume fraction and
particle size, are extremely difficult to vary in magnitude independently of
each other. Thus, for many formulations, although bimodal particle size
distributions may exist and indeed exert an influence on the properties of the
epoxy, it is likely that volume fraction could also vary in a significant
manner.
Although a relatively small number of researchers would still appear to
view particle size and size distribution as providing a not insignificant
contribution to toughness enhancement, most now recognise that the critical
morphological factor is rubber volume fraction. Generally the greater the
volume fraction of the phase-separated elastomeric component the greater
the improvement in toughness. Although this will partly depend on the
amount of rubber initially added to the formulation, the other factors
described previously, e.g. rubber molecular weight and acrylonitrile
concentration, curing agent reactivity etc. will also have an important
influence. It is particularly worth while noting that numerous studies have
shown that rubber phase volume fraction usually exceeds the volume
fraction of added rubber thus providing an insight into the composite nature
of the rubber domains. An anomalous effect such as this can only be
resolved if it is assumed that the rubber particles have a not insignificant
concentration of epoxy, either on the molecular level or as sub-inclusions
dispersed in the elastomeric phase.
ADDITIVES AND MODIFIERS 137
polyethersulphone
o 0
t<l§r@q:@°©t>-ol
II II
II
0
l8J I CH
3 0
II n
polyetherimide
Figure 4.9 Basic structures of polyethersulphone and polyetherimide.
140 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
References
Bascom, W.D., Cottington, RL., Jones, R.L. and Peyser, P. (1975) J. Appl. Polym. Sci. 19,
2545.
Bucknall, C.B. and Partridge, LV. (1983) Polymer 24,639.
Bucknall, C.B. and Gilbert, A.H. (1989) Polymer 30,213.
Cerere, J.A., Hedrick, J.L. and McGrath, J.E. (1986) 31st SAMPE Symp. 583.
Chan, L.C., Gillham, J.K., Kinloch, A.J. and Shaw, S.J. (1984) In Rubber-modified Thermoset
Resins, eds. Riew, C.K. and Gillham, J.K., Adv. Chern. Ser. 208, Am. Chern. Soc.,
Washington.
Diamont, J. and Moulton, R.S. (1984) 29th Nat. SAMPE Symp. 29,422.
Evans, A.G. (1972) Phil. Mag. 26, 1327.
Fisher, C.M., Gilbert, R.D., Fornes, R.E. and Memory, J.D. (1985) J. Polym. Sci., Polym.
Chern. ed. 23, 2931.
Gledhill, RA., Shaw, S.J. and Tod, D.A. (1990) 1nt. J. Adhesion and Adhesives 10,192.
Goobich, J. and Marom, G. (1982) Polym. Eng. Sci. 22, 1052.
Goodier, J.N. (1933) J. Appl. Mech. 55,39.
Green, D.J., Nicholson, P.S. and Emberg, J.D. (1979). J. Mat. Sci. 14,1657.
Griffiths, J.R. (1982) Chern tech 290.
Hedrick, J.L., Yilgor, I., Wilkes, G.L. and McGrath, J.E. (1985) Polym. Bull. 13,201.
Hu, H.P., Gilbert, R.D. and Fornes, R.E. (1987) J. Polym. Sci. Polym. Chern. ed. 25, 1235.
Kelly, B.K., Gilbert, R.D., Fornes, RE. and Stannett, V.T. (1988) J. Polym. Sci., B: Polym.
Phys. 26, 1261.
Kinloch, A.J., Gledhill, RA. and Dukes, W.A. (1975) In Adhesion Science and Technology,
ed. Lee L.H., Plenum, New York.
Kinloch, A.J., Shaw, S.J. and Hunston, D.L. (1983) Polymer 24,1355.
Kinloch, A.J., Maxwell, D. and Young, R.J. (1985) J. Mat. Sci. 20,4169.
Kirshenbaum, S.L., Gazit, S. and Bell, J.P. (1984) In Rubber-modified Thermoset Resins, eds.
Riew, C.K. and Gillham, J.K. Adv. Chern. Ser. 208, Am. Chern. Soc., Washington.
ADDITIVES AND MODIFIERS 143
5.1 Introduction
800 •
N
E •
~ 600
::: •
. •
\!I 400 •
• • ••
• ,•
• • ••
...-
• ••
200
•• ••
0
0 SO 100 ISO 200 250 300
Figure 5.1 The wide range of epoxy properties and associated glass transition temperatures
presently available.
great challenges to epoxy chemists, that is, a tough system with a good high
temperature capability.
In this chapter the fracture bahaviour of epoxy resins will be considered in
some detail. Since most of the recent advances in the understanding of the
fracture behaviour of these materials have been achieved as a result of the
application of fracture mechanics principles, most particularly linear elastic
fracture mechanics (LEFM) considerable use of this approach will be made
here. Attention will be given to assessing the influence of structure (cross-
link density etc.) as well as the test conditions (temperature, rate, etc.) on
the basic fracture properties of epoxy resins. In the later part of the chapter
the various methods currently employed to enhance the relatively brittle
nature of epoxies will be presented.
The following section will present a brief summary of some of the more
important LEFM principles necessary for characterizing the behaviour of
these materials.
G = P~ Be
c 2B Ba
where e is the compliance of the cracked body for a given crack length a
and Pc the critical load for the onset of crack propagation. The above
equation represents the basis for most Gc calculations with the basic
requirement being a knowledge of the variation of specimen compliance
with crack length. The compliance-crack length dependency can be
determined either analytically or experimentally depending upon the
complexity of the test specimen.
It is important to note that all modes of loading encountered in
operational service can be decomposed into the three basic modes denoted
I, II and III presented in Figure 5.2. Mode I corresponds to simple cleavage
MODEm
MODEl MOOED
Figure 5.2 Modes of loading: mode I opening mode, mode II inplane shear, and mode III
antiplane shear.
FRACTURE BEHAVIOUR 147
or tensile opening, Mode II to in plane shear and Mode III to antiplane
shear. The fracture energies associated with I, II and III modes of
deformation are denoted GIc , G uc and G IIIc respectively. Mode I
deformation is technically the most important of the three since it represents
the most commonly encountered loading condition and generally has the
lowest associated fracture energy. Mode I-type loading is also much more
readily simulated in laboratory conditions . Typical mode I test specimens
are presented in Figure 5.3. Of these, the single edge notch (SEN) and single
/·/"--"'1iI
•
I 1 2W
•.
(a)
(b)
4W
(c)
(d)
Figure 5.3 Typical test geometries used for characterizing the mode I fracture toughness of
epoxies. (a) compact tension; (b) single edge notch; (c) single edge notch bend; (d) double
torsion.
148 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
o
1.5
;:;;l 0.5
Io
e Uniform Crack
Non·uniform Crack
I
0.0 .l..-~_-'--~_'--~_L..-~---'_~---'
0.0 0.2 0.4 0.6 0.8 1.0
2 '4
BD <I> (m x 10 )
Figure 5.4 Variation of energy U with geometric factor BD<P for Araldite B single edge notch
bend specimens. The influence of a non-uniform crack front is apparent.
FRACTURE BEHAVIOUR 149
value for a relatively brittle epoxy of this nature. The figure also highlights
one important point, that is that most epoxy resins are well suited to the
application of linear elastic fracture mechanics. However, when
undertaking fracture mechanics tests, it is very important that a straight
uniform pre-crack be obtained. Generally speaking, it is not difficult to
introduce shart pre-cracks into epoxy specimens. Difficulty is often
experienced, however, in assuring a straight and uniform crack front. The
open circles in Figure 5.4 correspond to tests on specimens in which the
crack front was not straight. From the figure it is clear that the non-uniform
cracks result in greater levels of scatter and an artificially high value of GIc-
When pre-cracking epoxy samples with a razor blade and hammer care
should be taken to ensure that the crack propagates as a result of one single
blow rather than a series which may result in a non-straight crack front.
G, =
c
Kic
E for plane stress conditions
150 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
and
~~r
+.. . . "
K
j
~..
j' '
,
(a) (b)
Figure 5.5 Shear bands in a hydantoin-based epoxy tested in tension at 85°C. (a) Just prior to
crack propagation and (b) after failure (Cantwell et aI., 1988),
152 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
=-.;
I- 16
""'
rIO
W"
"
Figure 5.6 Strain field at the tip of a sharp crack in 3501-6 epoxy resin determined by in situ
strain measurements in a scanning electron microscope (from Bradley, 1989).
The inelastic deformation mechanisms that ·take place at the tip of a sharp
crack in an epoxy resin will clearly have a significant effect upon the manner
in which the crack subsequently propagates. Many workers have made
considerable use of the double torsion (DT) test geometry in an attempt to
characterize both the toughness as well as the mode and stability of crack
propagation in epoxy resins (Young and Beaumont, 1976; Yamini and
Young, 1977; Phillips et al., 1978; Scott et al., 1980). Extensive testing of a
FRACTURE BEHAVIOUR 153
number of pure and rubber-modified epoxies has highlighted three basic
modes of crack propagation, these being:
(i) Stable brittle propagation
(ii) Unstable brittle propagation
(iii) Stable ductile propagation.
(a) DISPLACEMENT
- - INITIATION
- - ARREST
(b) DISPLACEMENT
Figure 5.7 Schematic representations of (a) stable crack and (b) unstable crack growth in a
double-torsion test specimen.
154 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
distance c ahead of the crack tip. Then, by combining the analysis for the
local crack-tip stress field at a distance, ahead of the crack tip (this being
equal to c):
1 + R.
o = ooya
--
,-------;<";;:-"
YY y(2,) ( 1 + ~, ) 312
where (! is the crack tip radius and 00 is the applied stress with the classical
fracture mechanics equation
it was shown that the critical intensity factor KJc could be related to the value
for a sharp crack KJcs by the relationship:
3/2
( 1 + R. )
2c
1+ R.
c
Figure 5.8 shows the variation of the relative fracture toughness parameter
KrJ KJcs with the square root of the crack tip radius for a series of specimens
containing natural cracks (open circles) and machined blunt cracks (closed
circles) (Kinloch et al., 1983). In the former, the crack tip radius was taken to
be equivalent to the calculated crack tip opening displacement.
10 •
Theory •
..::'" 8 '-
--....
~
~
u
6 •
•
(J tc =350 MPa
2
c = 0.7 /-lm
0
0 5 10 15 20 25 30
Figure 5.S Variation of KId KIcs as a function of the square root of the notch tip radius, (!, for a
DGEBA resin cured with 5 phr piperidine. The calculated values of the critical stress Ote and
critical distance care 350 MPa and 0.7 microns, respectively (after Kinloch et al., 1984).
156 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
This crack blunting analysis implies that a stress of up to five times the
value of the material's yield stress is attained at the crack tip. Kinloch (1985)
suggested that this critical stress, ate> could be interpreted as a constrained
yield stress.
Epoxies, like all polymers, exhibit a viscoelastic response when tested under
both short- and long-term loading conditions. Two fundamental test
parameters that highlight this pronounced viscoelastic behaviour are test
temperature and loading rate.
5.5.1 Temperature
Tests undertaken at different temperatures have shown that the mode I
fracture toughness of untoughened epoxies increases, sometimes
dramatically, with increasing test temperature (Bradley et al., 1992). This
increase is clearly related to the deformation process taking place at the
crack tip. Increasing the temperature has the effect of reducing the yield
stress of the polymer which in turn results in a larger plastic zone size and
greater crack blunting. Bradley (1991) suggested that the degree of
temperature-dependency depends upon the relative difference of the test
temperature, T, and the glass transition temperature, Tg, of the polymer in
question. Tests undertaken at temperatures considerably below the Tg of the
polymer should always, therefore, result in brittle failure. By plotting the
FRACTURE BEHAVIOUR 157
fracture toughness, values GIc (or G q if the plain strain requirements were
not satisfied) of his DGEBA epoxies as a function of T - Tg he obtained a
relatively unique curve with all the data falling within a narrow band (Figure
5.9). In all cases, the initial increase in toughness was found to occur at
approximately 125°C below Tg with a dramatic increase occurring at about
60°C below Tg • Unfortunately, no such relationship was obtained for the
rubber-toughened counterparts of these resins.
An interesting and, as yet, unexplained temperature-related effect was
observed by Scott et al. (1980) following low temperature tests on a series of
amine-cured epoxies. Whilst undertaking DT tests in a nitrogen gas
environment they observed that at temperatures below -100°C crack
propagation became unstable with the value of G 1i rising rapidly with further
decreasing temperature. The authors concluded that this change in failure
mode could not be attributed to a simple crack blunting argument but might,
instead, result from increased energy absorption by molecular relaxation as
well as an interaction of the crack tip with the gaseous envionment near its
liquefaction temperature. Thomson (1981) suggested that any frozen water
present on the surface of the specimen might be able to permeate and
plasticize the resin around the crack tip. Young (1980) pointed out that since
N2 has been shown to influence the failure mechanisms occurring in
thermoplastics, it could have a similar effect on epoxy resins. He suggested
that a critical test would be to undertake such tests in a helium gas
environment, a medium which does not have any observable effect on
thermoplastic polymers.
[DJ
4.-----~--~----~----.---~----~
t.
A
C
..
3 0 E
•
•
C-' 1
o
o t. flo
D6 t. 0
OL-____~__~~~_D_t.~~O~O__~__~____~
-300 -200 -100 o
2.2
t;jj
c •
••
:a t;jj
2.0
.'"
~
c
oS :a
~
1.8
•
0:::
oS
~
c
~ .;::= 1.4
.
.s 1.6
• •
~ e •
't:!-= ~1.2
OJ
" ••
...
• •
"
1.0
0.8
0 1 2 3 4 5 6 7
Log (Test time) (s)
Figure 5.10 Static fatigue tests on a tapered double cantilever beam showing the increase in
fracture toughness with loading time (after Gledhill et al., 1979).
FRACTURE BEHAVIOUR 159
data for both the unmodified resins and the toughened systems falling on
what appears to be a unique curve. The shift factors determined from these
viscoelastic analyses were then applied to the fracture toughness data
yielding a GJc master curve.
This type of data reduction technique is potentially very powerful since it
permits the toughness of an epoxy to be predicted for any given ratel
temperature combination.
IS 2.6
~
....
<Il
<lI
t:
eo:
2.4
~
OJ)
o 2.2
...:l
In a detailed study, Glad (1986) showed that the extension ratio, A, within
the plane stress damage zone in thin epoxy films could be directly correlated
to the maximum extension ratio of a single network strand, Amm by the
relationship:
A - 1 = 0.32(Amax - 1)
This appears to be the first time that the deformability of an individual epoxy
network strand has been related to the deformation process at the crack tip.
Fischer (1992) showed that the calculated crack tip opening displacement
could be directly related to the effective molecular mass between cross-
links. He also suggested that the width and length of the deformation zone at
the crack tip are directly proportional to the molecular mass of the network
strands.
Considerable debate has raged in the literature over many years regarding
the existence or non-existence of a very fine nodular structure within epoxy
resins (Racich and Koutsky, 1976; Morgan and O'Neal, 1977; Dusek et ai.,
1978; Mijovic and Koutsky, 1979). It has been suggested that due to the fact
that epoxies are normally made by the addition of a curing agent to a pre-
polymer, a two phase structure could occur (Morgan and O'Neal, 1977).
Mijovic and Koutsky (1979) argue that nodules correspond to high cross-
linked regions ranging in size from 6 nm to 10 Jim. Much of the evidence for
the existence of this fine second phase structure is based on observations
obtained from scanning electron micrographs and replication techniques. If
such a fine structure does indeed exist in these thermosetting polymers, it is
FRACTURE BEHAVIOUR 161
certainly conceivable that they would have a significant influence upon the
mechanical response of the material itself. Indeed, Mijovic and Koutsky
(1979) established a correlation between the morphology of a DETA-cured
DGEBA epoxy and their ultimate mechanical properties. They suggested
that the size of the nodules, resulting from the use of various curing-agent
concentrations, can be quantitatively related to the amount of plastic flow
occurring during crack initiation. Dusek et al. (1978) etched the fracture
surfaces of an amine-cured DGEBA and revealed globular structures some
20 to 40 nm in size. Similar structures were observed, however, following
etching of amorphous polystyrene and PMMA. Further, the small angle
X-ray scattering curves for the cured epoxy did not differ from those of
common amorphous polymers. The authors concluded that nodules do not
exist in simple amine-cured epoxies of this sort. They did point out,
however, that with more complicated systems a more pronounced
inhomogeneity might be caused by thermodynamic incompatibility or by
non-alternating mechanisms of the curing reaction. It is also possible that
the presence of such a secondary phase may be due to incomplete mixing of
the hardener and resin during manufacturing (Ellis, 1992).
three point bend (Stalder, 1985) and compact tension test geometries
(Cudre-Maroux, 1988).
In Figure 5.12 two single edge notch fracture surfaces are shown along
with the associated crack velocity profiles . Figure 5.12a shows a smooth
fracture surface with few apparent markings. The corresponding crack
velocity profile shows that the crack accelerated to 230 mls and then
stabilized. The second fracture surface, Figure 5.12b, is much rougher in
appearance due to extensive crack bifurcation. The velocity profile again
shows a region of crack acceleration, however, in this case the crack reaches
430 mis, the limiting velocity in this material. Crack branching and therefore
a very rough three dimensional fracture surface occurs only when the crack
a
i
500
i>- 400
~b
I-
300
U
0
...J
fa
...
\iOI 200
':Ie
U 100
«
1:11:
u
5 10 IS 20 25 30 J5
0 1 TA CE (mm)
b
t
Figure 5.12 Fracture surfaces of single edge notch Araldite B specimens showing the variation
of crack velocity during failure . Top: a smooth fracture surface; bottom: a rougher surface due
to extensive crack bifurcation.
FRACTURE BEHAVIOUR 163
has reached the maximum velocity (determined by the sonic velocity of the
material). Since any excess energy within the specimen cannot be dissipated
by further accelerating the crack, the crack must bifurcate creating multiple
fracture surfaces and greater energy dissipation.
Under certain conditions, other features may be prominent on the
fracture surface of fracture mechanics-type specimens. One example is
parabolic markings similar to those presented in Figure 5.13. Parabolic
markings are associated with the initiation of secondary cracks in front of the
primary crack in the region of fast unstable crack propagation. The precise
shape and form of such markings depend upon the relative velocities of the
primary and secondary crack fronts. If both fracture fronts move with the
same velocity a hyberbola is formed whereas an ellipse is created if the
primary crack propagates more rapidly than the secondary (Doll, 1989). In
this case, the distance between the vertices of the markings is between 50
and 100 microns. Following impact tests on notched PMMA samples, Van
Noort and Ellis (1984) observed a lower density of markings with the
average distance between vertices being approximately 200 microns.
The above observations have been used to explain certain macroscopic
features on the fracture surfaces of unmodified and particulate-filled
epoxies (Cantwell et al., 1990). Figure 5.14 shows the fracture surfaces of a
hydantoin-based epoxy tested at a number of temperatures between 23 and
105°C (Tg of the resin = 125°C). A detailed examination of the failed
surfaces highlighted four zones: a defect (not always in evidence), a smooth
mirror-like region, a slightly less smooth area and a rough three-dimensional
zone. The mirror-like zone was found to correspond to the region of slow
crack growth where the crack advanced at a speed of the order of several
millimetres per minute. Fine lines oriented parallel to the direction of crack
propagation are sometimes observed in this zone. These lines or 'river-
markings' occur as a result of crack propagation on two slightly different
Figure 5.13 Parabolic markings on the fracture surface of a DGEBA epoxy (photo courtesy of
Lu Fan).
