You are on page 1of 45

Composites Science and Technology 31 (1988) 179-223

Failure Mechanisms in Toughened Epoxy Resins A


Review

Amar C. Garg
Department of Aeronautical Engineering, Indian Institute of Technology,
Powai, Bombay 400076 (India)

and
Yiu-Wing Mai
Department of Mechanical Engineering, University of Sydney,
Sydney, NSW 2006 (Australia)

(Received 27 February 1987; revised version received 5 June 1987;


accepted 27 October 1987)

SUMMARY

The subject of improving the fracture toughness of brittle epoxy resins is


receiving significant attention in order to improve the design strain of fiber-
reinforced composites for aerospace structural applications. Various rubber-
modified and particle-filled epoxy resins have been considered as candidate
materials. Such modified resins have been observed to yield a ten- to thirty-
fold increase in fracture toughness compared to the unmodified material In
order fully to utilize the potential of such materials, it is necessary to
understand the failure mechanisms leading to the improvement in toughness.
This paper provides a critical review of the existing theories that have been
proposed for the various toughening mechanisms related to modified epoxy
resins.

1 INTRODUCTION

Epoxies are among the best polymeric materials being used in fiber-
reinforced composites and in many structural parts. They are also used as
179
Composites Science and Technology 0266-3538/88/$03.50 © 1988 Elsevier Applied Science
Publishers Ltd, England. Printed in Great Britain
180 Arnar C. Garg, Yiu-Wing Mai

adhesives, tooling compounds, molding powders, and as potting and


encapsulating materials. However, maximum use of epoxies is being made in
the aviation industry as adhesives and as matrix materials in fiber-reinforced
composites for structural applications. Like metals, many of these materials,
and in particular fiber-reinforced composites, are notch-sensitive and lose
much of their structural integrity when damaged. Damage can occur either
at the time of manufacture or during service. Generally it initiates from pre-
existing defects (such as those due to fabrication processes, for example, or
resulting from impact damage due to dropping tools, etc.) or stress
concentrators (such as holes, corners, free edges and ply terminations).
During the application of load, damage initiation and propagation in a
laminated composite occurs in a series of events whereby individual modes
develop and interact with each other? The interaction of damage modes is
dependent on the relative orientation of plies within a laminate, the mode of
loading and the state of damage in a laminate. There are three most
important damage initiation modes in a laminated composite, viz. matrix
cracking, delamination and fiber fracture. The first two modes depend to a
large extent on the properties of the matrix.
Matrix cracks initiate in plies having tensile stress applied perpendicular
to the fibers. For example, the matrix cracks initiate in the 90 ° plies of quasi-
isotropic or cross-ply laminates at 25-30% fracture stress. ~- 5 This event is
often referred to as first-ply failure. 6'7 As the load is monotonically increased
or cycled, matrix crack initiation continues in 90 ° plies and in other off-axis
plies until a stage of crack saturation called the 'characteristic damage state'
(CDS) t is reached. During the development of the CDS, matrix cracks in
adjacent off-axis plies connect by growing short cracks along the matrix-rich
interface. The interfacial crack growth and subsequent coalescence with
cracks in adjacent off-axis plies lead to the development of delamination
under continued loading. Delamination also initiates from zones of high
interlaminar stresses such as free edges, notches, ply terminations, cut-outs
and other geometric discontinuities. Delamination may also develop during
manufacture as a result of incomplete curing, or through the introduction of
foreign particles, or as a result of impact damage, s'9
The growth of delamination is a major concern to the designers of
composite structures, as it redistributes the stress in the plies of a laminate
and may influence residual stiffness, residual strength and fatigue life.
Though it may not lead to catastrophic failure of a component, it is observed
to be the most prevalent life-limiting damage growth mode l's --lo and is one
of the major obstacles to the application of advanced composite materials in
primary aircraft structures.
Since delamination (a matrix-dominated failure mode) may have such
adverse effects on fiber composites, methods involving partial interlaminar
Failure mechanisms in toughened epoxy resins--a review 181

bonding have been developed to improve delamination fracture re-


sistance) ~'12 Alternatively, and more commonly, composites with tough-
ened matrices are used to overcome this problem. The currently available
epoxide resins are brittle and have Mode I fracture energies o f about 80-
3 0 0 J m -2, either in bulk ~3-~9 or in the delamination m o d e in a
composite, 2'1°'2°-26 or as an adhesive. ~5-18,27-30 On the other hand, the
fracture energy for crack propagation perpendicular to fibers may vary
between a few kJ m - 2 and several tens of kJ m - 2, depending upon the fiber
type. 31 The use of such brittle resins in composites imposes restrictions on
the structural designers. For example, an aircraft wing structure made with
fiber-reinforced composites is limited to a design strain of about
0"4%, 8'9'a2. which thus seriously limits realization of the full potential of
weight reductions offered by composites. By contrast, a tougher resin would
enable the allowable strains to increase to 0"6% or more. This would result in
a significant reduction of structural weight and fuel consumption. 8'9'32. In
addition, the development of toughened resins as adhesives would improve
joint strength. 33 In summary, toughened epoxy resins should improve
damage tolerance and lead to more durable and reliable structures.
In consideration of the application of resins in aircraft structures,
Johnston ~4 has recently specified the following desirable properties of
resins:
- - T h e r m a l performance range - 5 4 ° C to 93°C
--Solvents Resistance to dissolution and swelling
in organic solvents in stressed state
--Moisture Tg values sufficiently high to allow
satisfactory hot/wet strengths at 93°C
- - A u t o c l a v e processability Maximum 177°C/1.4 MPa
- - F r a c t u r e energy (G~c) 1.9-3 k J m 2
--Stiffness Young's modulus > 2.8 GPa; shear
modulus > 1"4 GPa
It is obvious from these specifications that the required fracture energy is too
high to be met by brittle epoxy resins. Thus, in order to achieve the desired
toughness of the resins, a significant a m o u n t of work is in progress in various
organizations around the world. There are two basic solutions for
improving the fracture toughness. One is to use high toughness thermo-
plastics 13'14 such as polyether ether ketone (PEEK). This resin melts at
330°C and its processing temperature is 400°C. The toughness of fiber-
reinforced P E E K composites is seen to be about ten times that of composites

*James and Harvill is° have recently shown that for wing structures a further weight saving
of some 8-12% can be achieved if the allowable strain can be increased from 0.4% to 0.6%.
182 Amar C. Garg, Yiu-Wing Mai

made with epoxide resins. Although it seems an attractive resin, the


structural fabrication cost may be prohibitive as a result of the high
processing temperature. However, the use of PEEK may be limited to critical
components requiring very high toughness. The other alternative is to
modify existing epoxies which already possess many of the desired
properties but are lacking in toughness. The latter is considered to be a more
attractive solution, existing epoxies being modified by the addition of either
rubber or inorganic fillers such as alumina and silica particles. In this process
rubber or inorganic filler in a minor proportion, typically between 5 and
20%, is incorporated as a dispersed phase in a rigid epoxy matrix. The
resulting composite has a higher fracture resistance (rubber-toughened
epoxy having much higher toughness, 2.0-4.0 kJ m-2, than particle-filled
epoxy, 0.5-1-0 kJ m -2) than the unmodified epoxy resin; impact strength,
elongation at break and work to break are also all increased. However, such
a gain is associated with some loss of modulus, tensile strength, try, and glass
transition temperature, Tg. On the other hand, these parameters increase
slightly in particle-filled epoxies.
The increase in fracture energy of a modified epoxy depends upon its
ability to dissipate work by various deformation mechanisms. In fact, a great
deal of controversy exists over the exact nature of work-dissipating
mechanisms that occur in such materials: that is whether the energy is
absorbed by the rubber or the matrix and whether there is crazing or voiding
at the matrix. The purpose of this paper is to review the existing toughening
mechanisms of rubber-modified and particle-filled epoxies. The fracture
mechanisms of unmodified epoxies will not be studied here since several
review papers on this subject have already been published. 34-4° Similarly,
rubber toughening of resins other than epoxies, such as polyesters and
polyimides, is excluded as this can be found in Refs 41-45.

2 EPOXIES

2.1 Unmodified epoxies

Any molecule containing the epoxy group - - C C - - is called an epoxy. '~6


The epoxy group when bonded chemically with other molecules forms a
large three-dimensional network. The process by which chemical bonding is
achieved is called curing, during which the fluid resin changes to a solid.
Various amines, anhydrides and Lewis acid (boron trifluoride) are employed
as curing agents to cure the epoxies. The properties of the resulting epoxide
resins depend upon the epoxy, the curing agent and the curing process. 47 -49
Failure mechanisms in toughened epoxy resins--a review 183

For example, when curing agents such as aromatic amines, acid anhydrides,
BF3MEA and dicyandiamide are used with the right epoxy resins and fully
cured, the epoxies attain high glass transition temperature, T~, and are thus
suitable for high-temperature applications.
For aerospace applications, only the resins that can withstand high
temperatures (~ 177°C) are acceptable. Such epoxies are polyglycidyl ethers
of the Novolacs (Union Carbide ERR 0100), triglycidyl p-amino phenol
(Union Carbide ERLA 0510) and tetraglycidyl diamino diphenyl methane
(TGDDM, Ciba-Geigy MY 720). The suitable curing agents are dicyan-
diamide (DICY), BFaMEA, diamino diphenyl sulfone (DDS, Ciba-Geigy
EPORAL), methyl dianiline (MDA), meta phenylene diamine (MPDA) and
piperidine. The chemical compositions of such epoxies and curing agents are
presented in Appendix A.
During the curing of the epoxide resins, the crosslinking between the
epoxy molecules and reactive groups on each end of the curing agent occurs.
The amount of crosslinking affects the resin properties significantly. Lower
crosslink density improves toughness by permitting greater elongation
before breakage and also results in reduced shrinkage during cure. On the
other hand, higher crosslink density yields an improved resistance to
chemical attack and raises Tg but lowers the failure strain.

2.2 Modified epoxies

Two major toughening methods have been considered for the epoxies. The
details of these are described below.

(a) Rubber-toughened epoxy


The toughening of thermoplastics by rubber has been in existence for over
40 years 5° but the addition of rubber to thermoset resins (epoxies) is only
about 15 years old. It was the work of McGarry et al. sl-53 which showed
that worthwhile improvements could be effected by adding certain liquid
rubbers to epoxide formulations. Since then a significant amount of infor-
mation on toughened epoxides has appeared in the literature. 15- 18,54-64
The type of rubber to be added to toughen the resin is determined by two
factors. 5° Firstly, there is a compatibility requirement: the rubber must
dissolve and become dispersed in the resin, but precipitate out when the resin
begins to cure. Secondly, there is a chemical requirement: the rubber must
react with the epoxide group. Both such requirements are met by carboxyl-
terminated butadiene-acrylonitrile (CTBN) rubber (see Appendix B) and
therefore it is widely used as a toughening agent for epoxies. To a lesser
extent both amino- and vinyl-terminated butadiene-acrylonitrile (ATBN
and VTBN) have also been used to toughen epoxies. '.3"65-6v'69 In addition,
184 Amar C. Garg, Yiu-Wing Mai

attempts have been made in using siloxane and fluorocarbon elastomers for
toughening. 6s.68
To manufacture a toughened epoxy resin, the rubber is first dissolved in
the resin at 50-60°C and the solution is subjected to vacuum to remove
trapped air bubbles. The hardener is then dissolved with stirring, and the
resin-rubber-hardener mixture is cast. A second application of vacuum
removes air bubbles entrapped at this stage, but care is necessary with a
volatile hardener such as piperidine. The resin is then cured following a
standard cure cycle.
In the modified epoxide resin, the size or diameter, d, and volume fraction,
Vp, of the dispersed rubber particles depend on the curing process, rubber
content and acrylonitrile content in the rubber. Slow-cure resin systems tend
to form large rubber particles between 1 and 10~tm in diameter. In some
cases, smaller particles less than 0.5/~m in diameter are also formed, s°
probably at the later stages of the cure reaction when molecular mobility is
reduced. Rapid-cure resins tend to form small rubber particles. An increase
in rubber content leads to an increase in d a n d Vp.s7"69 However, rubber as a
dispersed phase occurs only when the rubber content is less than 20 wt %.
Above this concentration rubber blends with epoxy 16 or it acts as a matrix
and the epoxy separates as spherical particles, s° Depending on the solubility
characteristics of the epoxy resin, increasing the acrylonitrile content over a
certain range reduces Vp and d. Also the fracture toughness of modified
epoxy is affected by acrylonitrile content. This is shown to be highest with
acrylonitrile contents between 12 and 1 8 w t % . 6a

(h) Particle-/illed epoxide res#l


Besides using rubber as a toughening agent for polymers, several inorganic
fillers such as alumina, silica, barium titanite, dolomite, glass beads and
aluminum hydroxide have been consideredf °-81 The composite resin is
manufactured by mixing the fillers directly with the resin and hardener and
then curing in a vacuum oven. Inorganic fillers have also been added to
rubber-toughened epoxies s2-86 so that high toughness, high strength and
high modulus may be obtained.

3 P R O P E R T I E S OF T O U G H E N E D EPOXIES

3.1 Rubber-toughened epoxy

The addition of rubber to a brittle resin modifies its various characteristics.


For example, it causes a reduction in stiffness, lowers T~, plasticizes the
matrix (if the rubber is not fully precipitated from solution in the resin),
Failure mechanisms in toughened epoxy resins--a review 185

reduces the yield strength and increases the linear thermal expansion
coefficient; but it also significantly increases the fracture resistance. Because
of such a gain in toughness, rubber-toughened epoxies are being used for
various engineering applications.
The changes in properties of a toughened epoxy resin depend upon
various parameters, such as the size and volume fraction of dispersed rubber,
the strain rate, temperature and other environmental factors. Though a
significant amount of information on the physical and mechanical behavior
of toughened thermoplastic materials like high impact polystyrene (HIPS)
and acrylonitrile-butadiene styrene (ABS) is available, 5° the information on
thermoset resins such as epoxies is rather limited. In particular, very little
information is available for high temperature epoxies toughened with
rubber.
T a b l e 115'16 shows the variations of initial tensile modulus, E, tensile
strength, a t, linear thermal coefficient of expansion, ~, and glass transition
temperature, Tg, with elastomer concentration for CTBN-toughened
D G E B A (DER-332 Dow Chemical Co.) epoxide resin. It is seen that E
decreases sharply at first, up to about 4.5% elastomer. Above this
concentration, it remains almost constant up to about 15% CTBN and then
falls sharply again. Similarly, a t and Tg both decrease with increase in
elastomer content (except Tg at 30% CTBN). However, the opposite trend is
true for cc These results are slightly different from the behavior of two similar
CTBN-toughened DGEBA (MY 750 and GY 250 Ciba-Geigy) epoxide
resins studied by Scott and Phillips 3~ and Lahiff, sv who noticed only a
marginal change in E and Gt with rubber content. These properties are also
affected by temperature and loading rates, o t and E are significantly reduced
with increasing temperature. 63,87 The effect of strain rate on E is seen to be
marginal but o-t increases with increasing strain r a t e . 62'63'88,89
TABLE 1
Mechanical Properties of a Rubber-modified Epoxy
(CTBN-toughened D G E B A ) ~6

CTBN Fracture Tensile Tensile Linear thermal Glass


concentration energy, Gic strength, a t modulus, E coefficient o f transition
(%) ( k J m - 2) (MPa) (GPa) expansion, ~ temperature, Tg
(per ~C × I0 5) (°C)

0 0' 12 72 3"3 7-8 80


4"5 1"05 70 2"3 8'0 70
10 2"72 56 2-2 8"7 65
15 3"43 45 2"0 9'6 62
20 3"59 20 1.0 10"2 60
30 2"00 17 0" 1 14.0 67
186 Amar C. Garg, Yiu-Wing Mai

35

30

25
z
20

1.5

w O.5

~. O0
0 5 10 15 20 25 30
PHR RUBBER
Fig. 1. Variation of specific fracture energy, G~¢, with rubber content. (+) Yee and
Pearson; 8s ( I ) Bascom e t al.; 16 ( 0 ) present work.

The fracture energy, Gic, increases with increasing elastomer concen-


tration up to about 16-17% and then decreases (Fig. 1 and Table 1). Such
behavior is seen as a result of the change of the elastomer from a dispersion
phase to a blend at high CTBN concentration. 16 These observations are in
agreement with those reported by Yee and Pearson 88 and Lahiff. 87 In
addition, Lahiff's data 87 for another CTBN-toughened epoxy (GY 250) also
support the trend for the toughness reported by Bascom e t al. 16 (see Fig. 1).
While increasing the temperature increases the fracture toughness 62'8v and
the fracture energy as shown in Fig. 2, increasing the strain rate decreases
toughness.~ v,62
The m a x i m u m fracture energy of rubber-modified epoxy is approxi-
mately 30 times that of the unmodified epoxy. 16 But despite this large
increase in the toughness of the bulk resin the same increase is not obtained
when the modified resin is used as an adhesive or as a matrix in fiber-
reinforced composites. 14-1s3~ A few such results, presented in Table 2, 31
show that the initiation fracture energies, Go, for the composite
(unidirectional carbon fiber-reinforced composite) made from unmodified
resin and the adhesive are very similar to that for the bulk resin. However,
the initiation fracture energy for the composite with a modified resin matrix
shows only a modest increase, whereas the increase for the bulk resin is
considerable. Also the toughness increase for the adhesive (glue-line
thickness --~200/~m) is less than that of the bulk resin. The suppression of
toughness in such cases has been explained ~5"16"3~ on the basis of the
dependence of fracture energy on epoxy film thickness. The full toughening
effect cannot be developed if the epoxy film is too thick, as in adhesives and
composites. 3~ Hunston e t al. ~° recently showed that the toughness in
Failure mechanisms in toughened epoxy resins--a review 187

-- 3.0

~ 2.5
~ 2.o
g T

to

0.5
o.o
-60 -~0 -20 0 Z0 ~0 60
TEMPERATURE('C)
(a)

t,5

/+.0

-- 3.5

~" 3.0

~- 2.5
~ 20
1.5

05

O0 . . . . .
-60 -40 -20 0 20 /+0 60
TEHPERATURE ('C)
(b)
Fig. 2. (a) Variation of fracture toughness, K=c,with temperature. ( , ) Kinloch et al.; 62 ( 0 )
present work. (b) Variation of fracture energy, Glc, with temperature (GY 250/piperidine/
CTBN X- 13).

rubber-toughened adhesives increases rather sharply with bond thickness to


a maximum value and then decreases with further increase of bond
thickness. They suggested that very thick bonds give toughness results
similar to those of bulk specimens. When the bond thickness decreases the
increasing constraint causes an increase in the deformation zone length
down the bond layer so that the toughness is increased. As the bond
thickness continues to decrease a point is reached where the deformation
zone volume is decreased so that the toughness decreases after passing
188 Amar C. Garg, Yiu-Wing Mai

TABLE 2
Fracture Toughness for Bulk Resin, as AdhesiVe Film and as
Matrix in Carbon Fiber-reinforced Composites 31

Rubber content Gc (kJm 2)


(%)
Bulk resin Adhesive film a Composite

0"0 0"33 0.28 0.28


3.2 b 1.4 0.33 0.37
6.2 b 2.2 1.35 0.36
9"0 c 3"2 1.50 0.49

a Adhesive film thickness ~ 200/am,


b Resin M Y 7 5 0 e p o x y + C T B N with acrylonitrile content
24.2% + 6% piperidine.
CResin M Y 750 epoxy + C T B N with acrylonitrile content
28.2% + 6% piperidine.

through a maximum. This explanation is physically sound and is in


agreement with that offered in Ref. 31.
In the preceding paragraphs, it was seen that the addition of rubber to
epoxy resin causes a significant increase in toughness. Such a trend is not
observed, however, when modified epoxy specimens are subjected to cyclic
stresses. This was shown by Shah et al. 54 by testing various specimens of
CTBN-modified D G E B A (DER 311, Dow Chemical Co.) epoxide resin.
Their results show that the fatigue crack propagation (FCP) rate is not
affected by the presence of rubber particles in the epoxy. Thus the addition of
rubber to epoxy resin is not expected to increase the fatigue life. This seems a
surprising result but such an observation has been explained by Hertzberg
and Manson 91 in terms of two competing mechanisms: matrix plasticity,
which would decrease fatigue crack growth rate, and reduced elastic
modulus (due to addition of rubber particles), which would increase fatigue
crack growth rate. This behavior is different from that of other rubber-
modified polymers, such a polyvinyl chloride (PVC), nylon and polymethyl
methacrylate (PMMA), where the increased elastomer concentration causes
reductions in fatigue crack growth rates. 91

3.2 Particle-filled epoxies

Particulate fillers are often used to reinforce epoxies for various engineering
applications in order to improve several mechanical and physical properties.
The fillers reduce cost, degree of shrinkage, exothermic temperature rise and
coefficient of thermal expansion, and raise the thermal conductivity and T~.
Failure mechanisms in toughened epoxy resins--a review 189

The addition of these fillers to the resin also affects its mechanical properties.
The stiffness, strength and fracture toughness of particle-filled epoxy resins
are higher than those of pure epoxy resin. 77'78 The changes in properties
depend upon the type of filler, particle size, volume fraction, surface
treatment of the particles, strain rate and the test environment. Various
studies of the fracture toughness (in terms of the critical stress intensity
factor, K~¢) of the particle-filled epoxy as a function of particle volume
fraction, V~, and surface treatment during fast 7°'75- 77,81 and slow 71 crack
propagation show that the K~¢ value of the composite is increased. In
addition, increases in Young's modulus, E, and yield stress, ay, have also been
reported. (See Table 3.)

TABLE 3
Influenceof Fiber Particle Sizeand VolumeFraction on E, % and K~ of an
Epoxy-Silica Composite 77

Volume Mean particle Kit E try


J?action, Vp size, d (MPam in) (GPa) (MPa)
(%) Ilm

40 300 1"76 + 0 ' 0 1 -- --


160 1 ' 7 4 _+ 0 - 0 2 -- --
100 1"87+0'01 -- --
60 1"83 + 0 " 0 2 -- --

0 -- 0-60 3-5 100


20 -- 1'30 6" 1 108
30 -- 1-62 7"7 121
40 -- 1'87 9"8 133
50 -- 2"21 12'5 150

40 -- 1 "90 -- --
(silane-treated)

The recent studies of Moloney et al. 77 have indicated that KI¢ of the
composite is little affected by the type of filler particle and its size (see Table
3). Their studies show that the matrix/particle adhesion, i.e. resulting from
treatment of particles with silane coupling agents, does not improve K~c
appreciably. Similarly, Broutman and Sahu 92 have shown that for a glass-
bead-filled epoxy resin, K~cfor treated and untreated glass beads coated with
silane coupling agent remained the same. By contrast, treatment with
silicone resin doubles the fracture toughness as a result of resin/filler
debonding and resin plasticization by the silicone. 92 In further work on
glass-filled epoxy resin, Spanoudakis and Young 76 show that although Kic is
not much affected by surface treatment the fracture energy, G~c, is. Poor
190 Amar C. Garg, Yiu-Wing Mai

16
50°C
lZ,

12

10 o
OC

08

-70°C
0,6

0~,

02

0 [ . _ ~ [ ~ I I t
01 02 03 0.~, 05 06
VOLUME FRACTION,Vf
(a)

5I 50Oc
4

J
2
x

I L 1 L I

01 02 03 04 05 06
VOLUME Ft~A£TIONVf
(b)
Fig. 3. Fracture energy, G,~, against volume fraction, Vr, of glass particles for (a) an
unmodified epoxy and (b) a rubber-modified epoxy. (After Kinloch et al. 84)
Failure mechanisms in toughened epoxy resins--a review 191

adhesion gives large Gie a s a result of a reduction in Young's modulus, E(i.e.


Gic = K ~ / E ) . However, good bonding is necessary for good strength. So, like
most brittle fiber-brittle matrix composites, high strength and high G~c are
mutually exclusive. 93'94
A simple and effective way to improve fracture energy without loss of
strength and elastic modulus is to incorporate inorganic fillers in rubber-
toughened epoxies. Figure 3(a) and (b) shows that GIe for the hybrid glass-
rubber-modified epoxy is much larger than for the glass-unmodified epoxy for
a given temperature and volume fraction of glass. Further work carried out by
Low et al. 95 confirms the advantage of adding fillers to rubber-toughened
epoxies. Table 4 shows that additions of metastable zirconia particles and
short alumina fibers considerably improve the fracture resistance of
unmodified epoxy (GY250 Ciba-Geigy). Unfortunately, the zirconia
particles do not give transformation toughening in this epoxy matrix as one
would expect in a stiffer matrix material like alumina.