164 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
(a)
(c)
Figure 5.14 Tensile fracture surfaces of a hydantoin-based epoxy resin (type XB 2900) cured
with a formulated anhydride hardener (HY 925) as a function of test temperature. Four regions
have been identified: (1) a defect (not always in evidence) (2) a slow growth mirror-like region
(3) a smooth region where the crack accelerates in an unstable manner and (4) a rough three
dimensional area. (a) 23°C; (b) 50°C; (c) 85°C; (d) lO5°C.
planes. The intersection of these two planes creates a welt frequently of the
order of tens of microns in diameter. River markings are useful in as much as
they indicate the direction of slow crack propagation. These lines can then
be traced back enabling the defect or cause of failure to be identified.
Beyond the mirror region extends another flat zone exhibiting parabolic
markings similar to those reported above. It was established that this region
corresponded to the zone over which the crack was accelerating in an
unstable manner. Once the crack achieved its limiting velocity (typically 400
to 500 mls in an epoxy tested at room temperature) and was therefore no
longer capable of accelerating, it bifurcated creating multiple fracture
surfaces and a rough fracture zone. The velocity of sound within this
material is approximately 1650 mls suggesting that the limiting crack
velocity is some 25-30% of this value. The rough zone was not observed in
the specimens tested at 85 and 105°e suggesting that the crack had either
stabilized at a velocity below the limiting value or else that the required
acceleration zone was larger than the specimen diameter.
FRACTURE BEHAVIOUR 165
The data presented in Figure 5.1 indicate that a large number of epoxies
offering an equally wide range of toughnesses are presently available.
Unfortunately, almost all of the epoxies for use in high temperature load-
bearing applications are brittle offering values of GIc of 300 J/m 2 or less. In
recent years, a considerable amount of work has been undertaken in an
attempt to enhance the toughness of these materials. These studies have
tended to concentrate on the addition of a second phase such as rubber
particles, thermoplastic particles or mineral fillers such as silica flour.
Modulus Kc Strength
Evans et al. (1985) suggested that substantial debonding of the second phase
glass particles may occur around the crack tip creating microcracks leading
to a reduction in the magnitude of the stresses in the process zone. A model
based on an idealized constitutive law was developed which predicted the
effect of filler volume fraction on toughness with some success. More recent
work (Smith, 1989) has suggested that the presence of these solid particles
results in a significant stress concentration in the surrounding matrix. At
high loads, the local stress exceeds the yield stress of the matrix resulting in
the formation of localized shear bands similar to those presented in Figure
5.15. The presence of such shear bands serves to blunt the propagating crack
and thereby enhance the measured toughness. Under conditions of extreme
crack tip ductility, such as at high temperatures and low strain rates,
extensive particle-matrix debonding may occur. Under these circumstances,
the presence of such a myriad of micro-cracks in a zone immediate to the
primary crack reduces the inherent toughness of the material resulting in a
significant drop in KIc (Cantwell et al., 1990).
Low (1990) showed that residual stresses playa significant role in the
failure process of filled epoxy systems. These residual stresses arise from the
thermal expansion mismatch between the epoxy and the dispersed phase.
He showed that compressive radial and tensile tangential stresses are
generated at the interface when the filler has a lower thermal expansion
coefficient that the matrix. It was indicated that the presence of such stresses
tends to enhance interface debonding and encourages catastrophic unstable
fracture resulting in a flat fracture surface.
Figure 5.15 Shear bands in a 16% silica-filled epoxy resin tested in tension at 85°C (courtesy of
Josh Smith).
FRACTURE BEHAVIOUR 167
2.5
...
!::'
eoi 2.0
0
Q.,
6 0
.'"'" 1.5
0
• •
...=
.c •
= 0
•
....
0
E-<
1.0 o CTBN
.a • PEI-PDMS
.....
Col
~
• PPO
0.5
0 5 10 15 20
Shear yielding
0.6
I
~
<l 0.4
0.0 .J......~____~_.-~--'-~----J'--~-'-~--.L_"----'
I~
Neat Resin
10% CTBN
I
~
4 • •
~~ •
~ 3
~
•
'-'
~
2 0
0
~
0
1 •
0
0
0
0 1000 2000 3000 4000
Resin Molecular Weight
Figure 5.18 Fracture toughness vs resin molecular weight for neat and CTBN-modified
DGEBA-DDS (after Levita, 1989).
d = r [( 2;rc )112 _ ~]
P P 3VF 2
Figure 5.19 The crack tip region of a silicalCTBN-modified epoxy resin tested at 105°C (photo
courtesy of Josh Smith).
172 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
5.9 Conclusions
Acknowlegement
WJC is grateful to the Fonds National Suisse for financing a sabbatical at the Czechoslovakian
Academy of Sciences, Prague, during which time part of the text was prepared.
References
6.1 Introduction
Epoxy resins are widely used for coating and encapsulating electrical circuit
components and electronic devices, where they serve to isolate a device
from adverse environmental effects of atmospheric gases, moisture, current
leakage, solvents, micro-organisms, mechanical shock and vibrations. In a
more recent application, known as micro-electronics packaging, they
provide the required insulation between closely-packed, fragile electrical
elements with good adhesion to both the substrate and the elements. It is this
application that now requires that the changes in the electrical properties of
epoxies during their curing and post-curing be understood at a fundamental
level, and that any effects of ageing on these properties be studied in enough
detail that procedures for eliminating these effects may be developed.
Although dielectric measurements for studying the progress of
polymerization were made as early as 1934 by Kienle and Race (1934), who
investigated polyesterification reactions, extensive literature on the subject
has appeared only since 1968. An excellent review of these studies up to 1986
has been given by Senturia and Sheppard (1986). For the reader's
convenience they have included, in this review, the details of measurement
methods and a comprehensive bibliography with 118 references on the
dielectric studies of various pure epoxies and their composites.
Since 1986, studies focused on dielectric monitoring of the progress of
chemical reactions during the processing of various epoxies have been made
by several groups (Day, 1986; Lane and Seferis, 1986; Bidstrup et aI., 1989;
Gatro and Yandrasits, 1989; Lane etal., 1989; Sheppard and Senturia, 1989;
Mangion, 1990; Kranbuehl, 1991; Parthun and Johari, 1992a; Alig and
Johari, 1992; Tombari and Johari, 1992). They have related the change in
the electrical properties to the increase in the viscosity of the epoxy resin,
thereby deducing that the increase in the glass transition temperature with
time allows modelling of the isothermal curing process - a subject of
considerable technical importance - by the Vogel-Fulcher-Tamman
equation. A collection of papers related to these observations has been
published in a special issue of Polymer Engineering and Science (vol. 29,
no. 5, 1989).
A somewhat different but more general treatment is based upon the
premise that in all addition reactions the growth in the size of a macro-
molecular addition product slows the rate of the very chemical reactions by
176 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
which it grows, and that this continues until neither chemical reactions nor
molecular diffusion occur at a perceptible rate. In this way, Johari and
coworkers (Johari and Mangion, 1991; Parthun and Johari, 1992a) prefer to
envisage the curing of an epoxy resin as a negative feedback process between
chemical reaction~-- and molecular diffusion, which ultimately and
irre¥ersibly vitrifies or solidifies a liquid. Thermodynamic evidence for this
vitrification is now available from the careful heat capacity measurements by
Cassettari et al. (1992); there is an irreversible, abrupt decrease in the heat
capacity of an epoxy resin over a narrow range of time, which remarkably
resembles the (reversible) decrease in heat capacity observed on
supercooling a liquid and rubbery polymer through their glass transition
temperature, Tg • The time at which this decrease occurs becomes less when
the reaction occurs at a higher temperature.
This chapter reviews the developments, since 1986, in our knowledge of
both the dielectric effects during the curing process of epoxy resins and the
electrical properties of their cured state. Much recent progress has come
from work on amine-cured epoxies, but many of the principles that have
been discovered relate equally to other reactions of linear chain and network
polymerizations, a subject of our continuing studies.
This chapter is organized such as to provide first (section 6.2) general
features of changes in the physical properties during the curing of a
thermoset polymer, in particular an epoxy resin. This is followed in section
6.3 by the description of a general formalism for the electrical properties of
an epoxy resin as they change with the curing time, temperature and
frequency used for the measurement. Section 6.4 provides the changes in the
dielectric permittivity, and section 6.5 those in the ionic conduction during
the process of sol-gel-glass conversion. The time and temperature evolution
of the dielectric properties is given in section 6.6 which is followed by the
chemical kinetics and dielectric behaviour in section 6.7. Sections 6.8 and
6.9 provide a description of the sub-Tg relaxations in partially-cured and
fully-cured epoxy resins and the effects of post-cure and physical ageing on
their electrical properties. A summary of the electrical applications of epoxy
resins is given in section 6.10.
(6.4)
and
(6.5)
where E'ion = (oo/weo) is the contribution to E" from dc conductivity. Here it
is clear that the permittivity of an epoxy resin during its curing process is
determined only by the dipolar orientation polarization but the loss is
determined by both the ionic or dc conduction and dipolar reorientation
process.
We first consider the change in the dc or ionic conductivity contribution,
E~o", during the curing of an epoxy resin. A number of the author's studies
on diamine-cured epoxies (Mangion and Johari, 1990c, 1991a, b, c; Johari
and Mangion, 1991; Parthun and Johari, 1992a, b) have shown that the
decrease in the dc conductivity of an epoxy resin during its curing is
described by a power law,
t (T) t (T)] x(T)
(J,(T t) = 0 (T t ~ 0) [ gel - c (6.6)
o 'c 0, c (T)
tgel
where oo(T, tc ~ 0) is the dc conductivity at the beginning of the curing
reaction at a temperature T, tgel is the time to reach the gelation point and
x is the critical exponent of the scaling equation used as a generalized
property function by Stauffer et al. (1982) and Djabourov (1988). (An
alternative equation oo(T, tc) = A exp (B/(to - t», which implies a
singularity at t = to, was used also by Johari and Mangion (1991), Mangion
and Johari (1991a, b, c) and Parthun and 10hari (1992a, b), but was found to
be a less satisfactory description of the measured conductivity than the
power law.) The significance of x in equation 6.6 lies in recognizing that for
a given time to reach gelation, its value determines the rate at which 00
decreases with time, i.e. the higher the value of x, the more rapid is the
decrease of 00 with the curing time. The description of the data by equation
6.6 allows extrapolation of the measured dc conductivity so that the time for
the gelation of an epoxy resin during both its isothermal and ramp-curing
can be estimated.
By combining equations 6.5 and 6.6, the dielectric loss during the curing of
180 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
where EO and Eoo are the limiting low- and high-frequency permittivities
respectively, t refers to the time for the decay of polarization for a chemical
structure that remains unchanged during the measurement time, or when
the curing reaction occurs too slowly in comparison with the time for
measurement. In equation 6.8 <l> represents a relaxation function whose
normalized form is given by
[<l>(t)]t, = exp [- (tlr)p]t, (6.9)
where f3 is an empirical parameter. For a Debye relaxation process, i.e. with
a single relaxation time, f3 is equal to unity. For the majority of viscous
liquids and solids, f3 is in the range zero to unity and here it is referred to as
Kohlrausch-Williams-Watts (KWW) function (Williams and Watts, 1970),
or stretched exponential parameter.
For a chemically and physically stable material, t and i in equations 6.8
and 6.9 have a strict meaning. It is the time for the observation of the decay
of polarization during which i, the average relaxation time, remains
invariant with time. For the curing process of a thermoset, i is not invariant
of time and therefore an approximation needs to be made. For frequencies
of 1 kHz or more which are often used in the studies of epoxy resins, t is of
the order of 1 ms or less and the effects from a change in both the chemical
structure and physical state of the epoxy resin during this period of 1 ms can
be justifiably assumed to be negligible at the usual curing temperatures
when the curing time is about one hour or more. This approximation allows
us to write equation 6.9 in a form used for a system with a time-invariant
chemical structure, as follows:
or
(6.12)
and
(6.13)
For a few epoxy resins whose 100 could be measured during the curing
process, Huraux and Sellaimia (1973), Sheppard and Senturia (1986),
Mangion and Johari (1991a,b) and Tombari and Johari (1992) have found
that the total decrease in 100 is by about 10% or less. This decrease therefore
has much less effect on 10* according to equation 6.8 in comparison with the
effect from the increase in r, which occurs by several orders of magnitude,
typically from 10- 9 s to 10+ 4 S. As a first approximation, therefore, consider
that both 100 and 10 remain constant with tc and equations 6.8 and 6.10 may be
00
combined to obtain
system such as the curing of epoxy resin is not easily measurable because of
the requirement that measurement be made as a function of decreasing
frequency till a plateau value at a low enough frequency is reached and that
during the period of measurement the chemical structure of the substance
remain unaltered. This requirement cannot be fulfilled for the curing of
epoxy resins. Therefore, it is necessary to devise an alternative procedure to
determine whether a fixed value of frequency used for the measurement is
low enough for e"(w, T, tc) to be equal to oo(T, tc)/weo within the
experimental error. We describe this procedure as follows.
The measured values of e' and e" of an epoxy resin may be transformed
into the complex electrical modulus, M*, formalisms by the equations:
M*(iw, T, tc) = (e*(iw, T, tc)r l = M'(w, T, tc) + iM"(w, T, tc) (6.16)
where
(6.17)
M' and Mil are the real and imaginary components of the complex electrical
modulus. When M* is entirely due to ionic conduction, it is related to
Maxwell or conductivity relaxation time, Ta , by the equation,
Since the rate of chemical reactions during the curing process of an epoxy
resin is sensitive to temperature, the change in the electrical properties with
both the temperature and time become important variables in the curing
process. Therefore, measurements are often carried out as a function of
ELECTRICAL PROPERTIES 183
both temperature and time. As a typical example of the behaviour observed,
measurement of the changes in electrical properties during the course of
curing of a stoichiometric mixture of diglycidyl ether of bisphenol A
(DGEBA) and hexamethylenediamine (HMDA) at six different
temperatures are shown in Figure 6.1, where the real and imaginary
components of dielectric permittivity and electrical modulus are plotted
against the logarithmic curing time.
For short periods of curing, E' slightly decreases first towards a plateau
value and thereafter in a stepwise manner to E' of about 4.4 as the curing
time, to approaches infinity. This step shifts towards longer tc as the
isothermal temperature for curing is decreased. The corresponding value of
E" first decreases from a near plateau value to reach a local minimum, which
is followed by a peak. As tc ~ 00, E" decreases to reach a limiting low value of
less than 0.02, which corresponds to that of a vitrified solid at a high
temperature. As the isothermal curing temperature is decreased, the
minimum becomes shallower, the peak becomes higher and both the
minimum and the peak shift towards longer curing times.
The corresponding plots of M' in Figure 6.1 show an increase in M' with
curing time, that occurs in two steps towards an ultimate value Moo. Both
steps shift to longer times with decrease in the curing temperature, and the
first step becomes smaller in height than the second step, while the
corresponding first plateau becomes broader on a logarithmic scale. The Mil
plots in Figure 6.1 show two peaks whose positions shift to longer tc and the
width of the first peak increases.
IO..---------,r-------,------, 0.3.----------,-------,------,
,.. ::.:~Y'
.: .
.:
'
Figure 6.1 The real and imaginary components of dielectric permittivity and electrical modulus
of the DGEBA cured with hexamethylenediamine measured for a fixed frequency of 1 kHz are
plotted against the curing time. The isothermal temperature for curing are: (1) 284.3 K,
(2) 296.5 K, (3) 304.2 K, (4) 312.3 K, (5) 324.3 K and (6) 336.5 K.
184 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
6
DGEBA-HMDA
E"
3
"''''''''''~ ,
21----=--
. ..:. . J
O~--~L-----~----~~--~~
3 5 7 9 11
E'
Figure 6.2 The complex plane plots of e* for the DGEBA cured with hexamethylenediamine
measured for a fixed frequency of 1 kHz during the isothermal cure at the same temperature as
in Figure 6.1. (li) calculated values from the parameters, t'1.e = 4.85 and y = 0.32 for (1);
4.35 and 0.32 for (2); 3.90 and 0.32 for (3); 3.48 and 0.31 for (4), 3.00 and 0.30 for (5) and 2.45
and 0.30 for (6).
ELECfRICAL PROPERTIES 185
DGEBA-EDA
0.3
343.3K
\.0,' ....
." " " - - - - - i
334.6K
'y.. .., . . . .
M"0.2
....
... 324.2K
\,..••••••••• ' ••••••=.- - - - - - I
r---."\......... .. . . . .
0.1
305.3K
DGEBA-DDM
-6
,..,
~
I
S -7
IJ1
---
0
b.O
.Sl
-8
-9L-----~------~--~--L-----~
in Figure 6.4. The curves calculated from equation 6.6 for values of ao(~ 0),
x and tgel have been shown (by the dashed line) to indicate the adequacy of
equation 6.6 as a description of curing kinetics.
Another description, which is an alternative to the power law of equation
6.6 has also been used by Mangion andlohari (1991a, b, c) and Parthun and
10hari (1992a). Its plausibility lies in recognizing that the increase in
viscosity during the early stages of the curing process determines the ionic
diffusion according to the Stokes-Einstein equation. This is particularly so
in view of 10hari and Pathmanathan's (1990) argument that the power law of
equation 6.6, when applied to temperature-dependence of relaxation time,
is equivalent to the well-known Vogel-Fulcher-Tamman equation over a
narrow range of measurements. This alternative equation implies an
approach of ao(t) towards a singularity during the curing process according
to
ao(T, te) = A(T) exp {- B(T)/[to(T) - tem (6.20)
where to is the point of singularity or the time taken to reach a value
ao(T, te) = 0 and A and B are temperature-dependent empirical parameters
which determine the rate at which conductivity approaches the singularity at
to. The data calculated from equation 6.20 are shown as a solid line in Figure
ELECTRICAL PROPERTIES 187
6.4. Mangion and 10hari (1990b, c, 1991a, b, c) and Parthun and 10hari
(1992a, b) have found that the value of to obtained from equation 6.20 is
close to the time of cure when the E" peak for 1 kHz frequency measurement
appears, and which in turn is closer to the vitrification time than to the
gelation time. In addition, Parthun and 10hari (1992a, b) have found that for
DGEBA cured with a variety of diamines, the value of tgel of equation 6.6 is
virtually independent of the frequency used for the measurement, whereas
that of to of equation 6.20 systematically changes. Figure 6.5 for the DGEBA
cured with ethylenediamine shows the variation of a(tc) measured at
-7~----~~----~-------r----~
... DGEBA-EDA
'e -g
"
00
~-1O
..s
-11 o
-7~------------~------------~
-8
...
'e -g
"
00
~-1O
..s
-11
-7~--------------~------------~~
...,
e -g
a.l
'6'
~-1O
-11
-12L-----~------~------L-----~
2 3 4
loglO (curing time/s)
Figure 6.5 The logarithmic plots of the measured conductivity against the curing time for the
DGEBA cured with ethylenediamine at 305 K. The frequency used for the measurement is
indicated. (0) data calculated from equation 6.6 or power law and full line from equation 6.20,
or singularity equation (after Parthun and Johari, 1992a.) Plot (1) is for measurements madefor
1 kHz frequency; plot (2) for 10 kHz; and plot (3) for 100 kHz frequency.
188 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
different frequencies against the curing time of the epoxy resin, where the
calculated tgel and to are marked by arrows.