TABLE 4
Mechanical Properties of Hybrid Particulate Epoxies 95

Material system Elastic Fracture Critical


modulus, E energy, Gic stress intensi O,
(GPa) (kJm - 2) .factor, Kxc
( M P a m 1/2)

Epoxy 2"78+0'28 0.23+0.10 0.80+0.10


Epoxy-rubber (100/15) a 2'18+0-16 2.24+0.10 2.24+0.05
Epoxy-rubber-zirconia (100/15/25) a 2'25+0'07 3.54+0.10 2'84_+0.50
Epoxy-rubber-alumina fiber (100/15/19) ~ 3"80+0.26 4.19+0.60 3.99-+0.60

a Weight percentage.

Though considerable information is available on the mechanical behavior


of particle-filled polymers,* very little is known about fatigue crack
propagation in such materials 91 and particularly the hybrids. In a recent
study, Gadkaree and Salee v2 have shown that the fatigue crack growth rate
for a bisphenol-A-terephthalate/isophthalate copolymer filled with an
alumino-silicate (particle size .,~200pm)decreases with increasing filler
content. Interestingly, for low filler contents, i.e. up to about 10%, the fatigue
crack propagation rate follows the Paris power law 96 but for high
concentrations it does not.

* For a review of the parameters affecting the strength and toughness of particlc-tilled epoxy
resins, see Moloney et al.l 5
192 Amar C. Garg, Yiu-Wing Mai

4 FAILURE MODES AND STABILITY OF CRACKING IN


TOUGHENED EPOXIES

4.1 Stability of cracking and crack growth behavior

Generally, both unmodified and toughened epoxies show similar crack


growth behavior. Three basic types of crack growth have been observed: (i)
brittle stable crack growth in which cracking is continuous; (ii) brittle
unstable crack growth during which the crack proceeds in a stick-slip
manner; and (iii) ductile stable crack growth in which cracking is continuous
but is dominated by gross plasticity in the specimen. Typical load-
displacement records for these types of crack growth behavior in compact
tension specimen geometry are given in Fig. 4 and are termed, respectively,
types C, B and A by Kinloch et al. 62 In between these three basic types,
transitional crack growth behavior is also reported (see also Fig. 4). Type A
crack growth is only observed at high test temperatures, say above 40°C, and
under these conditions it is questionable whether valid K~ values can be
obtained. For example, in a certain hybrid particulate epoxy composite
Kinloch e t al. 8~ obtained K ~ 3 " 5 M P a m - 1 / 2 and O-y<50MPa. The
required thickness must be larger than 2-5(K~jo-y)2, which is about 12ram
and twice the actual specimen thickness. A possible mechanism for this type

~00 TYPE C L00F TYPE E-B ~OOr TYPE B


Pc
z
-- 2 0 0
Q-

1 2 0 1 2
A (ram) A (ram) A (ram)

600 TYPE A-B 600

z
o~
coo

2O0


/ ~
z
Q_
~00

2OO

o i

I - - 8 I~
A(mm) A(mm}

Fig. 4. Load (P) versus deflection (A) curves for toughened epoxies associated with the
different types of crack growth in the compact tension geometry. Type C through to type A
occur with increasing temperature and decreasing stress rate of test. (After Kinloch el a/. ('2)
Failure mechanisms in toughened epoxy resins--a review 193

of ductile crack growth has also been suggested for rubber-toughened


e p o x i e s 97 in terms of the meniscus instability model of T a y l o r . 98'99
Although it is claimed that there is good agreement between the theoretical
and experimental values of the critical wavelength of the finger-like
meniscus instabilities, doubts must be raised about the calculated plastic
zone thickness based upon apparently invalid K~c measurements. Using
K~¢ ~ 3"5 MPa m -1/2 and % ~ 3 4 M P a 97 the required thickness for K~c is
about 27 ram, which again is much larger than the actual thickness (,-~6 mm)
used for the experiments. Under ductile fracture conditions the extent of the
plastic zone thickness is geometry-dependent and neither Kxc nor Gt¢ can be
meaningfully used to describe the fracture process. The specific essential
work of fracture dissipated in the fracture zone can be obtained by the
Cotterell-Mai experimental techniques. 1oo- 1o3 Such essential fracture work
is expected to be a material constant for a given sheet thickness. 1°4 The
occurrence of stable continuous type C crack growth is promoted by low
temperatures and high strain rates; conversely, type B 'stick-slip' unstable
crack growth is favored by high temperatures and low strain rates. The
addition of rubbers and other fillers can also change what is otherwise stable
continuous crack growth in unmodified epoxies to unstable stick-slip crack
growth in modified epoxies. 62'84 There is considerable debate as to the
mechanics and mechanisms of stick-slip crack growth and this concerns the
very important problem of crack stability which is discussed below.
The stability of cracks has been studied by a number of investi-
gators, 19'96"1°5-113 and it is shown that this depends on both the fracture
toughness of the material and the specimen geometry used. For cracking to
be stable the following mechanics equation must be satisfied, a°5`1°v'11~ i.e.

1 dG c > gsf (1)


G c dA -
where G c is the specific fracture resistance, A is the crack area, and gsfis the
geometric stability factor depending on the testing machine constraint. Thus
d2C/dA 2 2 dC
g s f = dC/dA " C dA (2a)

for displacement control, and


d2C/dA 2
gsf- dC/dA (2b)

for load control, where C is the specimen compliance defined as the load-
point displacement per unit load. Clearly, dGc/dA is a material property and
gsf is a function of the specimen geometry. The more negative the gsf, the
194 Amar C. Garg,Yiu-WingMai

more stable is the cracking. Characteristic values of gsf have already been
tabulated for a variety of testpiece geometries in Refs 96 and 106. So far two
explanations have been given for unstable 'stick-slip' cracking in epoxies.
Both require eqn (1) to be violated, i.e. dGc/dA is less than the gsf. W h y Gc
changes with A probably explains the mechanism of crack instability. Mai
and Atkins 19'96 first suggested that in terms of crack velocity, A, and crack
acceleration, ,4, eqn (1) can be rewritten as
1 dG~ ~I"
G~ dA A > gsf (3)

If G¢ decreases with crack velocity, A, sufficiently so that eqn (3) may not be
satisfied, then unstable cracking will follow. Indeed, Mai and Atkins 19 and
Andrews and Stevenson 113 have shown that for certain epoxies dG¢/dA is
negative. Kinloch and Young ~4 are uncertain whether this is the cause or
the consequence of 'stick-slip' crack propagation. Partly this is because
Gledhill and Kinloch ~ 5 could not p r o m o t e stable cracking by improving
the gsf and partly because they only obtained positive dGc/dA in the epoxies
they studied. It must be stressed, however, that cracking can only be made
stable in otherwise unstable crack systems if eqn (1) is satisfied. If dGJdA is
very negative the gsf may have to be very m u c h improved with external
stabilizers before any stable cracking can be obtained. 19'96 During 'stick-
slip' the crack velocity is not uniform and only a mean crack velocity
between upper and lower values can be estimated) 1° The corresponding
mean fracture toughness is given by G c = Q Gx/~G~, where G~ and GA are the
initiation and arrest toughnesses, and Q is a function of the specimen
geometry. 96 The positive dGJdA results obtained by Gledhill and Kinloch
only refer to the initiation,which must be higher for higher crosshead speeds
of the testing machine. There are no measurements for K~¢ and A during the
jump. Based u p o n these arguments it is felt that the negative dGJdA
explanation for unstable 'stick-slip' cracking in epoxies cannot be
completely ruled out. In fact this proposal has been recently supported by
Leevers 1lO and Maugis, ~~2 who both assume that if G~ decreasing with A is a
material property, 'stick-slip' occurs when a mean crack velocity is imposed
within this region. Both authors also discuss the effects of specimen
geometry (i.e. gsf) in controlling 'stick-slips'. Why G¢ should decrease with A
is not clear. Because of the low crack velocity at which stick-slip can be
observed in epoxies Maugis 112 suggested that the mechanism may be one of
internal friction ~16 and not one of isothermal-adiabatic transition*) t7

* In Ref. 117 the reference to isothermal-adiabatic transition as causing 'stick-slip' in epoxy


resin is incorrect. The instability crack velocity calculated is 0.2 m s-1 and is an order of
magnitude bigger than the observed velocity of ~0"01 m s-~.
Failure mechanisms in toughened epoxy resins---a review 195

An alternative explanation for 'stick-slip' crack propagation in epoxies is


offered by Kinloch and Williams, 118 who propose a crack tip blunting
model. Indeed, direct observation of the crack tip deformation in a SEM by
Bandyopadhyay et al. 59 shows that blunting occurs by localized shear
yielding and at fracture the shear yield strain is about 0.4. No crazing is
observed. The mechanics of crack instability is as follows. The crack tip
blunts by localized plastic flow and this relieves the effective stress intensity
factor at the tip so that a higher applied stress intensity is required to cause
crack initiation. Since the elastic energy stored in the blunt crack system is
much larger than that absorbed when the sharp crack propagates, i.e.
dGc/dA < 0, eqn (1) may be violated unless the gsf is negative enough to
offset the negative dGJdA effect.
In principle, it is possible to stabilize the crack provided the excess energy
can be absorbed elsewhere; but in practice it may be difficult to achieve this.
The mechanism for the negative dGJdA results from the varying degree of
plastic blunting for different epoxies. Low yield-stress epoxies have larger
crack tip radii and vice versa. Figure 5(a)--(c) shows the variation of the
K~c/K~c~ ratio with the crack tip radius, v/p, for the unmodified DGEBA
piperidine-cured epoxies, rubber-toughened epoxies and hybrid glass
particle-toughened epoxies, respectively. The theoretical curves for Kic/K~cs
are calculated from Refs 118 and 119:
Kic aaX/~ (1 + p/2c) 3/2
(4)
K,cs O"tc2%//~ (1 + p/c)
where K~cs refers to the propagation of a 'sharp' crack satisfying the
condition of a critical stress, ate , at a critical distance, c, ahead of the crack
tip, 12° a a is the applied stress and a is the crack length. Figure 6(a) and (b)
shows the dependence of Klc/Kic s o n the tensile yield stress, aty, of the
unmodified, rubber-toughened and hybrid particle-toughened epoxies. The
transition from stable cracking to unstable stick-slip crack propagation for
the unmodified and rubber-toughened epoxies occurs at O'yt ~ 9 5 MPa,
indicating that there is a possible common deformation mechanism in these
two materials. For the hybrid particulate composites this transition takes
place at aty < 120 MPa, which suggests that the addition of inorganic fillers
to epoxies makes cracking more readily unstable. 84 The transition from type
C to type B crack growth is characterized by a critical radius, Pc, of the blunt
crack tip which can be estimated from
Pc = tK~/ Eaty (5)
by assuming p = 6, the crack tip opening displacement. I~sA19 Using K~c at
the transition, and E and aty from Ref. 119, it can be shown that Pc = 2.9, 12.1
and 9.1 am, respectively, for the unmodified, rubber-modified and hybrid
196 Amar C. Garg, Yiu-Wing Mai

IZ

Eq (~)
%6 o

v.

ore =350 MPo Otc= 2 0 0 M P O


~m*'a c :O,71~rn c=lOum
i A
10 2'0 30 0 I0 2AO 3O
¢6 (~.'~-~-) -

(a) (b)

l q(~)

% =t+25 MPo
c :2l/* ~m

I
5 110

(c)
Fig. 5. Variation of Kic/Kic~ ratio with x/P for (a) unmodified epoxies, (b) rubber-modified
epoxies, and (c) hybrid glass particle-toughened epoxies. (After Kinloch and co-
workers. 84.~o)

particle-toughened epoxies. Any crack tip radius bigger than Pc gives types
A and B crack growth characteristics.
While the crack tip blunting model offers a simple mechanism for, and an
elegant analysis of, crack instability in unmodified and toughened epoxies it
must be mentioned that Cherry and Thomson 12 ~ have observed slow stable
crack growth prior to unstable fracture in a certain epoxy system and this
Failure mechanisms in toughened epoxy resins--a review 197

&
0%

0o

-%
y.

HiGH TEST TEMPERATURES LOW TEST TEMPERATURES


/SLOW RATES /FAST RATES

1 1

60 120 180
TRUE TENSILE YIELD STRESS, aty(MPa)
(a)

3 I

HII3H TEST LOW TEST


TEMPERATURES TEMPERATURES

2 x
@v
A

@a
D

o
o

0 7'5 150
O~y (MPo)

(b)
Fig. 6. Relationship between KI~/KI~ ratio and true tensile yield stress, a,y, showing the
various types of crack growth. (a) Unmodified and rubber-modified epoxies and (b) hybrid
particle-toughened epoxies. (After Kinloch and co-workers. 8'~'119)
198 Amar C. Garg, Yiu-Wing Mai

result is in direct conflict with the crack tip blunting theory. It is also
pertinent to ask whether the blunting model in which dG~c/dA or dK~¢/dA is
negative is any different to the negative dG~¢/dA proposal. Figure 6 shows
that in the instability region of type B crack growth dK~c/da,y is negative and
in the stable type C crack growth region dK~¢/da,y ~0. It is c o m m o n
knowledge that in polymeric materials a,y increases with strain rate (g). When
crack tip blunting occurs the crack tip is stationary and the strain rate in the
crack tip zone (gs) is of the order e.rk/K,96 where e.y is the yield strain. When
crack propagation eventually occurs the strain rate for the moving crack is
~ m ( = anE
" 2 ~y/K~c
3 2 ) and is proportional to the crack tip velocity 6. Thus the aty

axis in Fig. 6 may be thought of as the 6 axis. Since e;m > ~ there is an effective
increase in aty due to d when the blunt crack extends. At the same time, K~
decreases from its blunt tip value to Kt~ for a sharp crack so that dKtJdA (or
dK~Jdaty ) and dKtJId,4 are both negative and unstable cracking ensues by
violation of eqn (1). It appears, therefore, that the crack tip blunting theory
and the negative dGJd,4 model are qualitatively equivalent. Type C crack
growth is stable because dKtJdA m0; however, the stable type A crack
growth cannot be explained in terms of eqn (1) for it only applies to linear
elastic solids. Crack stability analysis in the presence of large plastic flow
remote from the crack tip has been given by Atkins and Mai, 96 Turner ~22
and Paris et alJ 23 The remote plastic flow work accompanying crack growth
acts as a sink for excess energy absorption and this has the effect of
stabilizing otherwise unstable fractures. As mentioned before, it must be
questioned whether valid Ktc measurements can be made in this region.

4.2 Time-temperature superposition of fracture energy, Gtc

It is shown in the previous section that the transitions between the three
different types of crack growth can be controlled by varying either the test
temperature or the applied crosshead/strain rate. This result is just a
reflection of the viscoelastic nature of epoxies. As such the time-temperature
superposition principle in viscoelasticity is expected to apply to the specific
work of fracture measurements, G~c. Hunston and co-workers have
vigorously pursued this line of approach. 9°'1~9'~ 24. By plotting the fracture
energy, G~¢, against time-to-failure, tf (taken as the time from initial load
application to onset of crack propagation), at various temperatures the
curves can be shifted along the tf axis to obtain the best overlap in order to
give a master curve for G~c. It is shown that this shift factor, a v, can be
described by the Arrhenius equation
Failure mechanisms in toughened epoxy resins--a review 199

Zt~
Stress retoxohon Dynamic mechanica[
~ Ophr 0 Ophr
20 " ~ 41, 5phr (bimodal • 5 phr
distribution) ~1 5 phr (bimodo[
• 17 5phr distribution)
16 ~ + 175phr
Yietd

12 15phr

o 13
2 t+

0 ÷

(T =Tg(lOO*g)}
-L,
-1/+0 -100 -60 -20 20
-(Tg-T)
Fig. 7. Shift factor, a x, as a function of temperature needed for time-temperature
superposition of G~c for rubber-modified epoxies. Rubber contents are indicated after each
test method. (After Hunston et a/ri °)

where R is the universal gas constant, T o is the reference temperature at


which the master curve is to be determined, and AE is the activation energy
of the particular failure process involved. Figure 7 shows aT for various
temperatures using To = T~, which is the glass transition temperature of the
epoxies. Though obtained from different test techniques aT is independent of
the compositions of the epoxies. Using the shift factors and To = T~, the
master curves for G,c of an unmodified epoxy and a rubber-modified epoxy
are shown in Fig. 8. These master G~c curves can be fitted to the following
equation:

Glc = Gic~ + fit?' exp R 70 (7)

Hunston and co-workers 119'12+ have suggested that G~c+ measures the
limiting toughness at low temperatures and high stress rates, fl gives the
magnitude of the toughening and depends on the rubber content, 124 and m
assesses the time-dependence of the fracture energy and appears to be a
function o f the rubber particle size distribution. 124 Equation (7), therefore, is
not only useful for the prediction of Gnc given any tf and T, but it can also be
used to compare the fracture properties of epoxies of different composi-
tions.t19 However, much further work is required to establish the usefulness
of the parameters G,c~, fl and m for the characterization of the structure-
property relationship.
200 Amar C, Garg, Yiu-Wing Mai

,&
SLow rote tests: T
• -600C "
x -50 Rubber modified epoxy J
A -~.0 (15 phr odded rubber Vf=018)
o -JO
v -20
. 0
6 • 20
v 30
• 50 . /
i:pQct tests: /

3 To:Tg(I00°C) /
L-~

_ . x ~ O Unmod,fledepoxy /
, . :,~-#~ . .* 4. v. ? . r ~ ~ ,
-20 -15 -I0 -5

tog (tf/nT)(S)
Fig. 8. Fracture energy, G]c,versus reduced time-to-failure,tt/ax. (After Hunston et al.9°)

4.3 Failure modes in toughened epoxies

The fracture behavior of toughened (rubber/inorganic fillers) polymers may


involve several mechanisms, each one contributing towards the total
fracture toughness of the material. Schematically, such possible mechanisms
in a composite are depicted in Figs 9 and 10. The possible modes are (Fig. 9):
shear band formation near rubber particles (1); fracture of rubber particles
after cavitation (2); stretching (3); debonding (4) and tearing (5) of rubber
particles; transparticle fracture (6); debonding of hard particles (7); crack
deflection by hard particles (8); cavitated or voided rubber particles (9);
crazing (10); plastic zone at craze tip (11); diffuse shear yielding (12); shear
band/craze interaction (13); and pinning of crack front (Fig. 10).
Several such failure modes may occur simultaneously in a toughened
polymer, depending upon the type of particles and the matrix. Each such
mechanism contributes to the absorption of energy and some of the
relationships needed for the estimation of these energies are given below.

(1) Stretching and tearing or debonding o f a rubber particle


A simple model to estimate fracture energy involved in such a mechanism
has been considered by Kunz-Douglass et al. 5~ The increase in toughness,
AG.c, contributed by the elastic energy stored in rubber during stretching
Failure mechanisms in toughened epoxy resins--a review 201

0 0 o

0 0
0
0

~ m m

Fig. 9. Crack toughening mechanisms in rubber-filled modified polymers: (1) shear band
formation near rubber particles; (2) fracture of rubber particles after cavitation; (3) stretching,
(4) debonding and (5) tearing of rubber particles; (6) transparticle fracture; (7) debonding of
hard particles; (8) crack deflection by hard particles; (9) voided/cavi tated rubber particles; (10)
crazing; (11) plastic zone at craze tip; (12) diffuse shear yielding; (13) shear band/craze
interaction.

Fig. 10. Crack pinning mechanism. The bowed crack front is at the verge of breaking away
from pinning. (After Lange. 7°)
202 Amar C. Garg, Yiu-Wing Mai

which is dissipated irreversibly (e.g. as heat) when the particle fails either by
debonding from the matrix or by tearing is given by 57
AG,¢ = 47 Vp[1 - 6/(22 + 2 + 4)3 (8)
where 2 is either the extension ratio at the time of debonding or rubber
tearing, Y is either the energy per unit area of interface required to d e b o n d
rubber from the matrix or the rubber tearing energy, and Vp is the volume
fraction of rubber particles.

(2) Brittle fracture of a particle


Assuming the particle to be elastic the specific fracture energy, 7p, of the
particle of mean radius, [, is given by ~25

2[22
7p = X ~/ ar (9)

where t/is the stress concentration factor at the particle equator and ar is the
fracture stress. The n u m b e r of particles, N, of mean radius, [, per unit area of
surface is 126
N = 3(Vp/n[2) (10)
Thus the increase in toughness due to the fracture of particles is
AG,~ = 3t/z[EVpe, 2 (11)
where q (= af/E) is the fracture strain of the particle. For a uniaxial stress
field t/= 2; hence
AGI~ = 12lEe 2 Vp (12)

(3) Crack deflection


The crack is diverted by the particle, resulting in an increase of crack surface
area. This causes an increase in fracture energy given by
AG,c = 3YmVp/2 (13)
where 7m is the specific fracture energy of the matrix. This relationship is
derived on the assumption that the increase in fracture surface area created
by the deviation of the crack is equivalent to half the surface area of the
particle (i.e. 2rtP 2) minus the matrix area, roy2, if the particle was not there.
Also, the n u m b e r of particles is given by eqn (10). Since the actual deviation
of crack may be less, eqn (13) may overestimate the fracture toughness
increase.
A fracture mechanics theory has recently been presented by Faber and
Evans127,~ 28 on the toughness increase due to the tilting and twisting of the
crack front as it meets these particles. Slender rods or fibers with high aspect
Failure mechanisms in toughened epoxy resins--a review 203

ratios are more effective than disc-shaped particles or spheres in deflecting


and twisting the propagating crack and hence in toughening.

(4) Crack-pinning
The obstacles (hard filler particles) create the obstruction to the propagation
of the crack front and cause an increase in toughness by bowing out the
crack front between the particles (Fig. 10). Lange 7°'v4 has given a relation for
the increase in fracture energy due to pinning as

AGx¢ = T/2b (14)


where T is the line energy of the crack front and 2b is the interparticle
spacing. The interparticle spacing can be obtained from 7°

b = d(1 - V,)/3V, (15)


where d = 2~.
For a penny-shaped crack, Lange ~29 showed that the line energy is
27
T = ~- 7m (16)
Thus
AG,¢ = f7m/3b (17)
This equation predicts a linear relationship between the increase in fracture
energy and the ratio i/b. However, this is not observed in practice except for
the case of a glass-filled alumina, la°
The analysis was consequently modified by Evans, s~ who calculated the
increase in strain energy required to bow out the crack front. The increase in
toughness is the derivative of this strain energy increase with respect to the
interparticle spacing. Moloney et al. v7 have reported good agreement of this
modified theory with experimental results for alumina- and silica-filled
epoxies at low volume fractions. At high volume fractions the agreement is
not as good, probably because the 'bow' crack front is semi-elliptical rather
than semi-circular as assumed in the theory. 77.1a~