The changes in £* of an epoxy resin during its curing, but after the dc
conductivity's contribution to £" has become negligible, are of some interest
here. These changes as seen in Figure 6.2 show that the fixed frequency
values of £' and £" of an epoxy resin during its curing resemble the £' and £"
plots of chemically and physically stable dipolar liquids and solids measured
at a fixed temperature but with varying frequency. This, as mentioned in
section 6.3, is a reflection of the fact that relaxation functions are invariant of
w or r as the choice of independent variable. Mangion and lohari (1990b, c)
have used equations 6.14 and 6.15 to calculate y, the curing parameter, by
using a modification of procedure given by Moynihan et al. (1973) and
Bendler and coworkers (Montroll and Bendler, 1984; Dishon et al., 1985;
Weiss et al., 1985), and Parthun and lohari (1992a,b) have calculated it by a
procedure given by Muzeau et al. (1991). The data calculated by Parthun
and lohari have been shown along with the experimental values in Figure
6.2. The agreement seen here demonstrates the adequacy of the formalism,
given in section 6.3, for the curing behaviour of epoxy resins.
Given that the measured dielectric data are satisfactorily described by
equation 6.15, it now becomes possible to calculate the relaxation time at
any instant during the curing of an epoxy resin. This is so because in the plots
of Figures 6.1 and 6.2, each data point corresponds to a unique value of
wr(te) in equation 6.14, which in turn corresponds to a unique value of N*.
Thus both the time-dependence of £* , and of r(te) can be calculated when !1£
and yare known. lohari and Mangion (1991) were first to show the adequacy
of such calculations, which was followed by the extensive work by Parthun
and lohari (1992a, b) who used a variety of curing agents to demonstrate the
variation of r(te) with the curing time of an epoxy resin. As a typical example
of this calculation taken from the work of the latter group, the calculated
values of dielectric relaxation time, and the permittivity and loss of the
DGEBA cured with propylenediamine have been plotted against the curing
time in Figure 6.6.
The range of relaxation time as determined by the above procedure is of
course limited by the insensitivity of N*(wr) to wr for a given value of yin
equation 6.15 but wr can also be varied by independently varying w which
then allows the determination of a wider range of r over a broader time scale
during the curing process. This is a remarkable feature in that it allows the
determination of virtually the entire change in r from picoseconds to
kseconds as an epoxy resin reaches its ultimately cured state. By using
frequencies of 50 Hz, 1 kHz and 100 kHz, Part hun and lohari (1992a, b)
have been able to obtain the curing time dependence of r of a number of
epoxy resins, a typical example of which is shown in Figure 6.7. The data
obtained from different frequencies of measurement overlap over a narrow
region of curing time, but all lie on the same curve which is now much
ELECTRICAL PROPERTIES 189
1r---------.---------,---------,
o 1
-1
-5
-6
-7~--------L---------~------~
10.--------.~~----~--------.
8
£'
·w
bO -1
..9
-2~------~--------~--------~
2 3 4 5
loglO (curing time/s)
Figure 6.6 The relaxation time, the permittivity and loss of the DGEBA cured with
propylenediamine measured for a fixed frequency of 1 kHz is plotted against the curing time
during their isothermal curing at (1) 294.9 K,(2)303.8 K,(3)315.5 K,(4)323.7 K,(5)335.9 K
and (6) 346.0 K. (0) calculated, and (e) measured, data.
DGEBA-PDA
-J
-5
-7
where ni is the concentration of ions with ionic charge Zi and mobility f-li' f-li is
often related to the inverse of viscosity according to Stokes' law and
assuming that ni remains unchanged during the curing process, a is regarded
as inversely proportional to the viscosity and its decrease to the increase in
the viscosity of the epoxy resin. This subject has been ably reviewed by
Senturia and Sheppard (1986) who have correctly pointed out that the
variation of a with viscosity is an oversimplification for an epoxy resin, for
here the viscosity approaches a formally infinite value at gelation, while
sufficient ionic and molecular degrees of freedom persist in the gelled state.
After a reanalysis ofthe work on this subject, they deduced that the decrease
ELECTRICAL PROPERTIES 191
in a with curing time is related to the difference (T - Tg), which itself
rapidly decreases as curing progresses. Thus after the maximum rate of
conversion which was taken to be at about 30-40%, the conductivity
becomes more sensitive to changes in Tg with the curing time. Its magnitude,
they concluded, provides no indication of the approach of gelation of epoxy
resin.
An alternative conclusion has come from the detailed studies of Mangion
and lohari (1990b, 1991a, b, c) and Parthun and lohari (1992a, b) who have
shown that the dc conductivity decreases according to a scaling law of
equation 6.6, and its value approaches zero at the gelation time of the epoxy
resin. However, this approach towards a zero value, they point out, is
interrupted by the contribution from ac conductivity which rises, with the
progress of cure, towards a peak value in a fixed frequency measurement.
They have argued that for the anticipated appearance of a singularity for a,
the time required to reach gelation should be independent of the frequency
used for the dielectric measurement, provided the change in the frequency
does not alter the relative contributions from other additional relaxation
processes. This is, of course, expected since the dc conductivity due to ionic
diffusion is invariant of the frequency for measurement, or E" inversely
scales with the frequency of measurement. They have therefore measured
the dielectric properties for a frequency of 50 Hz, 1 kHz, 10 kHz and
100 kHz as the curing ofDGEBA with ethylenediamine, propylenediamine
and hex am ethylenediamine progressed and have found that the gelation
time, tgeb obtained from equation 6.6 is independent of the frequency used
for the measurement.
A note of caution is necessary here, namely, that this new approach to
determining the time for gelation should not be regarded as undermining the
value of the now widely quoted (by Senturia and Sheppard, 1986) previous
efforts which concluded that no electrical 'events' accompany gelation, for
none does. The determination of gelation time by electrical measurement
described here relies on accurate determination of the dc conductivity and
the analysis of its variation with the curing time, which could not be done
until recently. It is also necessary to recognize the fact that the (event of)
approach towards formally a zero conductivity at gelation is prevented by
the significant contribution from ac or dipolar conductivity of the gelled
state. There is, of course, a need for confirmation of the value for gelation
time obtained from equation 6.6 by measurements using a suitable
mechanical relaxation procedure.
The fact that a single frequency measurement during the isothermal curing
can provide information on both the gelation time and the relaxation time is
192 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
I
(Tcure)-l
-16~------------------------------~
(temp)-l
Figure 6.8 An illustration ofthe change in relaxation time for the a- and p-relaxation processes
of the network structure at a given instant during the curing of an epoxy resin. For simplicity,
the plots for the p-process are drawn to have the same slope and to merge with the a-process at
a frequency of 107 Hz. The pre-exponent for all plots has been kept the same.
6r---------r---------,---------,18
DGEBA-PDA
4 12 "'"
'",..ro
x
>< "
,...,
0 i
-"'-
2"'-"'-'"
\.
'" o OQ"'"
<l e.
P?
2 6 ~
.~
0 0
275 300 325 350
T/K
Figure 6.9 The temperature-dependence of the dielectric parameters obtained from 1 kHz
measurement during the isothermal curing of DGEBA with propylenediamine.
jj.f = (f'(t -> 0) - f'(t -> 00)), y = curing parameter, tgel = gelation time and tpeak (f")
= time when the relaxation time of the epoxy resin is equal to 159 ms. (1) jj.f; (2) y; (3) t gel ;
(4) tpeak (f").
c'(t -':> 0). The decrease in tgel> and tpeak(c") with increasing temperature is
obviously a reflection of faster chemical reaction rates with increase in the
curing temperature, that convert the sol to a gel and gel to a vitreous state.
The decrease in y with increase in the curing temperature is unexpected in
that for chemically and physically stable substances the parameter for the
stretched exponential relaxation function, {3, approaches unity as the
temperature is increased. In this respect, the curing parameter, y, differs
from the K-W-W parameter, {3, and this difference justifies both its
distinction from {3, and its significance as a kinetic parameter for a kinetic
process. In technical terms, the decreasing value for y implies that c"(tc)
becomes more spread out during the curing period and that c" of the cured
sample is greater for a higher curing temperature than for a lower curing
temperature, if contributions from the {3-relaxation process remained
constant. Note that the use of y in equation 6.15 does not imply that its
Fourier transform into the frequency domain is meaningful.
Let us now consider whether a useful relation between the chemical kinetics
measured in terms of the extent of reaction and the electrical properties of
ELECfRICAL PROPERTIES 195
an epoxy resin can be found. Attempts for seeking such a relation require
that each of the seven effects noted in section 6.3 be considered, as both the
dynamic and static properties from near zero frequency to optical
frequencies are affected by the chemical reactions. Although in principle the
magnitude of these effects can be determined both theoretically and
experimentally, the effort required for doing so is enormous. So, as a
compromise, a procedure that includes only the most prominent effects can
provide a useful relation between the chemical kinetics and dielectric
properties of an epoxy resin, and these effects are changes in the
conductivity, a, and the average dipolar relaxation time, r, of an epoxy
resin.
A number of studies up to 1986, which have been ably reviewed by
Senturia and Sheppard (1986), have related the resistivity, or the inverse of
conductivity, to the extent of epoxide conversion, a. By assuming that a is
inversely proportional to log (a), they have provided a model for the curing
kinetics expressed by the equation,
(8 log a/8th = - K(1 - a)m (6.22)
where a is the reciprocal of resistivity measured for a fixed frequency, K is
the rate constant and m is the empirical reaction order parameter whose
value varies substantially depending upon the epoxy resin, the experimental
method and the temperature range.
In a fundamentally different approach Mangion and Johari (1991a, b)
have concluded that the dipolar relaxation time, rather than the
conductivity, is related to the reaction kinetics. Their studies have shown
that the dipolar relaxation time increases with the progress of reaction and
that the logarithmic plot of r against tc is sigmoidal in its shape up to r = 103_
104 s, beyond which measurement of r is prevented by the limit on the time
itself which is required for such measurements. Parthun and Johari (1992a)
have analytically examined the dependence of r on the curing time and have
shown that the shape of the plot of the extent of epoxy conversion against
logarithmic curing time is related to that of the corresponding plot of the
average dielectric relaxation time, instead of the resistivity of an epoxy resin
measured for a fixed frequency. This relation is given by an empirical
equation:
r(T, tc) = r(T, 0) exp [San(T, tJ], (6.23)
o~----------------------~
DGEBA-EDA
-2
-4
~
Ol
o
- -6
-8
•
•
-10~--~--~----~--~--~--~
o 0.2 0.4 0.6
(X
Figure 6.10 The relaxation time calculated from the dielectric behaviour of DGEBA cured
with ethylenediamine at 296.2 K (.) compared against the prediction from equation 6.22. Full
line is calculated from equation 6.23 with r(T,O) = 6 ns, r(T,ao) = 52 sand n = 1.95.
alters the shape of the sigmoidal plot so that reT, tgel ) can be made to
correspond with aCT, tgel ».
When n = 1, the extent of cure is directly
proportional to In (r(T, te».
The measurement in the MHz frequency range by Tombari and 10hari
(1992) has confirmed that for the curing of epoxy resin DGEBA with
ethylenediamine at 296.2 K, equation 6.23 or 6.24 provides a satisfactory
description of the calorimetrically measured extent of reaction and the
dipolar relaxation time with r(O) = 6.0 ns, r( 00) = 52 sand n = 1.95. A
typical plot showing such an agreement is given in Figure 6.10. There is a
need for further measurements of dielectric and calorimetric behaviours of
the epoxy for demonstrating the usefulness of this relation.
0.4
DGEBA-EDA
0.3
..-..
8
w
,
00.2
~
:--
-w
0.1
~
~
0
-4 -2 0 2 4
10g(<.O't)
Figure 6.11 The normalized plots of dielectric loss against frequency for the DGEBA-
ethylenediamine mixture at several fixed instants during its curing at 296.2 K. The instant of
measurement for each plot in ks is (e) 0.9; (+) 2.7; (D) 4.5; (.) 6.3; (0) 8.7; (l\.) 11.1; (83)
13.5; (lSI) 15.9.
198 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
peak, which is not shown here, would indicate that it is instead the a-process
which emerges and progressively separates from the {3-process as it shifts
towards a lower frequency. For clarity, these features are illustrated in
Figure 6.12, where the variation of the relaxation rates for the a- and
{3-processes is shown against the reaction time at a fixed temperature. The
plot illustrates a pattern of the development of relaxation processes and
their separation when an epoxy resin undergoes isothermal curing. As the
reaction proceeds, the relaxation process with a unimodal distribution of
times acquires a bimodal distribution of the a- and {3-relaxation processes,
each showing a peak for a fixed frequency of measurement. This feature of
splitting of the relaxation process, we note, bears a remarkable resemblance
to that observed on supercooling a liquid whose a-relaxation process also
progressively separates, as reversible molecular clustering raises the liquid's
viscosity, from the {3-relaxation process in a time-temperature plane (Johari,
1976).
Recent studies have shown an additional remarkable feature of the sub- Tg
relaxation in epoxy resins, namely the development of the slower or high
temperature {3-relaxation process at the expense of a faster or low
temperature, y-process. This is shown in Figure 6.13 where plots of the E"
against the temperature after different curing times are drawn. In this figure
/
/
Iu-process
/
/
-9
log(time/s)
Figure 6.12 An illustration of the change in relaxation time with increase in the reaction time of
an epoxy resin, as the a- and t3-processes evolve at a fixed temperature.
ELECTRICAL PROPERTIES 199
18 1 2
14 DGEBA-DDS
-w 10
2
I
0
~
-w
bD
.3 -1
-2
75 175 275 375
-w 4
X
'"0......
2
0
75 175 275
T/K
Figure 6.13 The dielectric loss of the sub- Tg relaxation regions (i.e. y and (3) and the a-relations
regions ofthe DGEBA cured with 4,4' -diaminodiphenyl sulphone. The bottom figure is for the
enlargement of the corresponding regions of the y- and f3-relaxation processes. Plots 1 to 6 in
both figures refer to an accumulated equivalent ageing time of 0.2, 1.7,3.5,5.6,9.0 and 15 ks,
respectively, and plots 7 to 9 to 24, 45 and 84 ks all at a temperature of 398 K.
500~--.---.---~--~----~--~--~---,
400 DGEBA...,...DDS
Q
300
200
100
0 II
0
.....
X 9
S
-w
7
l' .~Y
I~~~~~~----------~
1
2 3 4 5 6
log(AECT/s)
Figure 6.14 The peak value for f" for the a-, {3-, and y-relaxation processes during the curing
of the DGEBA with 4,4'-diaminodiphenyl sulphone (solid notations) and with 4,4'-
diaminodiphenyl methane (open notations) is plotted against the accumulated equivalent
curing time (log (AECT» at 398 K.
450 I I
2.0
DGEBA-DDk _0-
0_0
Tg 0_0 -
0- 0 -
425 0""
!
0
~
b"
E-< 400 l- 1.5 t>
t..c(ca!c) x 3 _--..... '"
~-~
/r,~~
r,/
375
I
t..c(meas) "\
I
----- 1 1 1.0
350
0 2
!oglO[ time/hi
Figure 6.15 Plots of the glass transition temperature, Tg , and dE = (Eo - Eoo), of DGEBA
cured with 4,4' -diaminodiphenylmethane against the time for isothermal ageing at 403 K. Data
for T (0) are taken from Plazek and Frund (1990). (L',) the dEvalues as measured, (.) the dE
g values as calculated. The scale for dE is on the right.
202 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
the end carbon atom of the epoxy molecule and the terminal atom of the
curing agent. The volume swept out by the hindered rotation of the OH
group about the C-O bond is much smaller than the volume swept out by
rotation of the epoxide-resin group about the newly formed C-N covalent
bond. Therefore, the increase in dE at the early stage of ageing in Figure
6.15 seems at first sight to be mainly due to an increase in the number of OH
groups formed as a result of chemical reactions, and in fact, the mechanical
relaxation peak corresponding to the j1-relaxation process in epoxy resins
observed by Ochi et al. (1987) has been attributed to the motion of hydroxyl
groups. However, a variety of studies on nonstoichiometric compositions
of cured epoxy resins by Mangion et al. (1992b) have shown that the height
of the dielectric j1-relaxation peak remains the same both for stoichio-
metrically starved (2 moles DGEBA-O.75 moles DDM) and saturated
(2 moles DGEBA-1.25 moles DDM) epoxy resin, although the number
of -OH groups formed on curing in the former case is much less than in the
latter case. This suggests that the dielectric j1-peak is not due to the motions
of only the -OH group in the network structure; other groups also move and
contribute to the j1-process.
Sidebottom and 10hari (1990) have calculated the increase in the dE of the
sub-Tg relaxation which is based on the premise that the number of OH
groups formed is equal to the number of cross-links in the network structure
and that the latter can be satisfactorily determined from the observed
increase in the glass transition temperature using the configurational
entropy theory by DiMarzio (1964). They found that the calculated dE is one
third the value evaluated from the experiment as shown in Figure 6.15.
Therefore, the formation of an OH group as a result of chemical reactions in
the amorphous solid state of a cured epoxy resin is only partly responsible
for the increase in dE. Other local dipolar segments also undergo rotational
diffusion. These dipolar segments are likely to be in the vicinity of the OH
group, but their exact chemical constitution cannot be unambiguously
identified by dielectric measurements alone. It is probable that the
formation of OH groups creates localized regions of relatively loose
molecular packing, where molecular or segmental diffusion persists.
One of the remarkable features of the change in the electrical properties
of cured epoxy resins is that when the Tg of the network structure approaches
a constant value, with increase in the ageing time, as is seen in Figure 6.15,
the dE of the sub-Tg relaxation process begins to decrease. This effect is
qualitatively similar to that observed for molecular liquids (Johari, 1973,
1976) and linear chain amorphous polymers (Pathmanathan et al., 1989;
Muzeau and Johari, 1990). It now seems sufficiently well-established that
the nature of changes in the electrical properties of cured epoxy resins over
a very long period of ageing is similar to that for other amorphous polymers,
although over a short period of ageing the nature of these changes in the two
cases contrast with each other.
ELECfRICAL PROPERTIES 203
Epoxy resins are now increasingly used in the various areas of electronics, a
use that is partly due to the versatility of the procedures by which they can be
adapted in a controlled manner for encapsulating, potting, thin-film coating
and embedding or packing electronic circuits or light emitting diodes, and
partly due to their suitable electrical properties. One of the earlier
monographs on the subject is Plastic Coatings for Electronics (Licari, 1970).
Since then a number of edited books (May, 1988; Lai, 1989; Seraphin et al.,
1989) containing articles on the subject of electrical applications of epoxy
resins have become available. Amongst these, perhaps the most
comprehensive account, with 190 references, is given by Lee (1988) in
chapter 9 of the second edition of Epoxy Resins, Chemistry and Technology
(ed. May, 1988). This may be consulted for details of the characteristics of
epoxy resins for electronic applications and procedures for their use.
Amongst the primary reasons for the use of epoxy resins in electronics is
their dielectric permittivity which should be 3-6 at ambient temperatures
and low frequencies, their dissipation factor, tan 6, of 10-3 to 10- for 60 to
1000 Hz frequency, their dielectric strength of about 120-180 kV/mm and
the volume resistivity of 1019_10 12 ohm/m, with relatively low sensitivity to
temperature. In addition, an epoxy resin should possess a number of other
physical properties in order to make it suitable for electrical and electronic
applications, such as good adhesion to metals and silicon, low permeability
Table 6.1 Typical epoxy resin compositions for 100 parts by weight of DGEBA epoxy resin
Low temperature potting Glass beads (18) Butyl glycidyl ether (2)
epoxy (0.125mmdia) Diethylamino propyl amine (2)
Flame-resistant epoxy Chlorendic anhydride (90)
Miniaturized circuits epoxy Allyl glycidyl ether (10)
Polysulphide polymer (10)
2,4,6-Tris-( dimethylamino-
methyl)phenol (10)
Low exotherm potting epoxy Silica (100) Phthalic anhydride (30)
Flexible potting epoxy Polysulphide polymer (50)
Triethylene tetramine (10)
Electrical encapsulating- Silica (5) clay (75) Diethylene triamine (12)
compound
Machinable compound Powdered aluminium (200) Allyl glycidyl ether (10)
Diethylene triamine (11)
High heat distortion Silica (100) m-Phenylene diamine (14)
(casting)
For low exotherm (casting) Powdered aluminium (80) Polysulphide polymer (25)
Pumice (70) Dimethylaminomethyl phenol
(10)
204 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Acknowledgements
I am grateful to Drs M.B.M. Mangion and E. Tombari and Mr M.G. Parthun for allowing me
here to include figures which have resulted from their research collaboration with me.