(5) Yielding near crack tip


The development of crazes or cavitation and/or voiding, shear deformation
and plastic zone near the crack tip can all be considered to cause yielding and
flow of the localized material. This has a blunting effect on the crack tip and
increases the fracture toughness of the material. Several models1 s,16,1~s,
32.133 to describe the effect of such yielding processes on toughness have
been given and are based upon either a critical plastic zone size, ryc, or a
critical opening displacement at the crack tip, 6c, being satisfied. By
204 Amar C. Garg, Yiu-Wing Mai

assuming a Dugdale plastic line zone model it can be shown that the two
parameters are related by
~¢ = 8~yrye/~E (18)
if Poisson constraint effects are neglected. Since the fracture energy, G~c, is
given by the p r o d u c t of ay anf 6 c, and blunting causes 6¢ to increase, the
fracture resistance is hence enhanced. In the crack tip blunting theory l~s
referred to in the previous section 3c is taken as equal to the crack tip radius,
p, at onset of crack growth. Equation (4) gives the toughness enhancement,
K~¢, as a function of p with two adjustable parameters, o't¢ and c. Figure 5
shows the success of this model in predicting K~¢ values for several
unmodified and toughened epoxies.
A similar fracture criterion for blunt notches in epoxies has been provided
by Narisawa et al., ~33 who state that the craze and fracture are initiated at
the elastic-plastic boundary, r = dy (ahead of crack tip), where high stresses
exist owing to stress triaxiality in plane strain conditions. The fracture
initiates when the dilatational stress at this point, dy, becomes critical.
The hydrostatic stress, S, at a point, r, ahead of a crack tip of radius p can
be obtained from slip-line theory and Tresca's yield criterion. Thus it can be
shown for a rigid, perfectly-plastic material with a Poisson ratio, v, of 0.5
that 133
S = ~y{1 + 2In(1 + r/p)} (19)
where Zy is the shear yield stress.
Fracture occurs when at r -- dy, S attains a critical value, S¢. Narisawa et
al. have determined S~ and dy for several epoxies at various loading rates.
This criterion can be used to predict the crack growth behavior in a m a n n e r
similar to the Kinloch and Williams l~s analysis. In terms of the principal
stresses, a~, o 2 and ~3, ahead of the crack tip, it can be shown that for plane
strain
S = ~((~l "{- ~2 -~- ~3) = 2(1 + v)aax~/3[2zcdy(1 + p/2dy)] '/2 (20)
At the onset of fracture S = S¢, so that eqn (20) becomes

K~¢s - Scx/Z~dy = 3(1 + p/Zdy)~/2/2(1 + v) (21)

Equation (21) is similar to eqn (4) in that S~ and dy now become the two
adjustable fracture parameters. But, unlike ate and c, both S~ and dy are
strong functions of loading rate)33 In this respect eqn (21) is not as useful a
fracture criterion as eqn (4), which applies for a wide range of loading rates
and test temperatures.
Garg and Mai ~34 have also derived the following failure equation for
Failure mechanisms in toughened epoxy resins--a review 205

epoxies based on the blunt crack tip model by considering both the complete
stress state at the point of fracture initiation and the pressure dependence on
fracture, i.e.
K,¢/K,~ = [x//-~-+ B - A ] (22)
where
A= (k - 1)(1 + v)(1 + q)5/2 (23a)
(1 + 2q) 2 - 2q - 4v(1 - v)(1 + q)2
and
k(1 + q)3
B = (1 + 2q) 2 - 2q - 4v(1 - v)(1 + q)Z (23b)
Here k = tree/tyro (ratio of compressive to tensile strength) and q = p/2c.
Equation (22) is given for plane strain, but for plane stress v = 0 in eqn (23).
Comparisons of the predicted K~c results for two epoxy resins using eqn (22)
and eqn (4) obtained by Kinloch and Williams 118 are given in Ref. 134.

5 TOUGHENING MECHANISMS

In Section 4.3 the possible failure mechanisms in toughened polymers are


discussed. However, a toughened polymer may involve only a few such
mechanisms depending upon the matrix and the toughening agent. The
plausible toughening mechanisms in individual toughened epoxies are given
below.

5.1 Rubber-toughened epoxies


Several theories have been proposed to explain the toughening mechanisms
in rubber-modified epoxies. 15-18'39'45'50-53"56-63'69'88's9 Many of these
are based on the mechanisms observed in rubber-toughened thermo-
plastics. 5° In such materials the toughening is explained by deformation
mechanisms involving crazing, shear yielding and interaction of crazes with
shear bands. The rubber particles have been treated as the sites for initiation
of crazes and shear bands. Also, the particles may act as obstructions to the
propagating crazes, thereby controlling their sizes. Thus, as a result of the
presence of rubber particles, smaller crazes are obtained. In fracture
mechanics terms this means the reduction of intrinsic flaw size which leads to
the increase in fracture toughness. However, such theories have met with
m i x e d success 57'62'88'89 when used to explain the toughening mechanisms in
epoxide resins.
McGarry and co-workers 51- 53 were the first to investigate the behavior
of rubber-toughened epoxies. They have explained that the toughening
206 Amar C. Garg, Yiu-Wing Mai

mechanism involved was similar to that of modified thermoplastics, i.e. by


generation of crazes in the vicinity of rubber particles. Such a suggestion was
thought to be supported by micrographic studies of fracture surfaces where
the stress whitened zone near the crack tip was observed. However, because
of the high hydrostatic stress beneath the blunt crack tip, micro-cavitation of
the rubber particles causes the occurrence of stress whitening which is
misinterpreted as crazing. Yee and Pearson 88'89 have recently confirmed
that such stress whitening is indeed caused by the initiation and growth of
voids in the rubber particles. This view is also supported by Kinloch. 119
Crazing as a toughening mechanism in rubber-toughened epoxies has also
been suggested by Bucknall, 5° in addition to shear deformation. It has been
observed that the unmodified epoxy deforms by a shear mechanism, i.e. by
shear banding, but rubber-toughened epoxy deforms by an additional
mechanism. During tensile creep tests of toughened epoxy specimens an
increase in volume was noticed in addition to shear deformation. It was
therefore proposed that the volume change could occur only by a massive
crazing of the specimen. In support of these findings, a micrograph of an
ultra-thin section of toughened epoxy resin being stretched on the stage of
an electron microscope has been shown. 5° The micrograph shows a fibrillar
structure which is interpreted as a craze. Once again, Yee and Pearson sS"s9
have expressed their doubts about such an interpretation for crazing, since
volume dilatation can also be caused by voiding of the rubber particles.
The formation of crazes around the rubber particles in thermoplastic
resins has been positively established as the major toughening mechanism,
but in thermoset resins, such as epoxies, it has been a controversial issue.
Doubts have been expressed by many w o r k e r s 57"62's8'89 about their
existence in thermoset resins. None of these authors have found any
evidence of crazes in such resins when fully cured, whether unmodified or
modified. The chances of craze formation in epoxies are considered to be
reduced owing to the very small chain lengths between chemical
entanglements (crosslinks) as the epoxies have high crosslink densities. 135
However, others 37'3s'136A3v have felt strongly about their formation in
thermoset resins. Morgan and co-workers 37"3s have extensively studied
micrographic features of various epoxies in different conditions. On
fractured unmodified epoxy specimen surfaces they have noticed the fibrillar
and nodular structures and cavities. They have interpreted the formation of
fibrillar and nodular structures as being caused by fracture of craze fibrils,
and the formation of cavities by the void growth and coalescence through
the center of the craze. Kinloch 119 suggests that such features could have
been microcracks and the nodular structure of the fracture surface of
unmodified epoxies might have been caused by the non-uniform structure
created by the curing agent. 62
Kunz-Douglass et al. 57 have proposed a different toughening mechanism,
Failure mechanisms in toughened epoxy resins--a review 207

i.e. rubber stretching and tearing. In their study of toughened epoxies, crack
propagation was followed with an optical microscope. As the crack opened
up, stretching and failure of rubber particles was seen to be the major source
of energy absorption in toughened epoxies. Also they did not notice any
significant shear yielding either in unmodified or modified epoxies and they
therefore ignored the contribution of shear yielding to toughness. Neither
did they obtain any evidence of crazing. Based upon such observations,
Kunz-Douglass et al. have proposed a simple model (eqn (8)) to compute the
fracture energy of rubber-toughened epoxies. However, as commented by
Yee and Pearson, 88'89 this quantitative theory can only account for an
increase in toughness by a factor of one or two, whereas the actual increase
in toughness due to rubber particles is seen to be an order of magnitude
greater (Table 1). This has been acknowledged by Sayre e t a / . 1 3 8 in a later
paper. However, they explained that the discrepancy was due to the rubber
tear energy, 7, used in eqn (8) being considerably underestimated in quasi-
static tests. In the fracture experiments the tearing rate for the rubber
particles would be three to four orders of magnitude larger so that 7 is at
least 10 times bigger. If so, better agreement would be obtained between
theory and experiment. In addition, they pointed out the importance of the
state of cure and mechanical properties of the rubber in toughening.
Kinloch et al. 62 have also observed a few discrepancies between the model
of Kunz-Douglass et al. and their own experimental observations. For
example, the model fails to explain the stress whitening phenomenon, the
high toughness at high temperatures and the transitions between various
types of crack growth. It also predicts the wrong trend for the time-
temperature dependence of G~c and it ignores the very important
contributions to toughness by rubber cavitation and shear yielding in the
matrix. It is difficult to resolve these differences because of the sensitivity of
the mechanical properties to curing hardeners and curing conditions.
However, it must be pointed out that the rubber stretching and tearing
model should not be dismissed completely. Bandyopadhyay and co-
workers 58.59 have clearly identified particles stretching and bridging behind
the advancing crack tip in rubber-toughened epoxies from direct scanning
electron microscopy. Toughening by crack interface bridging behind the
propagating tip has been observed in cementitious materials 139-14~ and in
ceramics. 142-144 This gives rise to the so-called R-curve behavior 11~ in
which the crack growth resistance as measured by either Gc or K c increases
with crack growth, Aa. It is to be expected that an R-curve would be obtained
if rubber particle bridging is present behind the tip of the crack. The
condition for a propagating crack is that at the crack tip the effective stress
intensity factor, Ke, is equal to that of the unmodified matrix, Km. Thus

K s = K~ + K , ( A a ) = K m (24)
208 Amar C. Garg, Yiu-Wing Mai

where K r is the stress intensity factor due to the rubber particles bridging
over a distance, Aa, and its magnitude is negative, K a is the applied stress
intensity factor and at fracture it measured the critical value of the fracture
resistance, K c. Therefore the crack growth resistance of the modified epoxy is
Kc = K m - - Kr(Aa ) (25)
K, depends on the properties of the rubber particles and K c reaches a
maximum value, Ko~, when full bridging is established over a saturated
distance, Aa s, and at this point the rubber particles at the original crack tip
position just begin to tear. From experimental evidence given in Ref. 138 it
seems that Aa~ is of the order of several millimeters and an R-curve should
have been obtained. Unfortunately, no such curves have been published in
the literature for ~rubber-modified epoxies. Kunz-Douglass et al. 5v have
shown that the increased fracture toughness in terms of G is given by
Glc = Gm(l - Vp) + AGIc (26)
Since AGj¢ is given by eqn (8) and corresponds to rubber tear, the G~¢ value
calculated is therefore related to the maximum value, K~, of the R-curve, i.e.
K 2 = EG~ = E[Gm(1 - Vo) + AGI¢] (27)
Scott and Philtips 31 have also studied rubber-toughened epoxies and have
discussed the possibilities of mechanisms such as tearing of rubber particles,
localized crazing and plastic zone effect by plasticization of the resin.
However, no conclusive toughening mechanisms have been identified. It is
only in more extensive recent studies by Bascom and co-workers, 15.18 Yee
and Pearson 88'89 and Kinloch et al. 62"63 have more definite toughening
mechanisms been verified. Bascom et al. have noticed in the fracture surface
of specimens of toughened epoxies a stress whitening zone in the slow crack
propagation region. Such stress whitening was observed neither for
unmodified epoxy nor for the modified epoxy in the region of fast crack
growth or during an impact test. Detailed microscopic examination of the
stress whitening region showed the presence of small closely spaced holes
{larger than the original particle size) in the matrix and yielding (seen as tear
markings in the epoxy) of the matrix resin around the particles. Such large
holes are interpreted as being caused by the dilatational deformation of the
particles and the matrix. Also, the dilatation of rubber particles nucleates
local shear yielding of the epoxy matrix, causing a significant crack tip
deformation. Thus volume dilatation of rubber particles and the
surrounding matrix as well as shear yielding of the epoxy are proposed as
two mechanisms responsible for the increase in fracture toughness of
elastomer-modified epoxies.
Yee and Pearson 88"89 and Kinloch e t a / . 6 2 A 1 9 have also proposed similar
Failure mechanisms in toughened epoxy resins--a review 209