References
7.1 Introduction
* © British Crown Copyright 1992IDRA. Published with the permission of the Controller of
Her Majesty's Stationery Office.
EPOXY RESIN ADHESIVES 207
relatively easy, without the necessity for high applied pressures during
the bonding operation.
(v) They exhibit good wetting properties when applied to well prepared
surfaces and exhibit relatively low shrinkage during cure.
A study of the literature concerning epoxies reveals the vast number of
formulation variables which can be employed for many applications
including adhesives (Lee and Neville, 1967; Potter, 1970; May, 1988). Many
different types of epoxide resins exist ranging from the essentially
workhorse epoxy, the diglycidylether of bisphenol A (DGEBA) to the more
complex systems based upon tri- and tetrafunctional resins (see chapter 1).
Perhaps rather more daunting, a large number of curing agents can be
employed to convert the epoxy prepolymer to a crosslinked network, with
a large number of these being capable of use in adhesive formulations
(chapter 2). In addition, adhesive formulations can contain and benefit from
further additives including fillers, toughening agents, coupling agents etc.
(chapter 4), as well as being available in liquid, paste, film or carried film
form, thus demonstrating the wide variation in epoxy adhesive types
available for the potential user.
In this chapter, an attempt will be made to discuss the factors pertinent to
the successful use of epoxy-based adhesives. Within this context,
consideration will be given firstly to the theories of adhesion, i.e. why
materials stick together including the important aspect of substrate wetting.
Substrate surface pretreatments, generally considered vital for adhesive
bonding with all adhesive types, including epoxies, will be briefly discussed
together with an account of the test methods available for assessing
important mechanical properties. Formulation aspects pertinent to epoxy
adhesive applications will also be briefly reviewed together with accounts of
the properties and, of particular importance, durability characteristics.
Finally some examples of where epoxy-based adhesives have been
successfully employed will be discussed.
7.2.2 Wetting
Possibly one of the single most important factors likely to influence the
strength of an adhesive joint is the ability of the adhesive to wet and spread
spontaneously on the substrate surface. This most important characteristic
can be quantified in terms of the contact angle which a liquid (adhesive)
forms when placed in contact with a solid surface as shown in Figure 7.1.
Considered initially by Young (1805) he related contact angle to the surface
energies of the two contacting materials, as indicated in equation 7.1 which
has been given his name.
sotid
Figure 7.1 Contact angle 0, of a liquid (adhesive) in contact with a solid surface.
EPOXY RESIN ADHESIVES 211
Indeed with the latter, surface treatments designed to impart polar surface
characteristics as a means of substantially increasing surface free energy are
generally regarded as mandatory. Interestingly, the criterion demonstrated
by equation 7.3 does in fact suggest that the application of molten poly-
ethylene to a solid cross-linked epoxy should result in the thorough wetting
of the latter by the former since polyethylene is now acting as the adhesive
and the epoxy as the substrate. In other words the surface energetics are
now reversed and experimental studies have in fact confirmed the validity
of equation 7.3 in this case (Kinloch, 1980).
Although metals and metal oxides, in theory, are regarded as occupying
the high surface energy end of the spectrum, in practice reality can be
remote from this theoretical utopia. The very factors which promote wetting
and good adhesion in the presence of an adhesive, also promote the adsorp-
tion of atmospheric contaminants such as hydrocarbons and moisture which
in turn can convert the substrate surface from one of high surface energy to
one of low surface energy. The importance of this phenomenon has been
demonstrated by the work of Gledhill et al. (1977) who studied the effect of
relative humidity on the wettability of mild steel surfaces. They found, as
shown in Figure 7.2, that the presence of atmospheric moisture, even at low
relative humidities, was sufficient to bring about a dramatic reduction in
surface free energy from what would be expected for a totally dry,
uncontaminated surface (>500 mJ . m -2). In addition, increasing relative
humidity from 7 to 88% was shown to result in a gradually decreasing value
of Ysv, suggesting that the ability of the s!Irface to allow wetting by an epoxy
eo
\
70
60
..
~
j
50
>
~
40
300
15 50 75 100
RH (%)
Figure 7.2 Influence of relative humidity on the surface free energy, Ysv, of a grit-blasted mild
steel surface.
EPOXY RESIN ADHESIVES 213
adhesive would depend upon the relative humidity prevalent at the time of
bonding. Joint strength values obtained from mild steel substrates bonded
using an epoxy adhesive under two relative humidity environments
demonstrated this to be the case, with joints bonded in the lower humidity
environment exhibiting substantially greater strength values.
In addition to the thermodynamic factors discussed above, wetting as a
kinetic process is also of importance, since although surface energetic
considerations may indicate the potential establishment of intimate
molecular contact between an adhesive and substrate, the kinetics of wetting
may be the crucial parameter. For example, a rapid curing epoxy, or one
having a particularly high viscosity, may not be able to fulfil the wetting
potential implied solely by thermodynamic considerations.
Topographical factors can also influence wetting. Wenzel (1963) has
shown that surface roughness can exert a strong influence on the contact
angle an adhesive exhibits on that surface. He demonstrated this argument
mathematically thus:
cos ()' = r cos () (7.4)
where r is defined as a roughness factor given by the ratio of the actual area
to the projected area of the surface, and ()' and () are the contact angle values
for rough and smooth surfaces respectively. Thus, with a smooth surface,
exhibiting () less than 90°, roughening would result in a reduction in ()' with
an apparent improvement in wetting. However, if () were greater than 90°
then an increase in surface roughness would have the reverse effect, with ()'
appearing to increase still further.
Surface treatments can be divided into three broad areas, these being
associated with solvent cleaning, mechanical abrasion and chemical
treatment.
7.3.4 Primers
In addition to the surface pretreatments described above, the use of
substrate surface primers is frequently recommended for the adhesive
bonding of many metallic alloys. Primers can be employed to fulfil a number
of requirements. These can include:
(i) Protection. Freshly prepared metallic surfaces are prone to contami-
nation by atmospheric contaminants including water, hydrocarbons,
etc. The virtually immediate coating of the pretreated surface with a
primer can protect that surface from atmospheric contaminants by
basically exchanging a surface of very high free energy, and thus prone
to contamination, for one of more moderate surface free energy but
highly compatible with the adhesive. After priming, the surfaces are
less prone to contamination resulting in a substantial relaxation in
the necessary time interval between surface pretreatment and
bonding. Such primers are frequently solvent-based systems with
solids similar to the intended adhesive.
(ii) Wetting enhancement. Adhesive systems used for metal bonding are
often composed of films carried on cloth supports such as polyester
or nylon. Because of their precise formulation, film adhesives often
suffer from having only a short period of time at the appropriate
bonding temperature before cure inhibits flow in a system which
already has an inherently high viscosity. The application of a primer
solution to a metallic substrate prior to adhesive bonding can there-
fore overcome the potential wetting problems which would clearly
be associated with the use of such adhesives.
(iii) Corrosion inhibition. Corrosion-inhibiting primers, generally com-
prising an epoxy together with strontium chromate in a solvent
base, are generally recommended for use with aluminium alloys
following anodising pretreatments (Noland, 1976). This combined
218 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
The mechanical testing of structural adhesive joints can be divided into two
broad classifications namely (i) small coupon scale tests bonded and tested
to well-defined specifications and (ii) evaluation at a structural level with
either a complete structure or a sub-assembly being assessed. The former
approach provides a means of assessing various adhesive joint qualities
associated with material components, surface treatments and bonding
procedures, allowing the generation of large amounts of data at relatively
low cost. On the other hand structural level testing provides much more
limited information at greater cost and is generally employed as a means of
verifying design principles originally based, to a degree, on preliminary
coupon type tests. Since tests at the structural level would generally be
conducted on designs specific to a particular application, and therefore
unique to that application, this section will be devoted to the small, coupon
type test techniques which can be applied to epoxy adhesives in addition to
other generic types.
Before briefly describing some of the main test methods which have been
employed, it is of interest to mention the three main loading modes which
adhesive joints can encounter in service and which test methods are
designed to simulate. These are shown in Figure 7.3.
Although this approach is generally used in a fracture mechanics context
(see chapter 5), it is useful to employ it here. The opening mode I situation,
usually obtained in joints subjected to cleavage or peel, can be regarded as
the most harmful loading mode which an adhesive joint can encounter.
Conversely, mode II, an in-plane shear mode, represents the most desirable
EPOXY RESIN ADHESIVES 219
I I[
Figure 7.3 Modes of loading. I, tensile-opening; II, in-plane shear; III, tearing, antiplane
shear.
means of load transfer with most adhesives, adhesive joints being optimised
to maximise shear properties. Mode III, which represents a tearing mode of
failure has been little studied and will not be considered further.
(0)
Cb)
'" (c)
Figure 7.4 Typical tensile lap-shear joint geometries (a) single lap-shear joint; (b) double lap-
shear joint; (c) scarfjoint.
220 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
7.4.1.2 Peel. Peel testing of adhesive joints is probably second only to lap-
shear in popularity, this arising from both ease of manufacture and test. In
addition, although exhibiting certain disadvantages, peel tests yield
qualitative information regarding the ability of an adhesive joint to
withstand the most damaging applied stress which an adhesive joint may be
subjected to in service.
Three main types of joint configuration have been employed. These are
generally referred to as the T-peel, Bell peel and the climbing drum method
(Figure 7.5). The T-peel is probably the most widely used and is prepared by
bonding together two adherents of similar thickness which, after cure, are
pulled apart in the manner indicated. The Bell peel and climbing drum
techniques involve the peeling of an adherend at constant radius around a
steel roller. Both methods involve considerable energy absorption in the
adherends so that the peel values obtained from these two techniques can
be, and generally are, considerably higher than those obtained, for example,
from the T-peel where deformation in the adherends is considerably less.
Due to the difficulties inherent in assessing both applied and failure
stresses, results from peel tests are generally reported as linear values,
e.g .N. mm .
-1
(0)
(b)
Ie)
Figure 7.5 Typical peel joint configurations. (a) T-peel; (b) climbing drum; (c) Bell peel.
inhibits the generation of peel loads of the type found in other joint
configurations such as the standard lap-shear specimen.
(1920) quite simply proposes that fracture occurs when sufficient energy
(denoted by the term G) is released from the stress field by growth of the
crack to satisfy the requirements of the new fracture surfaces. The second,
proposed initially by Irwin (1964), is based on the premise that the stress
field at the crack tip can be defined by a parameter called the stress intensity
factor, K, and states that fracture occurs when the value of K exceeds a
critical value. Although both approaches have been employed in the study
of adhesive joint fracture, the energy balance approach has received most
attention.
In recent years several specimen geometries have been designed with a
view to measuring the fracture resistance of structural adhesives (Kinloch
and Shaw, 1981a); most studies being devoted to epoxy-based adhesive
systems. Some of these designs are indicated in Figure 7.7. Probably the
most popular specimen employed to assess mode I fracture energy, G'e>
has been the contoured double cantilever beam specimen. Designed and
initially employed by Mostoroy and Ripling (1966), this specimen is
contoured in such a way as to provide a constant compliance. Since at a given
applied load the value of fracture energy remains independent of crack
length, G can be readily determined without knowledge of crack location; a
feature which can be extremely useful where difficulty is encountered in
determining the precise location of the crack tip. A number of variations of
the above geometry have been designed and employed.
In addition to mode I, specimens have been developed for measuring
mode II (in-plane shear) and combined modes 1111 loads. With regards the
last, the scarf-joint (Figure 7.8) has been designed to impart a combination
of both mode I and mode II load conditions on the adhesive layer; the value
of () dictating the precise load combination (Trantina, 1972). The so-called
independently-loaded mixed-mode specimen, also shown in Figure 7.8, was
designed to allow independent application of cleavage and shear loads so as
to provide both modes I and II or a controlled combination of the two
(Bascom et al., 1977).
Mode III loading has only received minor attention with regards adhesive
joint fracture studies.
/ /
9 /
~
(c)
Figure 7.7 Typical mode I fracture mechanics adhesive joint specimens. (a) Parallel double
cantilever beam; (b) contoured double cantilever beam; (c) double-torsion.
Ca)
(b)
o
Figure 7.8 Typical mode II fracture mechanics adhesive joint specimens. (a) Scarfjoint;
(b) independently loaded mixed-mode joint.
-
u 150
L
300
250
2DO
120 2~--~3---4&""---..I.5-~---'6
~
u
E
U.l
5 10 15 20
(TBN Cone (phr)
Figure 7.11 Influence of elastomer incorporation (CTBN) on the fracture energy of a
piperidine cured DGEBA epoxy.
-
~ 150
~
.-
:J
a
L.. 1
~
E
~
~130
:J
t...J
170
2 " 3 fo. 5
Cure time (h)
Figure 7.12 Effect of cure conditions (cure temperature and time) on a piperidine-cured
CTBN-modified DGEBA epoxy.
Figure 7.13 Typical relationship between lap-shear strength (aluminium substrates) and test
temperature for three types of epoxy adhesive. (a) Rubber-modified; (b) one-part, general
purpose; (c) epoxy phenolic.
234 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
c;< _ ~IC,: o~
:1\
4
E
8 3
I
M: \
l')
~ : 0---0- (a)
o tm
/
o
L__ ~~~~~==~I~~~e~=-~~~==~====~.~=-~(b~)
1 2 3
Bond thickness, t ,(mm)
Figure 7.14 Influence of bond thickness on the adhesive fracture energy of a piperidine cured
DGEBA epoxy. (a) Rubber-modified system containing 15 phr CTBN; (b) unmodified.
236 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
(--to
Volume of plastic deformation
Maximum
Decreasing,j, Decreasin~
More festriction I Less constraint
Bond Thid<ness, t ~
Corrpcrative!y
t.«2Fy) High ?'('I)I!/I/1jj Low
Approx Equal
tc(»2ry) Almost Nil =@ To GIG Bulk
7.7.1 Introduction
A major problem with the use of structural adhesives concerns the adverse
effects which certain environments can have on the mechanical properties of
the joint. Notable amongst these include temperature extremes, stress and
radiation. However, the most severe problem has been associated with the
influence of atmospheric moisture. Experience over many years, gained
from both accelerated tests and long-term exposure to warm/moist
environments, has shown conclusively the potentially disastrous results
which can result from the use of adhesively-bonded components in moist
environments (Hockney, 1970, 1972, 1973; Butt and Cotter, 1976). To
demonstrate the potential magnitude of this effect, Figure 7.16 shows the
influence of water immersion at a temperature of 60°C on the strength of
mild steel butt-joints. The adhesive formulation employed was prepared
from a diglycidylether of bisphenol A (DGEBA) cured using a tertiary
amine curing agent. Clearly this environment has a detrimental effect on the
strength of this particular joint, 1500 hours immersion resulting in only a
15% retention of strength. Additionally, the immersion conditions
indicated were also found to influence locus of failure with a pronounced
change from cohesive within the adhesive layer to apparent interfacial
occurring on increasing immersion time. Thermodynamic considerations
can, in fact, be used to predict such interfacial behaviour as will be described
shortly.
In this necessarily limited account an attempt will be made first to
highlight the major parameters which, from experience, have been shown to
influence joint durability, second to discuss the mechanisms which have
been deemed responsible for the adverse effects of moisture and third to
consider the various approaches which have been considered in an attempt
to improve the environmental resistance of epoxy adhesive bonded joints.
Finally, a brief discussion concerning other environmental factors such as
elevated temperature performance will be given.
EPOXY RESIN ADHESIVES 239
40
-
CJ
~
~
CI
c:
QI
L..
+-
VI
+-
c:
·0 0
.-'
I
+-
+-
::J
CO
10
O~------~5~OO~------1~OO~O------~~~
Interface
been confirmed, at least in the case of epoxy/mild steel joints, using modern
surface specific analytical techniques such as Auger and X-ray
photoelectron spectroscopy (Gettings et al., 1977). The epoxy/CFRP
interface is of particular interest in that both W A and W AL exhibit positive
values indicating interface stability in both dry and wet environments.
Indeed, as mentioned above, this stability is reflected in the locus of failure
generally observed when CFRP joints are aged in hot/moist environments;
the CFRP substrate and/or the bulk adhesive generally being influenced by
water rather than the interface.
7.7.3.2 Metal oxide stability. Studies have shown that, with the surface
pretreatments that have been and are currently employed with aluminium
alloys, the surface oxide layer is generally regarded as the 'weak-link' in the
presence of moisture, as opposed to the interface and adhesive. As
described in section 7.3, the three usual methods for pretreating aluminium
are chromic-sulphuric acid etching, chromic acid anodising and phosphoric
acid anodising. Work conducted by Noland (1975) using X-ray photo-
electron spectroscopy, showed that the oxide produced on aluminium alloys
by the chromic-sulphuric acid etch was unstable in the presence of moisture
changing to a weaker hydrated form. Furthermore, Noland observed that
epoxy/aluminium joints pretreated by this method exhibited an apparent
interfacial failure when the joints were exposed to hot/humid conditions.
However, XPS showed that failure had occurred in fact in the weak hydrated
oxide layer. Although using the thermodynamic arguments previously
described, interfacial instability would be expected between aluminium
oxide and an epoxy adhesive, it has been suggested that the apparent
discrepancy between this thermodynamic prediction and the observed locus
of failure can be reconciled by the existence of a mechanically interlocking
component (Schmidt and Bell, 1986). The three surface treatments
mentioned, particularly the anodising techniques, have been shown to
produce surface oxide layers exhibiting topographical characteristics
intuitively consistent with the ability to promote mechanical interlocking.
EPOXY RESIN ADHESIVES 243
7.7.4 Approaches to improved durability
Two main approaches can be considered as a means of alleviating the
undesirable effects of atmospheric moisture with epoxy bonded metal
joints. These are:
(i) Improve the moisture resistance of the most vulnerable part of the
joint, i.e. the interfacial zone.
(ii) Employ epoxy adhesive formulations which exhibit hydrophobic
characteristics which would restrict the transport of large concen-
trations of water to the vulnerable areas of the joint (Shaw et al.,
1988).
Both of these approaches will now be considered.
specific attention have included the influence of solvent type, solution age,
film drying conditions as well as, of course, the chemical nature of the
coupling agent. As far as the last is concerned, by far the majority of studies
and indeed eventual use in service, has been devoted to one specific
organosilane, namely y-glycidoxypropyltrimethoxysilane, the structure of
which is shown below.
/0"
CH2-CH-CH 2-O-(CH2h-Si-(OCH3h
non-silane joint
Figure 7.17 Effect of organosilane solution age a'nd solvent type on the strength of epoxy
bonded mild steel butt-joints following immersion in water at 60°C for 1500 hours (aqueous
solution) and 1000 hours (ethanol-based solution).
EPOXY RESIN ADHESIVES 245
shows the effect of two of the variables mentioned, namely solvent type and
solution age on the durability characteristics of mild steel butt-joints bonded
using a simple tertiary amine cured epoxy adhesive. Although rather
complex, this figure demonstrates several interesting features.
First, for joints pretreated prior to bonding with a 1% v/v aqueous
solution, the substantial improvement in moisture resistance over joints not
treated with organosilane is clear; the treated joints exhibiting joint strength
values of between 17.5 and 34.5 MPa following 1500 hours water immersion
at 60°C, in comparison with 5.8 MPa for the untreated joints. Second, a
maximum in joint strength is apparent over aqueous solution ages of
between 30 and 90 minutes. Third, changing from an aqueous to an ethanol-
based solvent system results in striking differences in behaviour. These can
be itemised as follows:
(a) The ethanol based solutions exhibit no apparent dependence upon
solution age.
(b) Joint strength values obtained from the ethanol system are sub-
stantially lower than their aqueous solution based counterparts.
(c) Joint strength values obtained from the ethanolic organosilane
solutions are virtually identical to those obtained from the untreated
joint.
Although such studies have provided clear evidence of the substantial
improvements in moisture resistance which can be conveyed to epoxy-
bonded joints by organosilanes it is clear that various experimental
parameters are of vital importance so as to avoid misleading and
pessimistically false impressions of their durability-enhancing charac-
teristics. Our inability to predict precisely the appropriate optimum
molecular and experimental parameters can be attributed, partly at least, to
the fact that at present no universally acceptable mechanism has been
proposed to describe their action. However, one particular theory, the
chemical bonding theory, has generally received the most widespread
support amongst adhesion specialists and thus a brief description here would
not be out of place.
The chemical bonding theory is the oldest, best known and possibly the
most easily understood of all the theories that have been proposed to
account for coupling agent action (Plueddemann, 1982). The concept is
strikingly simple, namely that in place of the predominantly secondary force
interactions that would exist between a high energy substrate, such as a
metal oxide, and an epoxy adhesive, a suitably employed coupling agent
would act as a 'covalent bridge' linking together the organic and inorganic
components of the joint. This covalent linking would be expected to provide
far greater moisture resistance than mere secondary force interactions
alone. The sequence of reactions which could be expected in the formation
of an interfacial 'covalent bridge', starting from the initial organosilane, can
246 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
R
I metal oxide
H 0 - Si-O H + ""'("""'"
()H OH OH
metal oxide
~
oI OH I
0
I
OH
I
R-Si-O-Si-O-Si-O-Si-R
I I I I
ORR 0
1. I.
-SI- -SI-
I I
1h Sh 24h 96h
CAE, chromic acid etch; CAA, chromic acid anodise; GB, grit
blast.
(Il)
(b)
c
(d)
o 100 1511
Exposure time (hI
Figure 7.18 Influence of hot/humid atmosphere (50°C/96% Rh) on wedge test crack growth
characteristics of epoxy bonded titanium joints. (a) Grit blast; (b) alkaline/hydrogen peroxide
etch; (c) chromic acid/fluoride anodise; (d) sodium hydroxide anodise.
which when cured with suitable curing agents, resulted in equilibrium water
concentrations as low as 0.2%. With regard to non-halogenated epoxies, the
amount of water the cured adhesive will absorb will depend upon many
factors including the type of resin and curing agent employed. In an
excellent review, Wright (1981) has shown that water uptake may differ by a
factor of ten between resin types and by a factor of three with a similar resin
but different curative. In fact the amount of water absorbed by a cured
epoxy adhesive can vary enormously from about 1% to 10% with 2-6%
being fairly typical for a wide range of formulations. Fluoroepoxies of the
type demonstrated above can therefore be considered as substantial
improvements over the majority of 'conventional' epoxy systems with
regard to hydrophobic character.
Further analysis of these systems, so as to determine whether the
EPOXY RESIN ADHESIVES 251
0
~ \0
\
!;.
£
@'
III
~70
IQI
0-
0
\0
\
f
~0
L.
'610
IS 0
1~
10 \ 0,--
0-
0 100 1 200 250
Test temperature (0 [ )
Figure 7.19 High temperature strength retention of a dicyandiamine cured, rubber-modified
epoxy adhesive.
252 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
greatly diminished with increasing temperature, with less than 10% room
temperature strength being retained at 140°C. This type of strength decline
can usually be related to the onset of Tg. Indeed Tg can, to a first
approximation, be considered as the high temperature limit for any
structural adhesive.
Improvements in high temperature capability can be obtained by
increasing the degree of cross-linking by a suitable choice of epoxy and/or
curing agent. The use of polyfunctional epoxies and curing agents can
produce cured products having Tg values in excess of 200°C, resulting in
substantial retention of properties at temperatures in excess of 150°C.
However, the propensity of epoxies to absorb, in some cases, considerable
quantities of water, which in turn can reduce Tg , together with the somewhat
brittle characteristics bestowed by the high cross-link density, has greatly
diminished this approach to high temperature capability. In addition, due to
their molecular structure, even the most highly cross-linked formulated
epoxy adhesives are unable to withstand long-term use at temperatures in
excess of approximately 175°C. For such applications other adhesive types
must be considered, including for example, the epoxy-phenolic-based
adhesives discussed in chapter 6 (section 6.2).
7.8 Applications
From humble beginnings in the 1940s and 1950s, epoxy adhesives are
nowadays employed in a diverse range of applications associated with a
number of important industries including aerospace, automobile, civil
engineering and electrical/electronics including of course the domestic home
market. Possibly the main impetus for their current popularity has been the
aircraft industry which has spearheaded both the development and
application of structural adhesives generally. Indeed, structural adhesives
were first used in aircraft construction in the early 1940s with a vinyl-
phenolic adhesive system being used to bond aluminium components on the
de Havilland Hornet aircraft. Since this time various adhesive types have
been used for a wide range of applications in both civil and military aircraft
(Alberici, 1983), with epoxy adhesives now enjoying prominence at the
expense of the older phenolic types.
Aircraft applications to date have included the bonding of reinforcing
doublers and stiffeners to both fuselage and wing panels; the former being
employed to reinforce holes for mechanical fasteners and window/door
openings. In addition, epoxy film adhesives are frequently employed for the
production of honeycomb sandwich structures where aluminium or Nomex
honeycomb is bonded to aluminium or composite skins. Such structures
have been employed to produce aircraft structures such as rudders, flaps,
elevators, ailerons, doors, floor panels as well as engine structures such as
EPOXY RESIN ADHESIVES 253
References
Anderson, G.P., Bennett, S.J. and De Vries, K.L. (1977) Analysis and Testing of Adhesive
Bonds, Academic Press, New York.
Adams, R.D. and Wake, W.e. (1986) Structural Adhesive Joints in Engineering, Applied
Science, London.
Ahearn, J.S., Davies, G.D., Sun, T.S. and Venables, J.D. (1983) Adhesion Aspects of
Polymeric Coatings, ed. Mittal, K.L. Plenum, New York.
Albericci, P. (1983) Durability of Structural Adhesives, ed. Kinloch, A.J. Applied Science,
London.
Allen, K.W. and Alsalim, H.S. (1977) J. Adhesion 8,183.
254 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
8.1 Introduction
Relative density Young's modulus Tensile strength Failure strain Fibre diameter
(GPa) (GPa) (%) (urn)
in particular, together with the high performance polymeric fibres offer the
best solution.
Where light weight is more important than the cost of materials, such as in
aerospace and sports goods applications, these reinforcements dominate.
However, where economic considerations are more important, in industrial
applications, glass fibres still have much potential. For the polymeric fibres,
• Kevlar 149
200
.
Asbestos (crocidolile)
~
.c 150-
Kevlar 49.
~:' '"
~ /
[ll"
Carbon fibre
5>
c
i"
• S glass Boron (filament
on tungsten). \
;;;
• Si·Ti·C·O
~HM
~c
·S,·C·O
2 100
u
"
0.
(f)
• Nylon
• E glass
50
•
.Polyester .
• Steel (wire)
Acrylic
.
~b.~.~ Alumina
o 5 10
Figure 8.1 Specific properties of fibre reinforcements. XA, 1M, HM, refer to differing grades
of carbon fibre. For other definitions see Table 8.1.
COMPOSITE MATERIALS 259
PBT is still only available in sample quantities for strategic applications and
so aramid fibres such as Kevlar, Nomex and Twaron are the main
contenders. However, differing fibre and composite performance are more
important criteria for the selection of the grade of aramid or carbon
reinforcement.
It is important to recognise that the strength of a material is determined
by the presence of Griffith flaws which reduces the theoretical strength
significantly. The fibre form reduces the preponderance of flaws so that
filament strength is larger than that of the bulk material. However, the
strength of a filament is still determined by the presence of defects and as a
consequence the reported values are gauge length-dependent and are
usually average values determined from either unimpregnated or resin-
impregnated bundles. Individual filaments within the bundle will have a
different strength so that the statistics of fibre strength contributes
significantly to damage mechanisms of the composite. A typical distribution
of filament strengths is given in Figure 8.2. A further consequence of the
distribution of flaws on the fibre surfaces is the need for protection by a resin
coating or so-called size during handling immediately after manufacture.
For this application emulsified epoxy resins are commonly employed for
carbon and glass fibres.
0.8
0.7
5 0.11
I 0.11
OA ·
0.3
Q.2
0.1
0.0 +----I~-r_____,r__--,--,----r---,.--_.__-._-_r_-_r___;
~ 1~ 1~ 1~ 1~ ~ ~ ~ ~ ~ ~ ~ ~
SlREHGnt /0'"
Figure 8.2 The probability of failure of single E-glass filaments under tensile stress showing the
statistical aspects of strength.
260 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
moo,,I
1500
C1l
D-
c:l
r:
c;,
:1 100
I
I
C1l
D-
I
filtered-clean c:l
c
---- --1300
t:
e! I
V)
::J
1ii strength -S
.!!i "0
0
'iii
c ~
<l>
l- normal
/
/
I I
1000 2000 3000
Heat treatment temperature °C
Figure 8.3 The effect of heat treatment temperature and clean room conditions on the
strength and modulus of PAN-based carbon fibres. (Redrawn from data in Watt, 1985).
COMPOSITE MATERIALS 261
room conditions have reduced the number of strength-reducing flaws. A
wide range of commercial fibres are now available of which the type-A fibre,
which has been heated to only 1000°C is often preferred. More recently,
intermediate modulus (1M) fibres have become available which combine
high strength with an improved modulus. A modulus approaching 800 GPa
(i.e. 80% theoretical) can be produced by hot-stretching techniques.
The microstructure of the carbon fibres is determined by the relatively
small size of the graphite nuclei and the curved ribbon-like molecular
structure of the oxidized acrylic fibre which yields a turbostratic graphite. A
skin-core structure is most pronounced for the high modulus fibres, as a
result of graphitisation at the high temperatures involved and the subtle
chemical differences in oxidised PAN fibre which arise from the diffusion-
controlled oxidation of the thermally cyclised polymer.
Surface oxidation is an essential post-carbonisation stage. This process can
be carried out chemically, thermally, or electrolytically, with the last being
the favoured method. The treatment is essential to the optimisation of the
adhesion of the fibre to the resin matrix and the performance of the
composite. If the surface treatment is too severe the composite material will
behave like a brittle monolith. Figure 8.4 demonstrates the need for a
carefully controlled surface oxidation to optimise the performance of the
1.5 ~--------------......,
(c)
0.5 I
100 300
DFT/%
8.2.1.2 Sizing resins for carbon fibres. To protect the fibres from damage
during handling and subsequent fabrication the fibres are dressed with an
epoxy resin size. This is usually applied from an emulsion, immediately after
the surface treatment. The role of these resins in the formation of the
interfacial bond has not been studied in detail but it is now recognised that
COMPOSITE MATERIALS 263
these may not have the same chemical structure as the resin matrix and the
presence of an interphase region needs to be considered.
foc-@-LH-@-NHt foSQ>-tNH-@NHf
(1) (2)
carbon fibres in that liquid crystal or nematic spinning dopes are employed.
The fibres are wet spun from 99.8% H 2S0 4 at 90°C, into an aqueous
coagulant bath and washed with a weak alkali to remove acidic impurities.
Residual sodium sulphate is considered to be responsible for the observed
variations in fibre-moisture absorption (3 ± 0.5%). The final morphology is
achieved by a subsequent but brief heat treatment for a few seconds in an
inert atmosphere at temperatures between 150 and 550°C when
recrystallization occurs. For high-performance, heat treatments at
temperatures of the order of 450°C lead to a perfection in the lateral order of
the polymer.
The spinning conditions result in a radial crystallite structure similar to
pitch-based carbon fibres. Because of the conformational arrangement of
the polymer chains, a pleated structure is formed.
The lower-modulus, 1,4-aramid with less lateral order is primarily used
for tyre cord and ropes. The higher modulus fibres which have a more
perfect crystal structure are used for composites. The irregular chain
conformation of the 1,3-aramid is not suitable for high stiffness, and has a
lower modulus. These are generally used as staple fibres for honeycomb
applications.
The highly aromatic molecular structure gives rise to high thermal stability
to the polymer. However, ultraviolet (UV) light can sensitise the oxidative
degradation process. Although the degradation products are self-screening,
it is essential, in applications where UV environments are encountered, to
employ protective films.
The polar nature of the molecule means that the equilibrium moisture
content is significant and steps are required to dry the fibres and to prevent
moisture ingress during use.
The aligned crystalline morphology is responsible for the poor fibre
transverse strength and the tendency to fibrillate. Consequently, the fibres
have poor compressive strengths. Composites also have relatively poor
transverse properties because of the limited adhesion chemistry between
fibre and matrix and between fibre crystallite planes. These properties can
be utilised in applications where impact resistance is critical, such as ballistic
protection.
Attempts have been made to improve the fibre-resin adhesion by using
plasma-type surface treatments and increase fibre integrity by y-irradiation
cross-linking, but commercially-treated fibres are not generally available
(Verpoest, 1987).
COMPOSITE MATERIALS 265
8.2.3 Glass fibres (Loewenstein, 1983; Jones, 1989)
A number of glass compositions are commonly used for the production of
glass fibres. The most important is E-glass, whereas for specialist
applications others can be used: for example, S-glass where a higher
modulus is required, ECR glass for acid resistance, and AR glass for alkali
resistance (i.e. cement matrices). C-glass is usually produced only in the
form of veil for chemically-resistant GRP linings.
The fibres (other than C-glass veil) are prepared by drawing from the
bushing below a molten glass reservoir. Since the strength of glass fibres is
determined by the distribution and size of surface defects the most
important and commercially sensitive stage in the process is the surface
treatment or size which is applied by rubber roller prior to assembly into
rovings. The rovings consist of strands, each containing 2000 filaments. The
rovings can be either used directly, rewound onto a bobbin (a small twist
being put in to improve handling) for further textile processing, or chopped
into short fibres for the production of chopped strand mat.
The surface 'finish' is a critical stage in the process, ensuring compatibility
with the resin or matrix and high-speed conversion into composite materials
with minimum catenation or fluffing of the strands, fibre breaks, and
damage. The 'finish' consists of an adhesion promoter called a coupling
agent, a protective polymeric coating or size, binder, and other additives
(e. g. fatty acid ester or amide) which ensure good handle ability . It is applied
in an aqueous medium and for epoxy resin matrices this will invariably be an
emulsified epoxy resin. Some manufacturers use polyester/epoxy resin film
formers for compatibility with a range of polyester and epoxy resins. Fibrous
mats or reinforcing tissues will have an additional polymeric binder, which
can be selected to control the wet-out of the fibres in the chosen fabrication
process.
The adhesion promoter which is generally a silane coupling agent, is
expected to generate a strong interfacial bond between the fibre and matrix.
They are generally of the form
R-(CHzkSi(OR'h
Vinyl CH 2=CHSi(OCH 3 h
/0"
*Epoxy CH2CHCH20CH2CH2CH2Si( OCH 3)3
CH 3
I
Methacrylate CH 2= C-COOCH 2CH 2CH 2Si(OCH 3 h
*Primary amine H 2NCH 2CH 2CH 2Si( OCH 3 h
*Diamine H2NCH2CH2NHCH2CH2CH2Si( OCH 3 h
*Mercapto HSCH 2CH 2CH 2Si(OCH3)3
Cationic styryl
Cationic methacrylate
Cycloaliphatic epoxide
1. Spray-up Chopped rovings (G) 15-30 Direct application of chopped Pipe linings
rovings and resin
2. Autoclave Continuous (G , C, A) 60 Forming of stacked prepreg in Aerospace components
vacuum bag
3. Filament winding Continuous rovings, -50 Wet resin carried by fibres onto Pipes and containers, geodesic
occasionally particulate mandrel at predetermined angles. shapes
fillers included (e.g. sand) Variants: prepeg tape, woven tape,
(G,C) dry preform manufacture
4. Pultrusion Continuous fibres, occasional 60--80 Wet resin-impregnated roving or I -beam or rods
CSM and fillers (G, C) cloth formed through a heated die
5. Compression moulding Chopped or continuous fibres 10--60 Pre peg or wet resin/preform pressed Medium-sized complex mouldings
or preforms and fillers to shape in a closed mould between
(G, A, C) matched dies
6. Resin injection, resin Continuous or discontinuous 25-30 Wet resin injected into woven, knitted Panels
(RTM) transfer fibre plus preform polymer or CSM preform within a closed
foam inserts (G, C, A) mould
7. Reinforced reaction Short-milled glass fibre or 5 Fibre-resin dispersion and hardener Large complex parts (e.g. auto
injection moulding fibre preform (G) 20 mixed at rapid mixing head and parts for impact resistance but
(RRIM) injected into mould. with low stiffness)
Chemistry controlled to give rapid
reaction
8. Structural RIM Continuous fibre preform 30--50 Similar to 6 with the more rapid RIM Large structural parts.
(G,A,C) chemistry Rapid production possible
Figure 8.5 Autoclave moulding arrangement for laminates based on epoxy prepreg. (After
Purslow and Childs, 1986).
270 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Difunctional:
0 ~:©-O-CH2-tH-CH2tO@+:&-O-CH2-CH~CH2
= 0), MY 750
f ©-
Diglycidyl ether of bisphenol A (n
CH2~CH-CH2
o 0
CH
CH 3
3 n OH
n
c
CH3
0
Trifunctional:
MY 0510
272 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
(3)
Amine hardeners react with epoxies to form hydroxy amines (3). At first
sight, this might be expected to promote adhesion to the fibres through the
hydroxyl groups, but the polarity promotes moisture sensitivity and
hardening or catalytic systems which provide a less polar molecular structure
are preferred. For industrial lamination, tertiary amine-promoted
anhydride cure is often employed. Nadic methylenetetrahydrophthalic
anhydride (NMA) is commonly employed with the diglycidyl ether
(MY750) and tetrafunctional glycidyl amine (MY720) with benzyl dimethyl
aniline (BDMA) or the similar Epikure K61b. This system provides a matrix
with lower water absorption, which can be attributed to the 'polyester'
copolymer molecular structure (4).
A polyether structure originates from the use of catalytic systems of which
the cationic boron trifluoride amine complex is favoured (5). For advanced
composites, BF3 complexes, often in combination with diamino diphenyl-
sulphone (DDS), act as latent hardeners becoming active at :::: 160°C.
Therefore, the resin viscosity profile with temperature can be adjusted by a
precure (sometimes referred to erroneously as B-staging) to optimise the
COMPOSITE MATERIALS 273
o o 0
x>
II + II +
C -NR3 0 C -NR 3
+ NR3 ~
II II II I
o o 0 0-
X X>
° 0
L"R'
/0,
8
>-
C CH 2-CH-CH 2"'"
II I
° ° °
"C/ /rF
C~
~o
(4)
OH
~ + I
0- CH 2 -CH """'"
I~
(5)
H 2N-C-NHCN
II
NH
H,N-C-NH-CN
+ . II
NH
OH
I
HO-CH
I
JVVVV\I\I\I\/'CH-CH, CH?
.'-....,. I .