mechanisms as being responsible for the increase in toughness of rubber-


modified epoxies, i.e. the crack growth resistance in the rubber-modified
epoxy arises from the large energy-dissipating deformations occurring in the
vicinity of the crack tip. The deformation processes are: (i) localized
cavitation in the rubber, or at the particle/matrix interface caused by
dilatation near the crack tip; and (ii) plastic shear yielding in the epoxy
matrix. Shear yielding is considered to be the major source of energy
dissipation. Owing to the interactions between the stress field ahead of the
crack tip and the rubbery particles, shear yielding is much more important
for the modified epoxies than for the unmodified epoxies. The shear
deformations near the crack tip produce blunting of the crack tip, leading to
the reduction of local stress concentration which consequently improves the
fracture toughness for crack propagation.
While discussing the deformation processes, Yee and Pearson as'89 have
proposed that shear band formation gives rise to shear deformation but
make no reference to homogeneous or diffuse shear yielding. Similarly,
Kinloch e t aL 62 mention the shear deformation to be plastic shear yielding
without specifying whether it is inhomogeneous (shear banding) or diffuse
shear yielding or a combination of both. However, it is a moot point, but it
seems plausible that the shear deformation process occurring in toughened
or pure epoxies would be of a combined nature, i.e. shear banding and diffuse
shear yielding.
In summary, the deformation process leading to toughening of rubber-
toughened epoxies can be described as follows.
The application of a tensile load to a notched specimen introduces a
triaxial stress field ahead of the crack tip and this creates a zone of high
hydrostatic tension or dilatation. The dilatation is further enhanced by the
curing stresses that are induced in the particle by cooling after cure. This
dilatation causes the cavitation of the rubber and the growth of the resultant
voids.
The other process which may initiate simultaneously is the formation of
shear bands due to strain inhomogeneities near the crack tip because of the
high stress concentrations. Additionally, the rubber particles produce stress
concentrations at their equators and also act as sites for the formation of
shear bands (rubber particles initiate shear deformation by producing local
increases in the octahedral shear stress). The shear bands produced at one
particle may terminate at another and keep the yielding localized. Since
there are a large number of particles producing shear bands, far more
yielding would be induced than would occur in the absence of particles. The
evidence of such a shear banding process has been provided by micrographic
studies of subsurface damages just below the fracture surface by Yee and
Pearson. 88,89
210 Amar C. Garg, Yiu-Wing Mai

The creation of voids by the cavitated rubber particles accelerates the


yield deformation process and also enhances the dilatation. For example, the
voids enhance further the stress concentration at their equators and this
tends to decrease the octahedral shear stress required for yielding. 14s The
voids thus grow and promote the formation of plastic zones between them,
enhancing the formation of shear bands. Such a yielding process produces
crack tip blunting which tends to reduce the local stress concentration and
delays crack propagation. Fracture occurs eventually through the voided
plane because the ligaments between the voids can no longer support the
tensile stress.
While the rubber cavitation and matrix shear yielding toughening
mechanisms described above are supported by experimental evidence no
quantitative expressions for these mechanisms have been given. In addition,
one cannot neglect the rubber bridging model, even though it may only play
a secondary role in toughening epoxies, since bridging has been observed in
some epoxies and it should be included. Recently, Evans et al. 146 have
proposed a toughening model for rubber-modified epoxies by considering
both crack bridging 57 behind the advancing crack tip and rubber cavitation
and shear band formation 62'63'ss's9 in a process zone ahead of the crack tip.
This is the first time a quantitative mathematical model has been produced
and it is based on the theories of transformation and microcrack toughening
of ceramic materials. 147'148 In particular, the model indicates the synergistic
effects of rubber stretching/tearing and plastic dilatation for cavitation and
void growth if these two mechanisms occur simultaneously. Quantitative
comparison with this model cannot be made since it requires careful
experimental evaluation of various microstructural parameters associated
with these different mechanisms. It is also noted that the model predicts the
possibility of an R-curve for rubber-modified epoxies, similar to eqn (25). Aa s
is approximately l'25K~/ay,2 2 which is about five times the process zone
width, ~47 i.e. for a rubber-modified epoxy 1~9 with K~ = 2-7 MPa m 1/2, O'y =
75 MPa, Aas = 2 mm.

5.2 Particle-filled epoxies

Possible toughening mechanisms suggested for particle-filled epoxies


include: 7o (a) an increase in fracture surface area due to the irregular path of
the crack; (b) plastic deformation of the matrix around the particles; and (c)
crack pinning. The first mechanism, however, may only account for a small
increase in toughness but not the substantial increase observed in many of
the particulate epoxy systems. The second mechanism always exists as the
particles may act as stress concentrators and induce the formation of shear
bands, causing localized yielding. This also produces crack tip blunting.
Failure mechanisms in toughened epoxy resins--a review 211

5.0

~,.0
÷

K~c o o

Km
3.0

2.0

1.0 i
1.0 20 3.0
%
Fig. 1 I. Increase in K~c relative to unmodified epoxy, Km, as a function of particle size to
interparticle distance, d/2b. - - . , Theoretical curve; 0 , resin A + alumina; O, resin A + silica;
[~, resin B + alumina; +, resin B + silica. (After Moloney et al. 77'79)

However, it is the third mechanism of crack pinning that is thought to be the


most significant source of toughening.
After studying the epoxide resins filled with particles such as alumina and
silica Moloney e t al. 77'78 concluded that the increase in toughness is the
result of crack pinning. Their experimental results compared well with the
predictions based on Evans' crack pinning model 81 as shown in Fig. 11. In
glass-particle-filled epoxy resins Spanoudakis and Young 75 have observed
that, in addition to pinning, crack tip blunting can also result in toughening
of the epoxy matrix. They showed that if the crack tip blunting effect could
be removed then toughening is due to crack pinning. However, the crack
pinning toughening mechanism cannot be applied to epoxies with weak filler
particles such as dolomite and aluminum hydroxide. 77 It was observed that
beyond a critical filler volume fraction (~20%) the crack passes through
these particles, leading to transparticle fracture.
In hybrid particulate epoxies (i.e. glass fillers in a rubber-modified epoxy
matrix) Kinloch e t al. a4 have shown that at low temperatures crack pinning
is the main toughening mechanism and the experimental results agree well
with theories put forward by Evans 81 and Green e t al. T M Fracture can be
described by the critical crack opening displacement criterion. At higher
temperatures the crack pinning mechanism is overshadowed by the large
plastic flow processes at the crack tip. Fracture is now described by the crack
(a)

(b)
Fig. 12. Fracture surfaces for (a) zirconia--epoxy showing crack pinning 'tails' and (b)
zirconia-rubber-epoxy showing crack initiations and arrests in stress whitened region. (After
Lowfl 5)
Failure mechanisms in toughened epoxy resins--a review 213

tip blunting model, and a two-parameter fracture criterion (i.e. a~c and c) is
required (see Figs 5(c) and 6(b)).
Recent work by Low et al. 95 on zirconia-epoxy and zirconia-rubber-
epoxy systems (see Table 4) has also confirmed that crack pinning is the
major toughening mechanism in zirconia-epoxies and that localized shear
flow in the matrix material at the crack tip and rubber cavitation becomes
the dominant mechanism of toughening in the hybrid particulate epoxies.
Experimental evidence to support these conclusions are given in Fig. 12(a)
and (b). In Fig. 12(a) the tails around the zirconia particles are clear
indications of crack pinning operating in the zirconia-epoxy system. In Fig.
12(b), taken in the stress whitening region for the hybrid particulate epoxy, it
appears that there are many sites of crack arrest and crack initiation. Since
initiation involves the debonding/tearing of rubber particles the toughness is
very much enhanced.
In conclusion it may be said that both crack pinning and crack tip
blunting are two mechanisms responsible for the enhancement of toughness
in particulate composites. It is probably difficult to separate the two effects,
although some approximate rules can be applied. 75'84
Although crack deflection processes by tilting and twisting of the crack
front are known to produce toughening in some ceramic materials 127'128
they have not been observed in particle-filled epoxies except when the filler is
a short fiber. Low et al. 95 have observed very large stress whitening zones in
an epoxy/rubber/short-alumina fiber hybrid composite resulting from a
range of failure mechanisms in addition to crack front tilting and twisting.
These are fiber bridging behind the advancing crack tip, rubber cavitation
and plastic shear deformation of the matrix. The development of this stress

K~
g_
tic

t..

._°2

I 2 3 4 5 6 ? 8 9 10
0omoge Zone Exfension Ao (ram)
Fig. 13. Crack growth resistance curve for an epoxy-rubber-alumina fiber hybrid
composite as given in Table 4.
214 Arnar C. Garg, Yiu-Wing Mai

whitening zone gives rise to stable crack growth and hence a rising crack
resistance curve as shown in Fig. 13.

6 CONCLUSIONS

The effect of toughening agents such as CTBN rubbers and particulate fillers
on epoxide resins has been reviewed. It is noticed that the effect of elastomer
addition may increase the toughness up to about 30 times the fracture energy
of unmodified epoxies. However, the gain in toughness for the toughened
epoxy as matrix in a fiber-reinforced composite or as adhesive is much
smaller than for the bulk epoxy. This reduced toughness is explained by the
dependence of toughness on epoxy film thickness. 31'9°
Several toughening mechanisms are reviewed. Although there exists some
controversy about the origin of the high toughness of elastomer-modified
epoxies, the evidence collected so f a r 62'63'88'89 suggests that the deform-
ation processes near the crack tip caused by (i) localized cavitation in the
rubber, or at the particle/matrix interface, and (ii) plastic shear yielding in the
epoxy matrix account for the high toughness in such materials.
As regards the toughening mechanisms in particle-filled epoxies, the
possibility of various mechanisms is discussed and it is suggested that both
plastic yielding of the matrix and pinning of the crack front are responsible
for increasing the toughness of such materials. The relative contribution of
the mechanisms depends on the test temperature. Various factors such as
size and type of particle, volume fraction, and bonding between particle and
matrix play a role in the operative toughening mechanisms. 75- 8o Effective
toughening also depends on the geometric arrangements of the inclusions,
whether a single inclusion or a cluster of inclusions, and whether they are
uniformly distributed or not. 149

7 RECOMMENDATIONS FOR FUTURE RESEARCH

In order to understand the behavior of toughened epoxies for use as a matrix


in fiber-reinforced composites and as structural adhesives, it is necessary to
determine the fracture mechanisms of such resins not only in bulk form but
also for their expected use. Thus, even if the resin shows excellent properties
in bulk form, it may not be useful in laminated composites. For example,
Bersch 32 has identified 24 resins which show higher strains to failure than
the current epoxies. However, only five of these provide a higher strain to
failure in composite form in the compression impact test, and none of them
is acceptable for use as a laminating resin because of deficiencies in other
Failure mechanisms in toughened epoxy resins--a review 215

properties, especially the elastic modulus. Thus, thorough investigations are


needed if toughened epoxy resins are to be used as laminating resins.
Previous studies have generally considered the fracture energies only for
Mode I fracture and hardly any work is available on shear modes, i.e. Modes
II and III. In fact, a combined mode fracture study validated by test is of the
utmost importance in order to predict the behavior of the most general case.
Most of the studies on toughened epoxies have considered D G E B A (T~ ~
80--100°C) epoxide resins. Such epoxies are not suitable for aviation
applications where a resin with higher Tg (~200°C) is desirable. Much
further work is therefore required for high temperature resins with and
without rubbers and other fillers. In particular, the usefulness of the
toughening agents has to be carefully assessed at these relatively high
temperatures.
So far, no information is available for the use of particle-filled epoxies as
matrices in fiber-reinforced composites (except for sheet molding com-
pounds having chopped glass fibers and polyesters with fillers such as
calcium carbonate). Since no loss in mechanical properties may occur as a
result of the addition of such fillers, the use of such materials as fiber
composite matrices may be promising.
A structure has to operate under various loading and environmental
conditions during its service. A detailed study of toughened epoxy under
both static and cyclic loading deserves greater attention. Since the resin is
environmentally sensitive the study should cover the useful temperature and
moisture ranges of the material.

ACKNOWLEDGEMENTS

A. C. Garg wishes to thank liT (Bombay) for providing the necessary


facilities and Y.-W. Mai thanks the Department of Defence, Materials
Research Laboratories, for the financial support of this work. The
experimental work on rubber, zirconia, zirconia/rubber and alumina fiber/
rubber toughened epoxies was performed by L. H. Lahiff and I. M. Low
(Table 4 and Figs 1,2 and 13). Dr S. Bandyopadhyay kindly provided Fig. 12.