N-C-N-CN
/ II
JVVVV\I\I\I\/'CH-CH, NH
I . +
OH
Scheme 8.1 Cure of epoxy resins with DICY. Alternative reaction schemes are discussed by
Amdouni et al. (1990).
COMPOSITE MATERIALS 275
temperature systems have clear advantages, not least the control of the
exotherm during cure, careful control of the curing schedule is required for
the production of good quality void-free laminates. Thus a dwell time at
'minimum' viscosity has to be incorporated, usually = 90°C for these
systems. The other disadvantage of the low temperature curing systems
based on DICY is the increased probability of residual DICY hardener in
the resin. Another observation is the inconsistent appearance of a micro-
structure in these cured systems. Much of the variability can be reduced by
ensuring that the hardener is micronised to a uniform fine particle size prior
to prepreg manufacture. However, microstructural variations can still
occasionally appear but these could also be attributed to phase separation of
the thermoplastic modifier.
DICY is used in catalytic quantities 4--6 phr but the curing chemistry is
complex producing a structure expected of a polyamine as indicated in
Scheme 8.1.
It should be pointed out that the complete reaction of the epoxy groups is
not normally achieved except after long times at high temperatures so that
higher-he at-distortion temperatures and chemical resistance can only be
achieved by appropriate postcuring. The mechanical properties of epoxy
resins are therefore determined by the hardener or curing systems and time-
temperature profile. The time temperature transformation curing diagrams
of Gillham (1986) are particularly valuable in this context. This is discussed
in chapter 1. The effect of post-curing on mechanical properties can be
generalised by appreciating that the heat distortion temperature and
Young's modulus increase with cross-link density and the concentration of
aromatic groups within the network. However, there are exceptions to this
where a higher glass transition can arise alongside a lower modulus, since the
glassy state is frozen in at higher molecular free volume.
8.3.3.2 Rapid curing resins for structural RIM. Much progress has been
made recently in the development of rapid curing epoxy systems for
alternative composites processing.
Stanford (Mortimer et al., 1992) has reported gel times of < 25 s at 40°C
using BF3 diol complexes (6) of Dianol 240 and 320 to cure DGEBA (Dow
Chemical DER332).
HO[CH'-iH-ol!Q\-VO\Lo-yH-cu,lOH
L
BF"
l R L~ CHP l R
(6)
(7)
The shortest gel time of 35 s was observed with the BF3 complex with diol
(8). The maximum cure temperature was 70°C.
(8)
Under a rising tensile stress parallel with the fibres, either the fibres or the
matrix will fail depending upon the relative failure strains of the fibres (ffu)
and matrix (fmu)' Two cases need to be considered (Figure 8.6). Case 1 when
Of _._._._"-
'"'"
Q)
bi
Strain Strain
(a) (b)
Figure 8.6 Individual fibre and resin stress/strain curves defining the failure processes
described in Figures 8.7 and 8.8. (a) Case 1, f mu > flu; (b) Case 2, f mu < flu'
278 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
the failure strain of the matrix is greater than that of the fibres (a typical
fibre-reinforced epoxy resin) and Case 2 when matrix failure strain is lower
(a typical fibre-reinforced ceramic).
Case 1
With uniformly strong fibres at high volume fractions, when the fibres fail all
the load is thrown upon the weaker matrix. The effective cross-sectional
area is reduced and the composite will immediately fail. The strength of the
composite alu is given by
(8.6)
where a~ is the stress in the matrix when the fibre breaks as shown in Figure
8.6.
At low volume fractions (typically less than 0.006--0.03 for GRP and
CFRP) the matrix can carry the load, leading to multiple fracture of the
fibres. The strength is given by
(8.7)
Equations 8.6 and 8.7 are plotted in Figure 8.7 where it is seen that a critical
fibre volume fraction (Verit ) for reinforcement exists. The matrix contributes
insignificantly to the failure strength of the unidirectional laminate in the
fibre direction. The strength is given approximately by equation 8.8.
Single Fracture
~
~
~
u
~
'"'
'"
~
~
.n
~
!
a.
.~
"
~
~
>:
a
mu
o 1.0
Vcri tical
Vf
Figure 8.7 Fibre volume fraction dependence of fracture strength for a unidirectional
composite, Olu, manufactured from a resin of failure strain higher than that of the fibres
(emu> efu)'
COMPOSITE MATERIALS 279
(8.8)
However, as shown in Figure 8.2 the fibres have a statistical failure
strength and strain, and in practice a cumulative weakening process occurs
with fibre breaks at random. The first fibre break may lead to immediate
failure if the flaw has sufficient stress intensity to cause the crack to
propagate. This situation only arises when the fibres are very strongly bound
to a matrix of low fracture toughness. Generally fibre-resin deb on ding
occurs to a limited degree at the fibre ends so that stress transfer through
shear enables the fibre damage to accumulate until the multiple fractured
fibres reach their ineffective lengths. The probability of fracture of adjacent
fibres is increased. This leads to a sequential failure and the production of a
Griffith flaw which can result in composite failure. The generation of a
critical number of adjacent fibre breaks can be inhibited by dispersing a
higher failure strain fibre into the bundle to increase the ultimate strength in
what has been termed the hybrid effect.
The tensile fracture surface has a brush-like appearance with varying
degrees of fibre fracture and pull-out making meaningful strength
predictions difficult. For further discussion of this problem the reader is
referred to the texts by Hull (1981) and Kelly (1973).
ease 2
Under these circumstances the matrix fractures first, throwing all additional
load onto the fibres (Figure 8.6). At low Vf as illustrated in Figure 8.8, the
v' 1.0
f
Vf
Figure 8.8 Fibre volume fraction dependence of fracture strength of a unidirectional composite
manufactured from a matrix of failure strain lower than that of the fibres (cmu < Cfu)'
280 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
fibres are unable to support the additional load and the composite fractures.
The strength is given by
(8.9)
where of is the stress on the fibres when the matrix fractures (see Figure 8.6).
At higher V f when V f > V f (= Omu/Ofu - Of + omu) the fibres are able to
support the extra load when the matrix fails. Additional loads cause multiple
cracking of the matrix. Further details are given by Aveston and Kelly
(1973). The strength of the laminate is clearly determined by the average
strength of the fibres O"fu and
(8.10)
In practice the distribution of values of 0fu means that some fibres may fail
before the onset of matrix failure and continual damage accumulation of
matrix cracks and fibre fracture occurs.
The rule of mixtures equation 8.13 is a poor estimate of the actual transverse
modulus and a number of other attempts have been made. The most
rigorous approach is that of Hashin and Rosen (1964) in which upper and
lower bounds may be calculated. The Halpin and Tsai (1969) equations are
generally applicable and allow for variations in packing geometry and
regularity with the factor;. These rules are compared in Figure 8.9 where it
is clearly seen that transverse fibres are inefficient reinforcing elements.
(8.14)
where
COMPOSITE MATERIALS 281
q- 20
E
z
~
ur
vi 15
:J
:;
"0
0
E
~
.0; 10
c
$
'i!?"
'c>"
<f)
5
~
f-
0 0.8
VI
Figure 8.9 The effect of fibre volume fraction, V" on the transverse tensile modulus of a
unidirectional composite. To accommodate the range of V, both epoxy resin (V, = 0.25-D.6)
and an unsaturated polyester (V, = 0.08-0.3) resins have been used. (e), Glass-polyester;
(0), glass-epoxy; (-), Law of mixtures (equation 8.12); (---), Halpin-Tsai equation (equation
8.14) which gives the best fit to the data. (Redrawn from Bailey and Parvizi, 1981.)
c:
.~
1il \,
~ \
::> '\
~ 2 '\
...
'\
CD
li? "-
"-
CD
>
!Il
"- ,
C "
'"
~ 1 "
......... ' .................
~-
Figure 8.10 The effect of fibre volume fraction, VI, on the transverse failure strain of a
unidirectional E-glass-epoxy composite. (-), predictions of the Kies theory.
Carbon 0.62 127 8.3 1.7 0.039 1.16 0.48 0.29 0.019
Glass 0.55 42 14 0.92 0.056 2.2 0.5 0.27 0.09
Aramid 0.62 83 5.6 1.31 0.039 0.35
, V is the Poisson ratio in the longitudinal (I) and transverse (t) directions.
COMPOSITE MATERIALS 283
associated with transverse fibre failure and fibre-matrix debonding. Design
should employ a careful selection of fibres to provide the required
performance.
Fibre surface treatments and coatings are currently receiving more
attention as a means of improving and optimising interlaminar shear and
transverse properties.
600
\
\
\
500 \
E 400
z 0
~
1u
:;'"
U 200
u:'" ~
\
\
100 /"0-..
Tsai·Hill '"
/0, = s~n~ e
criterion ~
o
Off-axis angle, e (degrees)
Figure 8.11 Effect of angle of test (8) on the strength of a unidirectional carbon fibre (high
strength) epoxy composite (Vf =0.5) showing the predictions of the Tsai-Hill model.
(Redrawn from the original given by Hull, 1981.)
284 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
For angles of (J to fibres between 0° and 90° the applied stress (at angle (J), G()
will have the following components and the failure will be determined by the
weakest:
G() = Glulcos2 (J
G() = Gtu /sin2 (J (8.19)
G() = iu/sin (J • cos (J
This is illustrated in Figure 8.11, where these equations are compared for
epoxy-based CFRP. At angles of 0-5° the failure is dominated by
longitudinal fracture, from 5-18° by intralaminar shear failure and from
45-90° by transverse fracture. In the region from 18-45° a mixed mode
failure occurs. The actual angles at which these transitions take place will
clearly be dependent upon the values of Gtu , Glu, iu. The Tsai and Hill
criterion which is based on maximum work theory however gives a better
prediction of the off-axis strengths in this region (Halpin and Tsai, 1969; Tsai
and Hahn, 1980; Hull, 1981; Tsai, 1985).
t I direction
t direction
If
'" '"
Splitting in
longitudinal (0°)
lamina
Transverse
crack in
transverse (90°)
lamina
Figure 8.12 A 0°/90%° crossply laminate showing transverse cracks which progressively form
under a rising tensile stress. Longitudinal splits in the 0° laminate also occur at high composite
strain.
125
~ 75
~
VI
VI
~
iJj 50
Stress Whitening
25
Strain (%)
Figure 8.13 Low strain part of a stress/strain curve for a 0°/90%° laminate showing the onset of
cumulative damage beginning with fibre-matrix debonding (stress whitening) and transverse
cracking (see Figures 8.12 and 8.14) (after Manders et aI., 1983).
286 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
(nOO (k)03L (1)036 (m) 0 51. (n)On (0)090 (p)1.08 (q) ISO (,)190
.20mm.
Figure 8.14 Progressive multiple cracking in the transverse ply of a glass·epoxy crossply
laminate. Note the presence of longitudinal splits at higher strain. The numbers represent the
strain at which the photograph was taken (after Manders et al., 1983).
the 90° ply at a transfer distance of = 2 mm from the crack. This results in a
black band either side of the transverse cracks as shown in Figure 8.14.
Multiple cracking of the inner ply continues as the load on the specimen is
increased until saturation has occurred when the shear-lag stress transfer can
no longer exceed the transverse strength of the remaining elements.
Under continued loading Poisson strains are induced which cause the
longitudinal ply splitting shown schematically in Figure 8.12. The onset of
ply splitting is also seen in the photographic record of the cracking sequence
in Figure 8.14.
Parvizi et al. (1978) used a modified shear lag analysis to calculate the
additional stress transferred back into the transverse ply at a distance from
the crack, y. These authors assumed that the cracks would form midway
between the previous ones. If iloo is the additional stress required to load the
transverse ply to its cracking strain then
iloo = ctluEtd/b[1 + exp (- q,1I2 t) - 2 exp (- q,1I2t)/2)rl (8.20)
where q, = EcGt(b + d)/EIEtbd2 and d is the semi-90° ply thickness, b is
the 0° ply thickness, tis the crack spacing and Oa is related to iloo according to
(8.21)
The nature of the accumulation of transverse cracks dictates that the crack
spacings predicted by these equations are a stepped curve. However, it has
been suggested that since the transfer distance y is small by comparison with
the crack spacing that the first cracks form statistically at the weakest points
in the 90° ply (see for example Manders et al. (1983) and Peters (1984».
Kimber and Keer (1982) showed that the average crack spacing is related to
the lower bound of the shear lag analysis.
The statistical aspects of crossply cracking can arise from variations in
fibre dispersion within the laminae, and could be worse when a memory of
COMPOSITE MATERIALS 287
the individual tows remains. Peters and Chou (1987) have analysed the
process in terms of an increasing element strength of a weak link theory and
have argued that the statistics can be extrapolated to obtain the true
transverse strength of the 'defect-free' ply.
tl -
where Tl is the stress-free temperature. (at - al) (Tl - T2) can be obtained
directly from the curvature of an unbalanced 0°/90° coupon in a manner
analagous to a bimetal strip.
(8.27)
Ec = 2n
5 0JCI2 Eod() (8.30)
D = Dx ( 1+
h
- yDy- + -h yDz)2
- (8.38)
1 Dx n Dx
where h, I, n are thickness, length and width respectively of the coupon.
M is the moisture concentration at time t and Moo is the equilibrium (i.e.
maximum) moisture content at infinite time.
292 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
D = Dx ( 1 + I +
h h)2
~ (8.39)
0
'\
\
10 - \
\
\
20 L \
\
6' I
30 r
\
e.... \151
\
0
- ~
~
::J
....Cll
ctS
a.
40 'J \
i;.
III \
\
- ~
E 50 \
Cll \
c I
(;
\ 0
0 60 r \
+= \
151
'iii
-
c \
~ 70 (;
\ 0
(;\ 151
/Jl
/Jl \
ctS 80 \
0> III \
.!: \
Cll 90 \ III
/Jl
ctS \
....
Cll \
100 \
(.)
Cll 'J ~x
0 \
\\
151
110 (;
\
120 \
(; (; \c:,
\
130 \
\
(; (; \
140
0 2 3 4 5 6 7 8
Moisture content (%)
Figure 8.16 The effect of moisture on Tg of a series of differing epoxy resins showing generality
of the effect of moisture plasticisation. Each symbol represents a different resin system.
(Redrawn from Wright, 1981).
294 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
250
200
x
~ 150
OJ
"""" 100
50
o ~~--~--~--~--~--~--~--~
o 2 3 4 5 6 7
M (wt%)
Figure 8.17 Change in Tg of Narmco 5208 epoxy resin as a function of equilibrium moisture
content. Each symbol represents the results of different authors where variations in cure give
differing 'dry' values. The wet Tg however tends to the same value in all work.
Figure 8.17 indicates the consistency of this observation with the variation
of Tg with differing moisture concentrations for nominally the same resin
system, as given by different workers (Mulheron, 1984).
The major consequence of this result is illustrated in Figure 8.18 where it
is seen that not only the glass transition temperature but also the modulus is
reduced. Since for most composite matrices, the average moisture
absorption at equilibrium leads to a reduction in Tg by about 50°C, 150-
\
"- '-
,
W
::J
::;
'-- ---
wet/- -, ......
"C
o '-
E "-
OJ
>
~
"-
Qi
\
II: \
\
\
'l
o 150 300"C
Temperature
Figure 8.18 The effect of moisture on the temperature dependence of the relative matrix
modulus of Narmco 5208 epoxy resin.
COMPOSITE MATERIALS 295
200°C resins can only be used where service temperatures are limited to 100-
150°C. This has led to the development of more highly functional resins
which have a higher cross-linked density, Tg and service temperature in
humid environments.
While simple resin systems appear to obey Fickian diffusion kinetics
under absorption and desorption conditions, the compounded resins
employed for advanced composites may show irreversible effects. For
example, the presence of residual DICY can increase the value of Moo
significantly. Moreover, on immersion in water, blister formation may occur
as a result of osmotic effects accompanied by leaching (Jones et al., 1987).
As shown in Figure 8.19 the presence of a second polymeric component may
lead to moisture-induced changes in the matrix-dominated expansion
coefficients and this can have severe effects on micro crack development in
laminates during thermal excursions (section 8.5.4) (Jacobs and Jones,
1991).
150
C
(!)
'(3
~a
u
c:: 100
.2
<n
c::
:.::
'0.x"
OJ
'f
~
Cil ~
E
OJ
.c:: 50
f-
o
o 50 100 150 200
Temperature (oC)
Figure 8.19 The temperature dependence of the transverse thermal expansion coefficients of
the bismaleimide-modified epoxy resin based carbon fibre composite (Narmco S24SC). Dry
(lower curve) and wet (upper curve). The cross hatching indicates the relative thermal strain
induced on cooling from Terit (dry) and Terit (wet) (lower and upper curves respectively). Tcrit is
the temperature to which a crossply laminate needs to be heated and then cooled to induce a
f
value of f:~ which exceeds tu and cause thermal cracking. Since Tg (dry) is below Terit (dry)
thermally-induced cracking does not occur but the reverse situation operates for the wet
laminate and thermally-induced cracking does occur (Jacobs and Jones, 1991).
swelling. Much debate has centred around the extent of this effect and the
accommodation of water in the free or unoccupied volume of the dry glass.
However, matrix swelling negates the restrained contractions which had
already occurred on cooling through the strain-free temperature from the
post-cure temperature after fabrication. As a result, the thermal strain is
reduced by moisture ingress as shown in Figure 8.20.
For other systems the effect is bigger and, for example with Fibredux
913G, the residual stresses are reduced to a very low level as a result of
moisture absorption.
1.1
1.0
•
:>..
•
L
== 0.9
"U
(,)
"-:;:::
:5
(,)
0.8
O. 7 '----'----'----L_"'--....L..---'----'_"'---'---L..---'~
0.8
?f- 0.6
8
::'E
0.4
0.2
Figure 8.21 Comparison of moisture absorption by a 0° carbon fibre epoxy laminate (Narmco
5245C) under isothermal conditions (continuous line) and subject to intermittent thermal
spikes to 150°C (saw-tooth). The environment was 96% RH at 50°C. The saw-tooth illustrates
partial drying during thermal cycling and the subsequent enhanced water gain.
298 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
mechanism of this phenomenon is not fully clear but may involve the re-
equilibration of the network molecular structure, since the critical thermal
spike temperature above which this phenomenon is observed appears to be
related to the glass transition of the wet resin.
8.8 Conclusions
Epoxy resins form the basis of the advanced composites industry as the
preferred matrix for carbon fibres. These systems are designed to give good
COMPOSITE MATERIALS 299
process viscosity profiles and involve latent hardeners for use in prepreg
technology. The micro mechanics of laminate systems is strongly dependent
on the properties of the matrix. However, the absorption of moisture and
the maximum achievable glass transition temperature can limit the
maximum service temperature of the composite to < 200°e.
8.9.1 Laminates
D Interfibre spacing
E(E) Young's modulus (average value)
G Shear modulus
L Length of composite coupon
T2 Measurement temperature in equation 8.25
Tj Strain-free temperature
V Volume fraction
b Outer ply thickness
c Specimen width
d 90-ply semi thickness
df Fibre diameter
rf Radius of fibre
t Crack spacing
y Shear lag transfer length
a Linear expansion coefficient
Yt Fracture surface energy of the 90° ply parallel to fibres
<5 Displacement of 0°/90° beam from flat
c Strain
a Stress
e Radius of curvature of a 0°/90° beam
T Shear stress
£ Tsai-Halpin geometry and dispersion factor
'YjO,'Yjl Orientation (0) or length (1) factors for random discontinuous
composite
v Poisson ratio
Diffusion theory
D Diffusion constant or diffusivity
M Moisture content at time t (in %)
Moo Moisture content at t = 00 (in % )
C Moisture concentration (g . mm -3)
Time
300 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Subscripts
a Applied stress (oa) or strain (e a)
f Fibre (e.g. Of stress in fibre)
m Matrix (e.g. em strain in matrix)
c Composite (or Lc critical length for a short fibre in a composite)
1 Longitudinal or 0° direction of a unidirectional ply
t Transverse or 90° direction of a unidirectional ply
u Ultimate value at failure
fu Ultimate value for failure of a fibre
mu Ultimate value for failure of the matrix
lu Ultimate value for failure of a longitudinal composite
tu Ultimate value for failure of a transverse composite
tIu Ultimate value for transverse cracking of transverse ply
tu Ultimate value for failure of a transverse ply
tI Property of transverse ply in longitudinal direction
It Property of longitudinal ply in transverse direction
11 Property oflongitudinal ply in longitudinal direction
Crit Critical value
Superscripts
th Thermal
Property in the component when other fails
P Poisson
Min Minimum value of property
References
Alexander, M.R. and Jones, F.R. (1991) In Composites, Design Manufacture and Application,
Vol. 1, eds Tsai, S.W. and Springer, G.S. Sampe, Corina, USA. Chapter 11, paper G.
Amdouni, N., Sautereau, H., Gerard, J-F. and Pascault, J.P. (1990) Polymer 31,1245-1253.
Aveston, J. and Kelly, A. (1973) I. Mat. Sci. 8,352.
Bacon, R. (1980) Phil. Trans. R. Soc., London A294, 437.
Bader, M.G. (1985) In Handbook of Composites, Vol. 4, ed. Kelly, A. and Mileiko, S.T.
Elsevier, North Holland, p. 177.
Bailey, J.E. and Parvizi, A.11981) I. Mat. Sci. 16,649.
Cox, H.L. (1952) Br. I. Appl. Phys. 3,72.
COMPOSITE MATERIALS 301
Denison, P., Jones, F.R. and Watts, J.F. (1988a) Surface and Interface Analysis 12, 455.
Denison, P., Jones, F.R. and Watts, J.F. (1988b) In Interfaces in Polymer, Ceramic and Metal
Matrix Composites, ed. Ishida, H. Elsevier, New York, pp. 77-86.
Denison, P., Humphreys, P., Watts, J.F. and Jones, F.R. (1989) In Interfacial Phenomena in
Composite Materials, ed. Jones, F.R. Butterworths, Guildford, pp. 105-110.
Dobb, M.G. (1985) In Handbook of Composites, Vol. 1, eds Watt, W. and Perov, B.V.
Elsevier, Amsterdam, pp. 673-7.
Dunsford, D.V., Harvey, J., Hutchings, J. and Judge, C.H. (1981) RAE Technical Report
TRSJ096, HMSO, London.
Drzal, L.Z., Rich, M.J. and Lloyd, P.F. (1982) J. Adhesion 16, 1.
Folkes, M.J. (1982) Short Fibre Reinforced Thermoplastics, Research Studies Press,
Letchworth, UK.
Garrett, K.W. and Bailey, J.E. (1977)1. Mat. Sci. 12,157.
Gillham, J .K. (1986) Encyclopaedia Polym. Sci. Eng. 2nd edn, Wiley, New York, pp. 519-524.
Halpin, J.G. and Tsai, S.W. (1969)AFML-TR67, 423.
Hashin, Z. and Rosen, B.W. (1964) J. App. Mechanics Trans. ASME 31,223.
Hull, D. (1981) An Introduction to Composite Materials, Cambridge University Press, UK.
Jacobs, P.M. and Jones, F.R. (1991) In Composites, Design, Manufacture and Application,
Vol. 2, Ch. 16. eds Tsai, S.W. and Springer, G.S. SAMPE, Corina, USA, paper G.
Jones, F.R. (1989) In Interfacial Phenomena in Composite Materials ed. Jones, F.R.
Butterworths, Guildford, pp. 25-32.
Jones, F.R., Shah, M.A., Bader, M.G. and Boniface, L. (1987) Sixth Int. Conf. on Compo
Materials, 2nd European Conf. Compo Materials ICCM VIlECCM2 Vol. 3, ed. Matthews,
F.1. et al. Elsevier, London, pp. 443-436.
Kelly, A. (1973) Strong Solids. Clarendon Press, Oxford.
Kelly, A. and Tyson, W.R. (1965) J. Mech. Phys. Solids 13, 329.
Kies, J.A. (1962) US Naval Research Report No. 5752.
Kimber, A.C. and Keer, J.G. (1982) J. Mat. Sci. Letters 1, 353.
Loewenstein, K.L. (1983) The Manufacturing Technology of Continuous Glass Fibres.
Elsevier, Amsterdam.
Manders, P.W., Chou, T-W., Jones, F.R. and Rock, J.W. (1983) J. Mat. Sci. 18,2876.
Mortimer, S., Ryan, A.J. and Stanford, J.L. (1992) Proc. 5th Int. Conf. FRC '92, paper 2
Plastics and Rubber Inst, London.
Mulheron, M.J. (1984) PhD Thesis, University of Surrey, UK.
Oberlin, A. (1985) In Developments in the Science and Technology of Composite Materials eds
Gunsell, A.R., Lamicq, P. and Massiah, A. Elsevier Applied Science, London. pp. 15-20.
Parvizi, A. and Bailey, J.E. (1978) J. Mat. Sci. 13,2131.
Parvizi, A., Garrett, K.W. and Bailey, J.E. (1978) J. Mat. Sci. 13,195.
Peters, P.W.M. (1984) J. Compo Mat. 18,845.
Peters, P.W.M. and Chou, T-W. (1987) Composites 18, 40.
Plueddemann, E.P. (1990) Silane Coupling Agents 2nd edn. Plenum, New York.
Purslow, D. and Childs, R. (1986) Composites 17,127.
Rand, B. (1985) In Handbook of Composites, Vol. 1 eds Watt, W. and Perov, B.V. Elsevier,
Amsterdam, pp. 495-575.
Rosen, M.R. (1978) J. Coatings Tech. 50,70.
Schapery, R.A. (1968) J. Compo Mat. 2,380.
Shen, C.H. and Springer, G.S. (1977) J. Compo Materials 11,107.
Thomason, J. and Morsink, H. (1988) In Interfaces in Polymer, Ceramic and Metal Matrix
Composites, ed. Ishida, H. Elsevier, New York.
Tsai, S.W. (1985) Composites Design, USAF Mat. Lab. - Think Composites, New York.
Tsai, S.W. and Hahn, H.T. (1980) Introduction to Composite Materials, Technomic, Westport.
Verpoest, I. (1987) In Proc. Europ. Symp. on Damage Development and Failure Processes in
Composite Materials, eds Verpoest, I. and Wevers, M., Leuven, Belgium, pp. 70-77.
Waltersson, K. (1985) Compo Sci. Tech. 22,223.
Wang, D., Jones, F.R. and Denison, P. (1992) J. Adhesion Sci. Tech. 6,79.
Watt, W. (1985) In Handbook of Composites, Vol. 1. eds Watt, W. and Perov, B.V. Elsevier,
Amsterdam, pp. 327-387.
Wright, W.W. (1981) Composites 12, 201.
302 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
Further reading
Chawla, K.K. (1987) Composite Materials - Science and Engineering, Springer-Verlag, New
York.
Hull, D. (1981) An Introduction to Composite Materials, Cambridge University Press,
Cambridge.
Jones, R.M. (1973) Mechanics of Composite Materials, McGraw-Hili, Washington DC.
Kelly, A. (1973) Strong Solids, Clarendon Press, Oxford.
Kelly, A. and Rabotnov, Yu.N. (1985) Handbook of Composites, Vols 1-4, Elsevier Science,
Amsterdam.
Tsai, S.W. (1985) Composites Design 1985, Think Composites, Dayton, Ohio.
Tsai, S. W. and Hahn, H.T. (1980) Introduction to Composite Materials, Technomic, Westport,
Connecticut.
Donnet, J.B. and Bansal, R.C. (1984) Handbook of Fiber Science and Technology, Vol. 3,
Marcel Dekker, New York.
Plueddeman, E.P. (1982) Silane Coupling Agents, Plenum, New York.
Weatherhead, R.G. (1980) FRP Technology, Fibre Reinforced Systems, Applied Science,
London.
9 Coatings and other applications of epoxy resins
X.M. CHEN and B. ELLIS
9.1 Introduction
The current use pattern for epoxy resins is similar to that suggested by Lee
and Neville in 1967 who estimated consumption as coatings 55%, composite
matrices 20%, castings 10%, adhesives 5% and miscellaneous 10%. These
figures may be compared with the 1990 and 1991 estimates given in chapter
1, Table 1.1, even though the categories are not identical. Lee and Neville
(1967a) pointed out that there are difficulties in obtaining exact
consumption estimates. However, the main difference is that coatings
have a slightly lower share of the epoxy market. Epoxy resins are usually
more expensive than their rivals, such as phenolic resins for coatings or
laminates. Therefore, epoxy resins find application because of their superior
properties, which include both processing and those of the cured resin. The
processing is convenient since it is possible to formulate compositions with
the required rheological properties, such as low viscosity, and there is also a
wide choice of hardeners so that it is possible to cure at ambient as well as
elevated temperatures. Because epoxy resins can be cross-linked without
the formation of low molecular weight products, volatiles are not evolved
during cure, and the resins have only a relatively low shrinkage during
curing. Their mechanical and electrical properties are superior to other
resins and they have good heat and chemical resistance.
There is a large range of epoxy resins available commercially and some of
the more important types are listed in Table 9.1. Exact specifications can be
obtained from resin suppliers and many mixtures are offered by
compounders. Selection of a resin-hardener system can be made by a critical
assessment of processing and end-use requirements, e.g. the electrical
properties, or resistance to UV radiation. It is possible to enhance a desired
parameter simply by altering the cure schedule without a change in the initial
composition of the resin-hardener system. Thus, it is possible to increase
the glass transition temperature (Tg) of the cured resin by post-curing.
Such effects can be appreciated by study of the time - temperature-
transformation diagram, TIT, given in Figure 1.2 and the detailed
discussion of cure given in chapter 3. Extensive data compilations of the
properties of epoxy resins are available and Table 9.2 is a selection of the
'tensile' mechanical properties and also an estimate of the heat distortion
temperature. The Young's modulus of epoxy resins is somewhat lower than
Table 9.1 Summary of commercially available epoxy resins a
- - O-CH,-CH-CH,
"(~rn-cHoJ9 ~ /\
ii. Others include:
Tetraphenol ethane derivatives Powder coating
Polyglycidyl ether of tetraphenol ethane: Shell 1031
"O~ )§l0"
5§5rn-rn~
HO OH
4. Amines
Reaction products of epichlorohydrin with
amines such as:
~
r$) NH2
C$JNH2
~ NH2
Table 9.2 Mechanical properties and range of heat distortion temperatures of cured epoxy
resins
Mechanical properties
Badge type:
Flexible 0.1-1 10-70 5-10
Rigid 1.5-3.5 25-90 2.5-4 100--200
Epoxy-novolac 1.5-3.5 70 2 150-250
Cycloaliphatic epoxy 3.5 90 2-10' Upto200
Phenol-formaldehyde 2-4 20-60 1.5 150-170
Unsaturated polyesters 1.5-3.5 40-90 <2 Up to 200
that of phenol formaldehyde (P/F) and unsaturated polyester resins but the
tensile strengths are comparable and may be higher. This is because the
extensibility, that is the fracture strain, of the epoxy resin is much higher
than that of the 'brittle' resins. It is possible to select from a range of
flexibilizers (see chapter 4) to 'tailor' the mechanical properties to suit the
application requirements. For instance the 'tensile strength' can be
significantly affected by cure treatment and the fracture behaviour of epoxy
resins should always be considered when an application demands optimum
strength. There is a detailed discussion of fracture behaviour in chapter 5
which is also relevant to the adhesive (chapter 7) and composite (chapter 8)
applications of epoxy resins.
With standard BPA-EpiCI resins the heat distortion temperatures are
similar to those of PIF and polyester resins, but with suitable hardeners and
a post-cure at elevated temperature it is possible to attain higher glass
transition temperatures and hence higher use temperatures. For instance a
BADGE-type resin cured with trimellitic anhydride and cured for 1 hour at
120°C followed by 2 hours at lS0°C may have a heat distortion temperature
of up to 2S0°C.
Cyclo-aliphatic resins have very good outdoor ageing properties. This is
because they do not contain aromatic rings which absorb the UV radiation
which leads to degradation processes. Also, the chlorine, and other halogen,
content of the cyclo-aliphatic resins is essentially zero and hence these resins
have very good electrical properties.
Data compilations such as that given in Table 9.2 are indicative of the
properties that may be achieved with epoxy resins. However, by comparison
with the PIF and unsaturated polyester resins it would not be possible to
assess the superior technical properties of the epoxy resins. Their processing
properties offer considerable advantages. The range of hardeners available
is a major asset since it is possible to formulate a curing system to suit almost
Table 9.3 Some typical curing agents and applications for BADGE-type resins
Hardener' Physical state Concentration Typical cure schedule Approximate heat Suggested applications 3
at ambient (phr)2 distortion temperature
COC)
1. Aliphatic amine
Triethylene tetramine Liquid 13 a. 25°C, 7 days a. - General purpose
b. 25°C, 1 day plus post- b. Up to 120 Co, A, Ca, M, F, Sand
cure at elevated T FW
leffamine D-400 Liquid 55 1l0-120°C, 30 min -30 Flexible resin Co, A,
Ca,FandS
2. Aromatic amine
DDM Solid 27 1 hat 150°C Upto 150 Glass cloth laminate
+3hat 180°C
DDS (orDAS) Solid 25 5 h at 125°C Upto175 A, Ca, M, E, Sand FW
+ 1 hat200°C
3. Polyamides Liquid Depends on type a. Ambient temperature Up to 100 a. Co,A,Ca,F,M
7 days andS
b. 1 h at 100°C Up to 100 b. CaandA
4. Anhydrides
Methyltetrahydro- Liquid 80 2hat20°C 130 Liquid hardener
phthalic anhydride +2hat 150°C processing
(MTHPA) A,Ca,M,SandFW
Trimellitic anhydride Solid 35 1 h at 120°C 250 High heat distortion
(TMA) +2hat 150°C temperature
A,Ca,M,SandFW
5. Miscellaneous
Dicyandiamide (DICY) Solid 4 1 h at 180°C 150 A, Ca, M, E, Sand FW
BF3-Monoethylamine Solid 3 1 hat 120°C 170 A, Ca, M, E, Sand FW
(BF3-MEA) +2hat 170°C
, See chapter 2 for fuller details. Many hardeners are mixtures and/or derivatives and are not included in this list. Many other hardeners may be used. Shell
list 30 curing agents for Epon 828. This table is only illustrative.
2 phr = parts per hundred of resin by weight.
3 Code: Co = coatings; A = adhesives; Ca = castings; E = electrical laminates; F = flooring; M = mouldings, S = structural laminates; FW =
filament winding.
308 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
9.2.1 Introduction
Surface coatings may be primarily applied for protective or decorative
purposes but usually they have dual or more functions, such as exterior
automotive coatings which are both protective and decorative. They may
have specialized functions, such as anti-fouling paints for ships' bottoms or
the protection of food and beverages in metal cans. There are also 'high-
tech' applications in the electronic and related industries such as coatings for
fibre optics, printed circuit panels and many other applications (Burkhart,
1988; Emerson etal., 1990; Bonneau and Burkhart, 1991). Coatings contain
a polymeric film-forming material, pigment, special additives and may
be dissolved in a suitable solvent mixture or dispersed in water. Powder
coatings do not contain a solvent. There is a range of film-forming materials
which may be classified as either non-convertible or convertible (Martin,
1972). The former includes shellac, cellulose nitrate or acetate, cyclized and
chlorinated rubber and vinyl and acrylic resins, which form films when a
solvent evaporates. Convertible materials include the drying oils, varnishes,
thermosetting resins such as alkyds, phenolics, epoxies and polyurethanes in
Table 9.4 Film forming materials for coatings
Application Epoxy Acrylic Alkyd Polyester PlFa AIFb Coal tar Alkyd Polyester AIFb P/F a Polyurethane
resins modified
aP/F is a phenol-formaldehyde resin; b AIF includes urea and melamine formaldehyde resins; C Non-convertible coatings include drying oils and
cellulose ethers and esters.
310 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
which the film former becomes cross-linked either by reaction with oxygen
or suitable curing agents or hardeners. A general text on surface coatings
is that of Paul (1986).
There are many different coating methods which can be adopted for
specific applications, depending on the coating thickness, coating speed and
type of finish required. A review of 25 methods with outline descriptions is
given by Scharenberg (1985), spray coating is discussed by Coeling (1985)
and powder coating by Richart (1982, 1985). A listing of the polymer
systems used for specific industrial coatings is given in Table 9.4. Epoxy
resin coatings are discussed in detail by Somerville and Smith (1973), Potter
(1970), Lee and Neville (1967) and Tess (1988a). Resin manufacturers offer
advice on the use of epoxy resins for the fomulation of coating compositions,
Ciba-Geigy (1988) and Shell (1991) offer a range of epoxy resins for surface
coatings, including Epon 828, with an epoxy equivalent-weight of 185-192.
This resin and also the Ciba-Geigy equivalent GY-250 have been
formulated as a heat-cured coating for strengthening glass (Chen et al.,
1992). Epoxy resins can be formulated to have much shorter cure times (1 to
30 minutes
at 240-135°C) than equivalent polyurethanes or polyesters (10 minutes
at 200°C) and have excellent adhesion to most substrates, provided the
surface has had an appropriate pre-treatment.
where R is the gas constant and T the absolute temperature. For the use of
this equation the vapour phase must obey the ideal gas law. For epoxy resins
the value may be determined from swelling measurements or by calculation
useing additive group cohesive energy constants (see Barton, 1983).
The simple matching of solubility parameters is only suitable for an initial
selection of a solvent because other factors such as polar attraction force and
hydrogen bonding complicate solvent-solute interactions. Tess (1988b)
gives useful diagrams for solvent selection and Barton's (1983) solubility
maps for epoxy resins can also be consulted. Solvents for epoxy resins
include xylene, methyl ethyl ketone (MEK) , methyl isobutyl ketone
(MIBK) and proprietary solvents, such as oxitol. Mixed solvents are often
used so that not only solubility but the rate of evaporation and also the
rheological properties of the solution are optimized for the specific
application. Solvent mixtures include toluene with acetone, MEK, MIBK
and also xylene, MIBK and 2-ethoxyethanol (Lee and Neville, 1967).
After the selection of suitable solvents and subsequent formulation of a
coating solution, other factors need to be considered carefully to ensure
successful coating application. These include the condition of the surface
prior to coating, concentration ofthe solution, curing technique, pot-life of
the solution, etc. It should be noted that the coating solution is a dynamic
system within which chemical reaction (cure) progresses with time. There-
fore care should be taken in its storage. Also, the rate of evaporation of the
solvent in the system depends not only on the vapour pressure of the solvent,
but also the curing method used. Evaporation will be faster at higher
temperatures, but then the rate of cure is higher and the film may harden
before all solvent molecules diffuse out of the coating. The trapped solvent
molecules will affect the properties of the film which may be impaired and
may lead to poor adhesion. Thus, it is always important to consider all the
factors in order to establish a set of optimized application conditions.