REFERENCES

1. K. L. Reifsnider, E. G. Henneke, W. W. Stinchcomb and J. C. Duke, Damage


mechanics and NDE of composite laminates, in Mechanics o f Composite
Materials (eds Z. Hashin and C. T. Herkovich), Proc. of IUTAM, Blacksburg,
VA, Pergamon Press, New York, 1983, pp. 339-90.
216 Amar C. Garg, Yiu-Wing Mai

2. A. C. Garg and O. Ishai, Hygrothermal influence on delamination behaviour


of graphite/epoxy laminates, NASA TM 85935 (1984), and Eng. Fract. Mech.,
22 (1985), pp. 413-27.
3. A. C. Garg, Composites, 17 (1986), pp. 141--9.
4. A. C. Garg, Eng. Fract. Mech., 22 (1985), pp. 1035-48.
5. O. Ishai, A. C. Garg and H. G. Nelson, Composites, 17 (1986), pp. 23-31.
6. S. W. Tsai and H. T. Hahn, Introduction to Composite Materials, Technomic
Publishing Co., Connecticut, USA, 1980.
7. S. C. Chou, O. Orringer and J. H. Rainey, J. Comp. Mater., 10 (1976),
pp. 371 81.
8. H. D. Carden, Impact dynamic research on composite transport structures,
NASA-CP 2321 (1984), pp. 113-59.
9. M. D. Rhodes, Damage tolerance research on composite compression panels,
NASA-CP 2142 (1980), pp. 107-43.
10. D. J. Wilkins, ASTM STP 775 (1980), pp. 168-83.
11. L. C. Jea and D. K. Felbeck, J. Comp. Mater., 14 (1980), pp. 245-59.
12. Y.-W. Mai, B. Cotterell and R. Lord, On fiber composites with intermittent
interlaminar bonding, in Progress in Science and Engineering of Composites
(eds T. Hayashi, K. Kawata and S. Umekawa), Japan Soc. of Composite
Materials, 1982, pp. 271-7.
13. R. Kim, W. Click, J. Hartness, M. Koenig, M. Rich and S. Soni, Improved
materials for composite and adhesive joints, Airforce Wright Aeronautical
Laboratories, AFB, Ohio, AFWAL-TR-82-4182, 1983.
14. N. J. Johnston, Synthesis and toughness properties of resins and composites,
NASA-CP 2321 (1984), pp. 75-95.
15. W. D. Bascom and D. L. Hunston, The fracture of epoxy and elastomer
modified epoxy polymers, in Adhesion 6 (ed. K. W. Allen), Applied Science
Publishers, London, 1980.
16. W. D. Bascom, R. L. Cottington, R. L. Jones and P. Peyser, J. Appl. Polym. Sci.,
19 (1975), pp. 2545 -62.
17. W. D. Bascom, R. Y. Ting, R. J. Moulton, C. K. Riew and A. R. Siebert, J.
Mater. Sci., 16 (1981), pp. 2657 64.
18. D. L. Hunston and W. D. Bascom, Failure behaviour of rubber toughened
epoxies in bulk, adhesive and composite geometries, in ACS Advances in
Chemistt 3' Series No. 208: Rubber Mod(fied Thermoset Resins (eds C. K. Riew
and J. K. Gillham), Amer. Chem. Soc., 1984, pp. 83 99.
19. Y.-W. Mai and A. G. Atkins, J. Mater. Sci., 10 (1975), pp. 2000-3.
20. D.J. Wilkins, A comparison of the delamination and environmental resistance
of a graphite/epoxy and a graphite/bismaleimide, NAV-CD 0037, Naval Air
Systems Command, Washington, DC (1981).
21. T. K. O"Brien, Characterization of delamination onset and growth in a
composite laminate, ASTM STP 775 (1982), pp. 140 67.
22. J. Koutsky, B. Riber, R. Ebewele, M. Luce and K. S. Han, Proc. Organic
Coatings, Appl. Poh'm. Sci., Am. Chem. Soc., 48 (1983), pp. 822 5.
23. P. E. Keary, L. B. Ilcewicz, C. Shaor and T. Trostle, J. Comp. Mater., 19 (1985),
pp. 154 77.
24. J. M. Whitney, C. E. Browning and W. Hoogsteden, J. Reinlbrced Plastics and
Composites, I (1982), pp. 297 313.
25. T. K. O'Brien, lnterlaminar fracture of composites, NASA TM 85768 (1984).
26. D. L. Hunston, Composite Technology Review, 6 (1984), pp. 176-80.
Failure mechanisms in toughened epoxy resins--a review 217

27. R. A. Gledhill, A. J. Kinloch, S. Yamini and R. J. Young, Polymer, 19 (1978),


pp. 574-82.
28. R. A. Gledhill and A. J. Kinloch, Polymer, 17 (1976), pp. 727-31.
29. S. Mostovoy and E. J. Ripling, J. Appl. Polym. Sci., 15 (1971), pp. 644-59.
30. Y.-W. Mai, J. Adhesion, 7 (1975), pp. 141-53.
31. J. M. Scott and D. C. Phillips, 3. Mater. Sci., 10 (1975), pp. 551-62.
32. C. F. Bersch, What we have done, in Proc. Critical Review: Techniques for
Characterization of Composites Materials, Army Materials and Mechanics
Research Center, Mass., AMMRC MS 82-3 (1982), pp.487-91.
33. J. F. Schie and R. J. Juergens, Aerospace America, Sept. (1983), pp. 44-9.
34. S. Bandyopadhyay, Fracture mechanisms in structural epoxies, paper
presented at 6th Polymer Techn. Convention, the Australasian Section of the
Plastics and Rubber Institute, Canberra, 5-7 October 1983.
35. A. J. Kinloch, J. Mater. Sci., 17 (1982), pp. 617-51.
36. R. J. Young, Fracture of thermosetting resins, in Developments in Polymer
Fracture (ed. E. H. Andrews), Elsevier Applied Science Publishers, London,
1979, pp. 183-222.
37. R. J. Morgan, E. T. Mones and W. J. Steele, Polymer, 23 (1982), pp. 295-305.
38. R. J. Morgan and J. E. O'Neal, Polym. Plast. Technol. Engng, 10 (1978),
pp. 49-116.
39. A. J. Kinloch, Metals Science, 14 (1982), pp. 305-18.
40. J. G. Williams, Metals Science, 14 (1982), pp. 344-50.
41. G. A. Crosbie and M. G. Phillips, J. Mater Sci., 20 (1985), pp. 182-92.
42. G. A. Crosbie and M. G. Phillips, J. Mater. Sci., 20 (1985), pp. 563-77.
43. E. H. Rowe, Proc. 34th Annual Techn. Conf., Reinforced Plastics/Composites
Institute, Society of the Plastics Industry, Inc., New York, 1979, p. 233.
44. S. J. Shaw and A. J. Kinloch, Int. J. Adhesion & Adhesives, 5 (1985), pp. 123-7.
45. A.J. Kinloch, S. J. Shaw and D. A. Tod, Rubber-toughened polyimides, in A C S
Advances in Chemistry Series No. 208: Rubber Modified Thermoset Resins (eds
C. K. Riew and J. K. Gillham), Amer. Chem. Soc., 1984, pp. 101-15.
46. L. S. Penn and T. T. Chiao, Epoxy resins, Lawrence Livermore Laboratory,
UCRL-79815, 1977.
47. R. J. Morgan, Structure-property relations of epoxies used as composite
matrices, in Advances in Polym. Sci., Vol. 72: Epoxy Resins and Composites I
(ed. K. Dflsek), Springer-Verlag, Berlin, 1985, pp. 143.
48. R. J. Morgan, Structure-property relationships and the environmental
sensitivity of epoxies, in Developments in Reinforced Plastics--1 (ed. G.
Pritchard), Elsevier Applied Science Publishers, London, 1980, pp. 211-29.
49. C. A. May, Resins for aerospace, Amer. Chem. Soc. Symposium Series 285,
1985, pp. 557 80.
50. C. B. Bucknall, Toughened Plastics, Elsevier Applied Science Publishers,
London, 1977.
51. J. N. Sultan and E J. McGarry, Polym. Engng & Sci., 13 (1973), p. 29.
52. F. J. McGarry, Proc. Roy. Soc. London, A319 (1970), p. 59.
53. J. N. Sultan, R. C. Laible and F. J. McGarry, J. Appl. Polym. Sci., 6 (1971),
p. 627.
54. D. N. Shah, G. Attalla, J. A. Manson, G. M. Conelly and R. W. Hertzberg,
Effect of monotonic and cyclic loading on some rubber modified epoxies, in
A CS Advances in Chemistry Series No. 208: Rubber Modified Thermoset Resins
(eds C. K. Riew and J. K. Gillham), Amer. Chem. Soc., 1984, pp. 117-35.
218 Amar C. Garg, Yiu-Wing Mai

55. C.K. Riew, E. G. Rowe and A. R. Siebert, in ACSAdvances in Chemistry Series


No. 154: Toughness and Brittleness of Plastics (eds D. Deanin and A. M.
Crugnola), Amer. Chem. Soc., 1976, p. 326.
56. W. D. Bascom and D. L. Hunston, Adhesive fracture behaviour of CTBN-
modified epoxy polymers, paper presented at Int. Conf. on Toughening of
Plastics, 4-6 July 1978, London, Paper 22.
57. S. Kunz-Douglass, P. W. R. Beaumont and M. E Ashby, J. Mater. Sci., 15
(1980), pp. 1109-23.
58. S. Bandyopadhyay and V. M. Silva, Crack propagation studies of a rubber-
toughened epoxy resin in the SEM, in Proc. 6th Int. Conf. on Fracture, New
Delhi (eds S. R. Valluri et al.), Pergamon Press, Oxford, 1984, pp.2971-8.
59. S. Bandyopadhyay, P, J. Pearce and S. A. Mestan, Crack tip micromechanisms
and fracture properties of rubber toughened epoxy resins, paper presented at
Churchill Conference on Deformation, Yield and Fracture of Polymers, 1985,
Cambridge, UK, Paper 18.
60. S. Bandyopadhyay, J. Mater. Sci. Lett., 3 (1984), p. 39.
61. R. S. Drake and A. R. Siebert, Quart. SAMPE, 6 (1975), p. 11.
62. A. J. Kinloch, S. J. Shaw, D. A. Tod and D. L. Hunston, Polymer, 24 (1983),
pp. 1341-54.
63. A. J. Kinloch, S. J. Shaw, D. A. Tod and D. L. Hunston, Polymer, 24 (1983),
pp. 1355-63.
64. E. H. Rowe, A. R. Siebert and R. S. Drake, Modern Plastics, 49 (1970), p. 110.
65. E. M. Yorkgitis, C. Tran, N. S. Eiss, T. Y. Hu, I. Yilgor, G. L. Wilkes and J. E.
McGrath, Siloxane modifiers for epoxy resins, in A C S Advances in Chemistry
Series No. 208: Rubber Modified Thermoset Resins (eds C. K. Riew and J, K.
Gillham), Amer. Chem. Soc., 1984, pp. 137-62.
66. L. C. Chan, J. K. Gillham, A. J. Kinloch and S. J. Shaw, Rubber-modified
epoxies: cure, transitions and morphology, ibid., pp. 235 60.
67. Idem. Rubber-modified epoxies: morphology, transitions and mechanical
properties, ibid., pp. 261--80.
68. J. Mijovic, E. M. Pearce and C.-C. Foun, Fluoro-elastomer-modified
thermoset resins, ibid., pp. 293 310.
69. S. C. Kunz, J. A. Sayre and R. A. Assink, Polymer, 28 (1982), pp. 1897-1906.
70. F. F. Lange, Fracture of brittle matrix particulate composites, in Composite
Materials, Vol. 5: Fracture and Fatigue (ed. L. J. Broutman), Academic Press,
New York, 1974, pp. 2-44.
71. R. J. Young and P. W. R. Beaumont, J. Mater. Sci., 10 (1975), pp. 1343-50.
72. K. P. Gadkaree and G. Salee, PoIym. Composites, 4 (1983), pp. 19-25.
73. R. Griffiths and D. G. Holloway, J. Mater. Sci., 5 (1970), pp. 302 7.
74. F. F. Lange and K. C. Radford, J. Mater. Sci., 6 (1971), pp. 1197-203.
75. J. Spanoudakis and R. J. Young, J. Mater. Sci., 19 (1984), pp. 473-86.
76. htem. ibid., 19 (1984), pp. 487 96.
77. A. C. Moloney, H. H. Kausch and H. R. Stieger, J. Mater. Sci., 18 (1983),
pp. 208-16.
78. htem. ibid., 19 (1984), pp. 1125-30.
79. Idem. Interfacial properties of filled epoxy resins, in Adhesive Joints (ed. K. L.
Mittal), Plenum Publishing Co., New York, 1984, pp. 883-904.
80. A. C. Moloney and H. H. Kausch, J. Mater. Sci. Lett., 4 (1985), pp. 289-92.
81. A. G. Evans, Phil. Mag., 26 (1972), pp. 1327-44.
Failure mechanisms in toughened epoxy resins--a review 219