The presence of pigment will have a considerable effect on the flow
properties of a coating solution. Nielsen (1977) discusses the effect of fillers
on viscosity and suggests a generalized relationship for the viscosity of a
dispersion, related to that of the fluid 'I,
/ _ 1 +AB¢2
'I '10 -1 - B1jJ¢2
'O""-@-°frn,-rn,+ ]0
The hydrocarbon 'end' is soluble in the resin and the polar part of the
molecule is hydrophilic where the length of the ethylene oxide chain can be
chosen to suit the application. There are very many such emulsifying agents
available commercially, and mixtures may be used to ensure that an
emulsion is stable (Ash and Ash, 1980-1983). It is essential that the particle
or droplet size of the emulsion is less that one micro metre to ensure that the
emulsion does not settle out. Also, the emulsifying agent should be chosen
to ensure that the droplets do not aggregate or even worse, coalesce. The
stability of an emulsion is affected by shearing and hence the surface active
agent should form a protective layer around each droplet. Changes in
temperature also affect the hydrophile-lyophile balance (HLB) of the
surfactant (Shinoda and Friberg, 1986).
COATINGS AND OTHER APPLICATIONS 315
To recapitulate from this discussion it can be easily appreciated that two
basic criteria for the selection of emulsifiers must be satisfied.
• Reduce the interfacial tension between water and resin
• Form protective films around the resin droplets.
Although analytical methods can assist in the selection of emulsifier, a
direct screening of the available emulsifiers is essential in practice to ensure
that all requirements are satisifed.
There are three sets of conditions that an epoxy coating emulsion must
comply with:
1. Preparation, stability and shelf-life
2. Application of the coating
3. Film formation.
The essential factors for the first set have already been outlined. For the
second set the rheological properties of the emulsion must be such that the
emulsion can be applied either by dipping, spraying or electrodeposition.
With dipping the film thickness is determined by 'drainage' of the emulsion
as the coated product is removed from the emulsion. The 'quality' of the film
formed depends on even drainage, such quality involves control of the
coating thickness as well as the avoidance of surface blemishes. It is
particularly important that there is even coverage, which is difficult to
achieve by spray application at places adjacent to corners, which should be
avoided if possible by alternative product design. The drip that forms in the
region of the 'neck' as the product leaves the coating bath can be trouble-
some. With spraying it is essential that fine droplets are formed since when
large they splash where they hit which can lead to a 'more or less grainy
surface' which eventually leads to a film with imperfect specular reflection.
The rheology of the emulsion will affect size reduction in the spray
equipment and also the flow of the emulsion on the coated surface which can
lead to the same type of defects that occur with dip coating, except for a
'neck' blemish. Design of equipment for the support of the product during
spraying is required so that the uncoated region adjacent to suspension
hooks or grips is unimportant or can be subsequently treated appropriately.
Film formation involves the evaporation of water which requires more
heat than for the removal of an organic solvent. The rate of diffusion of
solvent or water through a polymer film is also important since it is essential
that all of the low molecular weight species are removed prior to curing.
With evaporation of water, a stage will be reached when the emulsion
inverts and the epoxy resin becomes the continuous phase. Such a phase
inversion will depend on the concentration of water present in the film and
the temperature as well as other factors. The phase inversion temperature
(PIT) is discussed by Shinoda and Friberg (1986). Evenness and film quality
may be improved by the use of an organic solvent to reduce the viscosity of
the resin prior to cure. The flow properties of the uncured film will depend
316 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
on the pigment, its type, concentration and particle size distribution as well
as the presence of other additives. The flexibility of the coating can be
improved by the addition of alkyd, polyester or acrylic resins, polymeric
amines and polyamides may be used as curing agents. A can coating has
been formulated from a higher molecular weight epoxy resin (M. wt -15-
20 000) which was initially dissolved in solvent and then dispersed in water
and cured with either a melamine or urea formaldehyde resin. The baked
films were 'flexible' and resistant to the pasteurization conditions required
for beer cans. Water-based epoxy resin coatings can be formulated for
electrodeposition, both anodic and cathodic (Brewer, 1985). These coatings
were developed for the applicance and automobile industries with cathodic
systems being replaced by the anodic coatings because higher corrosion
resistance can be produced. With electrodeposition (ED) the loss of paint
due to overspray is eliminated, so that the process offers advantages
compared with spraying since with ED it is possible to obtain more uniform
coatings even in recessed regions. Very many formulations have been
proposed for water-based coatings for dipping, spraying or electro-
deposition.
It is possible to build hydrophilicity into an epoxy resin and an example is
the reaction product of an epoxy resin with dibasic acids or anhydrides, and
many different combinations have been evaluated. The synthesis of
hydrophilic epoxy resins may be illustrated by the following reactions with
both epoxy and hydroxyl groups,
oII
C-OH
R'"
'C-OH
II
o
oII 0
II
OH
O-C-R'-C-OH
+ *-CH 2-tH-CH 2-*
- *-CH 2-CH-CH
I 2- *
The carboxyl acid groups are, when neutralized, hydrophilic and the higher
the acid content the more readily dispersible these resins are in water
containing amines or inorganic bases. The degree of hydrophilicity IS
controlled by the extent of neutralization of the pendant acid groups.
9.3.1 Tooling
Epoxy resin tooling was initially used in the aircraft and car industries but
has now been extended to mould and tool making for many different end-
products, such as ship and boat building, domestic equipment and pottery.
A combination of techniques similar to those used for casting and laminating
are used for the production of moulds and foundary patterns. The resins are
usually formulated with a liquid resin cured with a room temperature
hardener. The properties are modified by the use of additives which not only
improve specific properties but also usually reduce materials' costs. Tools
can usually be fabricated in a fraction of the time required to machine metal
and the costs of epoxy tooling are much lower than conventional metal tools.
Epoxy resins are especially advantageous for the construction of prototypes
and tooling for short runs. Lee and Neville (1967) have a chapter on epoxy
resin tooling in which much detailed information is given which is still
relevant. Special resins for tooling applications can be formulated and
details are available from resin manufacturers. Some of the more important
tooling applications of epoxy resins include:
1. Moulds for various purposes such as form master moulds for vacuum
forming and injection moulding.
2. Jigs and fixtures to assist accurate positioning and checking shapes
and dimension of components for such operations as drilling, cutting,
and welding.
3. Metal-forming tools to replace the much more costly steel tools for
prototype and short run production. Examples of such tools are
stretch blocks, press tools, drop-hammer punches and dies. A rubbery
resilient facing on one of the tools of the set may be required, e.g. for
a drop hammer.
4. Foundry patterns ranging from master moulds from the original
pattern, duplicate patterns made from the master mould to core
boxes.
The benefits of using epoxy resins instead of traditional steel or aluminium
alloy as tooling materials are numerous. In addition to facile forming of
complex shapes the resins have low cure shrinkage and close tolerance can
be maintained. Also, the inert nature of epoxy resin to both heat, when
operational temperature is below its glass transition temperature, and
chemical attack, which may corrode steel, ensures shape stability. Other
important advantages include lower costs than for metals, lighter weight,
also the shape of the tool can be readily adjusted, and lead-times are shorter.
Epoxy resin formulations for tooling applications are normally room
temperature curable systems made up with liquid BADGE-type resins with
COATINGS AND OTHER APPLICATIONS 319
polyamines (normally aliphatic) as hardeners. Depending upon specific
application, diluents, plasticizers, fillers can also be introduced into the
system for the purpose of reducing cost and improving required physical and
chemical properties. In cases where special toughness is required, laminates
of epoxy resin with various fibres can also be used. Open mesh 'tooling
fabrics' are available and higher stiffness can be attained by the use of carbon
fibre reinforcement (chapter 8).
flooring. Seamless flooring has a 'low' viscosity due to the use of a reactive
diluent and can be applied in thickness of 10--60 mm. They can be used to
coat Portland cement floors, wood or metal after a specific surface
preparation before application of the resin. Industrial epoxy flooring or
epoxy concrete contains 75-85% particulate filler, usually sand (10--100
mesh). With high filler loadings the viscosity is high and the composition is
applied by trowelling with thickness of 3-25 mm CI8 to 1 inch). These
compositions have low shrinkage and may be used to repair cement floors.
Epoxy terrazzo floors contain coloured stone or marble chips and a fine
particulate filler, such as calcium carbonate, and are trowelled to give 6-
12 mm (I;cI;2 inch) thick coatings. After setting overnight they are wet-
ground to a polished terrazzo surface which is readily maintained.
The advantages of epoxy flooring may be summarized as follows:
1. Ease of seamless application, room temperature cure with low
shrinkage
2. Excellent adhesion to many substrates
3. Chemical resistant and durable
4. Tough, flexible and abrasion-resistant
5. Dust-free, skid-resistant and readily maintained.
9.3.2.2 Road and bridge coatings. Epoxy resin coatings have been used
to protect and repair road and bridge surfaces. Spalling is a commonly
encountered problem on the surfaces of roads and bridges. It is caused by
the combination and accumulation of many different factors such as freeze-
thaw cycles, spillage by vehicles, road gritting and de-icing operations.
Epoxy resins that are used for the treatment of these conditions are
usually liquids cured by amine hardeners at ambient temperatures. Coal tar
is used as extender and often in high proportions in their formulation. When
coated onto the structure and cured, the resin formulation forms a film
which prevents the spalling of concretes on the road surface and bridge
structure. Furthermore, the use of such formulations to seal cracks can
significantly slow down their growth by bonding the damaged sections into
an integral part. It also limits water penetration into the structure of the
bridge. Ingress of water to steel reinforcement or framework causes
corrosion. Hence epoxy resin coatings can give substantial protection
against such damage. Some other applications of epoxy resin road and
bridge coatings include:
1. Repair of scaling concrete roads with 'thin' coatings
2. Repair of 'slippery' but otherwise satisfactory asphalt or concrete
road surfaces
3. Fuel spillage protection
4. Lightweight surfacing for bridge decks
COATINGS AND OTHER APPLICATIONS 321
of course is often an advantage since the epoxy resins bond well to metal
inserts. At the moulding temperature the cured product is flexible and
ejector pins can cause damage to the moulding. This can be avoided by
suitable design so that excessive local stresses are avoided during ejection of
the moulded part.
Compared with casting the moulding of epoxy resins offers the advantages
of simpler handling, higher filler loadings and also fibrous reinforcement. Its
disadvantages include higher capital costs and storage problems with the
uncured moulding powder. Their shelf-life may be limited due to incipient
cure if the storage temperature is not low, for example it may be only several
weeks at 22°C but much longer in sealed packages refrigerated at 7°e.
Epoxy moulding powders are formulated with a mixture of solid or liquid
BPA resins which may be blended with epoxy novolacs. The latter improves
the flow properties and has higher functionality and hence decreases gel and
cure time. The curing agents are often diaminodiphenol methane (DDM) or
acid anhydrides (PA or THPA), but others may be used for specific
purposes, for instance chlorine-containing curing agents have been
suggested for improving flame resistance. The moulding powders contain
suitable particulate fillers, such as silica flour, and can be reinforced with
fibres, 1.5-25 mm (116 to 1 inch) long. Zinc stearate is often included as a
release agent, to assist ejection of the cured product from the mould.
Thermal conductivity can be increased by use of metal powder as fillers, but
should be avoided when the electrical properties are important. The density
of the filled resins ranges from about 1.6 to 2.2. Lower densities, 0.75-1.00,
can be obtained by the incorporation of microballons, composed of either
clay, U/F or P/F resins or glass (Lee and Neville, 1967a). A major
application of moulding powders is the encapsulation of electrical
equipment and electronic devices. It is essential that during moulding the
resin flows so that the component is not damaged or displaced. Typical
moulded parts include deposited carbon resistors; chokes; ceramic,
polyester and glass capacitors; metallized thin-film resistors; pulse trans-
formers; toroid, solenoid and relay coils; printed circuits, solid-state circuits
and many more products (Lee, 1986, 1988). Epoxy resin moulding powders
also have applications in chemical plant, such as valves and filter plates
where their chemical resistance offers major advantages. Structural
components with good thermal and chemical resistance coupled with good
mechanical properties and light weight are also moulded from epoxy resins.
9.3.4 Embedding
Embedding is used to protect and/or decorate electrical or electronic
devices. Allied and related terms are casting, potting, impregnation and
encapsulation. Moulding has been discussed in the section on moulding
powders. The electrical properties required for electrical applications have
COATINGS AND OTHER APPLICATIONS 323
been dealt with in chapter 6. Protective functions of the embedding resin
include exclusion of moisture, oxygen, salt spray, solvent, other chemicals
and microorganisms. The encapsulation of a device also protects it from
mechanical shock and vibrations and provides suitable locations for
mechanical handling. Alternative materials to epoxy resins include
silicones, polyurethanes, polyesters, polysulphides and allyl resins (Lee,
1988). The epoxy resins have low viscosities and shrinkage since volatile
compounds are not formed during cure. By suitable formulation it is
possible to obtain a range of properties, from the rigid to flexible,
elongations at break ranging from 3 to 40%. BPA resins and also mixtures
with epoxy novo lacs are used for potting. Cycloaliphatic epoxy resins have
good dielectric properties and excellent weatherability, which is due to the
absence of chlorine and aromatic rings. The cycloaliphatic resins require
acid anhydride hardeners and longer high temperature cures. The pro-
perties of the resin can be modified by the inclusion of fillers and flexibilizers
(chapter 4). Typical particulate fillers include silica, 'talc, clay and calcium
carbonate, mica and chopped glass. Metal powders can be used to increase
the thermal conductivity of the product. Lower density products can be
produced by the inclusion of micro spheres of saran, PIF resin or glass.
9.3.5 Miscellaneous
It is almost an impossible task to list all the existing applications of epoxy
resins since there are currently many new developments. However, it is
worth mentioning several areas in which epoxy resins are used, although on
a small scale, but nevertheless for important specialized applications, such
as textile finishes (Tanaka and Shiozaki, 1988).
9.3.5.1 Oil field applications. Epoxy resins are used in drilling operations
to consolidate poor well formations, especially where the geological
structures are primarily sandy. Resin compositions are adsorbed onto the
well walls and subsequently cross-linked. Such treatment is particularly
effective in minimizing sand production, which could eventually result in the
well being discarded.
9.3.5.3 Epoxy resin foams. Blown foams can be formed with epoxy resin
formulations where a blowing agent is present. During the initial stages of
324 CHEMISTRY AND TECHNOLOGY OF EPOXY RESINS
cure such an agent either evolves a gas or evaporates to leave voids in the
cured network. For example, hydrazide derivatives can be mixed into the
epoxy formulation; with the heat generated by cure or provided by heating
elements, they decompose to form nitrogen which is the blowing agent.
References
Allen, R.T.L. and Edwards, S.C. (eds) (1987) The Repair of Concrete Structures, Blackie,
Glasgow, UK.
Ash, M. and Ash, I. (1980-1983) Encyclopedia of Surfactants, Vol. 1-3. Chemical Publishing
Co., New York.
Barton, A.F.M. (1983a) CRC Handbook of Solubility Parameters and Other Cohesion
Parameters, Ch. 1,2. CRC Press, Boca Raton, USA, pp. 1-21. (1983b) Many references to
epoxy resins, see Index.
Becker, W.E. (1979) Reaction Injection MOUlding Van Nostrand Reinhold Co., New York.
Bonneau, M. and Burkhart, A (1991) MEPPE Focus (Jan.), pp. 1-6.
Brewer, G.E.F. (1985) Ency. Polym. Sci. Eng. (Suppl.) Wiley-Interscience, New York,
pp.675-fJ87.
Burkhart, A (1988) Hybrid Circuit Technology, pp. 222-228.
Chen, X.M., Ellis, B. and Seddon, A.B. (1992) Paper prepared for publication but submission
delayed until after patent application.
Ciba-Geigy (1988) Araldite Resins and Hardeners for Surface Coating Systems, Ciba-Geigy
Plastics, Cambridge, UK.
Coeling, K.J. (1985) Ency. Polym. Sci. Eng., Vol. 3, Wiley-Interscience, New York, pp. 567-
575.
Ellis, B., Chen, X.M. and Seddon, AB. (1991) British Patent Application.
Emerson, J.A., Sparapany, J.J., Martin, A.R., Bonneau, M.R. and Burkhart, A (1990)
IEEE, 40th Components and Technology Conference Vol. 1, pp. 6()()....605.
Grulke, E.A (1989) Polymer Handbook, VII 3rd edn ed. Brandrup, J. and Immergat, E.H.
Wiley, New York, pp. 519-559.
Hawkins, W.C. (1989) Enc. Polym. Sci Eng. 2ndedn Vol. 15, Wiley, New York, pp. 560-563.
Lee, H. and Neville, K. (1967) Handbook of Epoxy Resins, McGraw-Hili, New York, pp.I-16.
(1967b) Ch. 23, pp. 1-5, Ch. 24, 7-40.
Lee, S.M. (1986) Ency. Polym. Sci. and Eng. Wiley-Interscience, New York, Vol. 5., pp. 792-
828, see p. 818.
Lee, S.M. (1988) Epoxy Resins, Chemistry and Technology 2nd edn, ed. May, C.A, pp. 799-
810.
Manzione, L.T. (1989) Enc. Polym. Sci. Eng. 2nd edn. Wiley-Interscience, New York,
pp.72-100.
COATINGS AND OTHER APPLICATIONS 325
Martin, S.R.W. (1970) Paint Technology Manuals, Gen. ed. Taylor, C.I.A. and Marks, S.,
Part 3, collator Keenan, H.W., Chapman and Hall, London, p. 16.
Maslow, P. (1979) In Plastic Mortars, Sealants and Caulking Compounds. ed, Seymour, R.B.
ACS Symposium Series 113. American Chemical Society, Washington, D.C. pp. 39-60.
Misev, T.A. (1991) Powder Coatings, Chemistry and Technology. Wiley, Chichester, UK.
(1991a) pp. 304-324.
Nielsen, L.E. (1977) Polymer Rheology, Ch. 9. Marcel Dekker, New York, pp. 133-150.
Paul, S.l. (1986) Surface Coatings, Science and Technology, Wiley, Chichester, UK.
Potter, W.G. (1970) Epoxide Resins, Ch. 8, Iliffe Books, London, pp. 144-167.
Richart, D.S. (1982) Kirk-Othmer Enc. Chern. Tech., 3rd edn, Vol. 19. Wiley-Interscience,
New York, pp. 1-27.
Richart, D.S. (1985) Ency. Polym. Sci. Eng. Vol. 3, Wiley-Interscience, New York, pp. 575-
601.
Scharenberg, R.T. (1985) Ency. Polym. Sci. and Eng., Vol. 3, Wiley-Interscience, New York,
p.552.
Shell (1991) Technical Bulletins, Shell Chern. Co., Houston, Texas 77027, USA.
Shinoda, K. and Friberg, S. (1986) Emulsions and Solubilization, Wiley, New York.
Somerville, G.R. and Smith, I.T. (1973) Epoxy Resins, Chemistry and Technology, 1st edn,
Ch. 7, eds May, C.A. and Tanaka, Y., Marcel Dekker, New York, pp. 451-484.
Stephenson, R.C. (1989) Enc. Polym. Sci. Eng. 2nd edn (Suppl.) Wiley-Interscience, New
York, pp. 867-870.
Tanaka, Y. (1988) In Epoxy Resins, Chemistry and Technology, 2nd edn, Ch. 2, Table 49,
pp. 212-231.
Tanaka, Y. and Shorzaki, H. (1988) In Epoxy Resins, Chemistry and Technology, 2nd edn,
Ch. 12. ed. May, C.A. Marcel Dekker, New York, pp. 956-1048.
Tess, R.W. (1988a) Epoxy Resins, Chemistry and Technology, 2nd edn Ch. 8. ed. May, C.A.
Marcel Dekker, New York, Ch. 8, pp. 719-782. (1988b) see pp. 726-735.
Utracki, (1988) In Rheological Measurement eds Collyer, A.A. and Clegg, D.W. Elsevier,
Amsterdam.
General index
For reference to specific curing agents and hardeners see the separate index