82. D. Maxwell, R. J. Young and A. J. Kinloch, J. Mater. Sci. Lett., 3 (1984),


pp. 9-12.
83. A. J. Kinloch, D. Maxwell and R. J. Young, ibid., 4 (1985), pp. 1276-9.
84. ldem. J. Mater. Sci., 20 (1985), pp.4169-84.
85. R. J. Young, D. L. Maxwell and A. J. Kinloch, ibid., 21 (1986), pp. 380-8.
86. M. E Tse, J. Appl. Polym. Sci., 30 (1985), pp. 3625-47.
87. H. M. Lahiff, Rubber-toughened epoxy resins, BE Thesis, Department of
Mechanical Engineering, University of Sydney, 1986.
88. A. F. Yee and R. A. Pearson, J. Mater Sci., 21 (1986), pp. 2462-74. (Also see
NASA-CR 3718, 1983.)
89. R. A. Pearson and A. F. Yee, ibid., 21 (1986), pp. 2475-88.
90. D. L. Hunston, A. J. Kinloch, S. F. Shaw and S. S. Wang, Characterisation of
the fracture behaviour of adhesive joints, in Adhesive Joints (ed. K. L. Mittal),
Plenum Publishing Co., New York, 1984, pp. 789-807.
91. R.W. Hertzberg and J. A. Manson, Fatigue of Engineering Plastics, Academic
Press, New York, 1980.
92. L. J. Broutman and S. Sahu, Mater Sci. & Eng., 8 (1971), p. 98.
93. A. G. Atkins, J. Mater. Sci., 10 (1975), pp. 819-32.
94. Y.-W. Mai and F. Castino, J. Mater. Sci., 19 (1984), pp. 1638-55.
95. I.-M. Low, Y.-W. Mai, S. Bandyopadhyay and V. M. Silva, in Proc. 1987
Australian Fracture Group Symposium (ed. Y.-W. Mai), Sydney University,
Sydney, 1987, pp. 77-89.
96. A. G. Atkins and Y.-W. Mai, Elastic and Plastic Fracture: Metals, Polymers,
Ceramics, Composites, Biological Materials, Ellis Horwood/John Wiley,
Chichester, 1985, Chapter 7.
97. A.J. Kinloch, D. G. Gilbert and S. J. Shaw, J. Mater Sci., 21 (1986), pp. 1051-6.
98. G. I. Taylor, Proc. R. Soc., A201 (1950), p. 192.
99. P. G. Saffman and G. I. Taylor, ibid., A245 (1958), p. 312.
100. B. Cotterell and J. K. Reddell, Int. J. Fract., 13 (1977), pp. 267-77.
101. y.oW. Mai and B. Cotterell, Int. J. Fract., 32 (1986), pp. 105-25.
102. Y.-W. Mai, B. Cotterell, G. Vigna and R. Horlyck, Polym. Eng. & Sci., 27 (1987),
pp. 804-9.
103. Y.-W. Mai and B. Cotterell, J. Mater Sci., 15 (1980), pp. 2296-306.
104. Y.-W. Mai and B. Cotterell, Eng. Fract. Mech., 21 (1986), pp. 123-8.
105. C. Gurney and Y.-W. Mai, Eng. Fract. Mech., 4 (1972), pp. 853-63.
106. Y.-W. Mai and A. G. Atkins, J. Strain Analysis, 15 (1980), pp. 63-74.
107. S. Yamini and R. J. Young, Polymer, 18 (1977), pp. 1074-80.
108. H. T. Corten, Fracture mechanics of composites, in Fracture--An Advanced
Treatise, I1"ol. 7 (ed. H. Liebowitz), Academic Press, New York, 1972, pp. 675-
769.
109. Y.-W. Mai, A. G. Atkins and R. M. Caddell, Int. J. Fract., 11 (1975), pp. 939-53.
110. P.S. Leevers, Theoretical and Applied Fracture Mechanics, 6 (1986), pp. 45-55.
111. Y.-W. Mai and B. R. Lawn, Ann. Rev. Mater. Sci., 16 (1986), pp.415-39.
112. D. Maugis, J. Mater. Sci., 20 (1985), pp. 3041-73.
113. E. H. Andrews and A. Stevenson, J. Mater. Sci., 13 (1978), pp. 1680-8.
114. A. J. Kinloch and R. J. Young, Fracture Behaviour of Polymers, Applied
Science Publishers, London, 1983, Chapter 8.
115. R. A. Gledhill and A. J. Kinloch, Polym. Eng. & Sci., 19 (1979), pp. 82-8.
116. F. A. Johnson and J. C. Radon, Eng. Fract. Mech., 4 (1972), pp. 555-76.
220 Amar C. Garg, Yiu-Wing Mai

117. Y.-W. Mai and N. B. Leete, J. Mater. Sci., 14 (1979), pp. 2264-7,
118. A. J. Kinloch and J. G. Williams, J. Mater. Sci., 15 (1980), pp. 987-96.
119. A.J. Kinloch, Mechanics and mechanisms of fracture of thermosetting epoxy
polymers, in Advances in Polymer Science--Vol. 72 (ed. K. Dfisek), Springer-
Verlag, Berlin, 1986, pp. 45-67.
120. J. G. Williams, Fracture Mechanics of Polymers, Ellis Horwood/John Wiley,
Chichester, 1984.
121. B. W. Cherry and K. W. Thomson, J. Mater. Sci., 16 (1981), pp. 1913-24.
122. C. E. Turner, in ASTM STP 677 (1979), pp. 614-28.
123. P. C. Paris, H. Tada, A. Zahoor and H. Ernst, in ASTM STP 668 (1979),
pp. 5-36.
124. D. L. Hunston and G. W. Buliman, 'Viscoelastic fracture behaviour for
different rubber-modified epoxy adhesive formulations', to be published.
125. J. Gurland, Fracture of metal matrix particulate composites, in Composite
Materials, Vol. 5: Fracture and Fatigue (ed. L. J. Broutman), Academic Press,
New York, 1975, pp. 45-91.
126. E. E. Underwood, Particle size distribution in Quantitative Microscopy (eds
R. T. DeHoff and F. N. Rhines), McGraw-Hill, New York, 1968, pp. 149-200.
127. K. T. Faber and A. G. Evans, Acta Metall., 31 (1983), pp. 565-76.
128. Idem. ibid., 31 (1983), pp. 577-84.
129. F. F. Lange, Phil. Mag., 22 (1970), pp.983-92.
130. F. F. Lange, J. Amer. Ceram. Soc., 54 (1971), pp. 614-20.
131. D. J. Green, P. S. Nicholson and J. D. Embery, J. Mater. Sci., 14 (1979),
pp. 1657-61.
132. S. Hashemi and J. G. Williams, J. Mater. Sci., 20 (1985), pp. 922-8.
133. I. Narisawa, T. Muralyama and H. Ogawa, Polymer, 23 (1982), pp. 291-4.
134. A. C. Garg and Y.-W. Mai, Failure prediction in toughened epoxy resins,
Composites Sci. & Techn., 31 (1988), pp. 225-42.
135. A. M. Donald and E. J. Kramer, J. Mater. Sci., 17 (1982), pp. 1871-9.
136. J. Lilley and D. G. Holloway, Phil. Mag., 28 (1973), pp. 215-20.
137. S. H. Carr, What we should be doing, in Proc. of the CriticalReview, Techniques
for the Characterization of Composite Materials, Army Materials and
Mechanics Research Center, Mass., AMMRC, MS 82-3 (1983), pp. 493-9.
138. J.A. Sayre, S. C. Kunz and R. A. Assink, Effect of rubber cross-link density and
tear energy on the toughness of rubber-modified epoxies, in ACS Advances in
Chemisto, Series No. 208: Rubber Modified Thermoset Resins (eds C. K. Riew
and J. K. Gillham), Amer. Chem. Soc., 1984, pp. 215-34.
139. A. Hillerborg, M. Modeer and P. E. Petersson, Cement and Concrete Research,
6 (1976), pp. 773-82.
140. M. Wecharatana and S. P. Shah, Cement and Concrete Research, |0 (1980),
pp. 833-44.
141. B. Cotterell and Y.-W. Mai, J. Mater. Sci., 22 (1987), pp. 2734-8.
142. Y.-W. Mai and B. R. Lawn, J. Amer. Ceram. Soc., 70 (1987), pp. 289-94.
143. P. L. Swanson, C. J. Fairbanks, B. R. Lawn, Y.-W. Mai and B. J. Hockey, J.
Amer. Ceram. Soc., 70 (1987), pp. 279-89.
144. R. F. Cook, C. J. Fairbanks, B. R. Lawn and Y.-W. M ai, J. Mater. Res., 2 (1987),
pp. 345-56.
145. A. L. Gurson, J. Eng. Mater. Techn., Trans. ASME, 99 (1977), pp. 2 15.
146. A.G. Evans, Z. B. Ahmad, D. G. Gilbert and P. W. R. Beaumont, Acta Metall.,
34 (1986), pp. 79 87.
Failure mechanisms in toughened epoxy resins--a review 221

147. R.M. McMeeking and A. G. Evans, J. Amer. Ceram. Soc., 65 (1982), pp. 242-6.
148. A. G. Evans and K. T. Faber, J. Amer. Ceram. Soc., 67 (1984), pp. 255-60.
149. J. C. M. Li and S. C. Sanday, Acta Metall., 34 (1986), pp. 537-43.
150. A. M. James and W. E. Harvill Jr, Advanced Composites, Lockhead Horizons,
1986, pp. 31-43.
151. A.C. Moloney, H. H. Kausch, T. Kaiser and H. R. Beer, J. Mater. Sci., 22 (1987),
pp. 381-93.

A P P E N D I X A: S T R U C T U R E OF SOME EPOXY RESINS A N D


C U R I N G A G E N T S 46

(a) Epoxies
1. Polyglycidyl ethers of the Novolacs (Union Carbide ERR 0100)

/o /o\
O--CHz--CH--CH 2 O--CH2--CH--CH 2 O--CHz--CH--CH 2

2. Triglycidyl p-amino phenol (Union Carbide ERLA 0510)

/ \0
O--CH2--CH--CH 2

/% y\ /o
CH2--CH--CH 2 CH2--CH--CH 2

3. Tetraglycidyl diamino diphenyl methane (TGDDM, Ciba-Geigy


MY 720)

CH2--CH--CH 2
\ N ~ C H 2 ~ N /
/ L_Z_/ L~_/ \
C~HO2 -/- C H - - C H 2 CH2--CH - -~C H 2
,NO
222 Amar C. Garg, Yiu-Wing Mai

4. Diglycidyl ether of bisphenol-A (DGEBA)


CH3
/% r /o\
CHs

(b) Curing agents


1. Dicyandiamide (DICY)
NH2
H2N--C~N--C~N
2. Borontrifluoride-monoethylene amine (BF3MEA)
F
[
F--B--: NH2--CHz--CH 3
I
F
3. Diamino diphenyl sulfone (DDS, Ciba-Geigy EPORAL)
O
H2N ~_Z_/ IL X_Z_/ NH2
O
4. Methyl dianiline (MDA)

H2N ~ ~ - CH2 ~ ~ - NH2

5. Meta phenylene diamine (MPDA)


NH 2

~-NH 2
6. Diethylene triamine (DETA)
H 2N--CH2--CH2--NH--CHi--CH2--NH 2
7. Piperidine
(CH2)sNH
Failure mechanisms in toughened epoxy resins--a review 223

A P P E N D I X B: C H E M I C A L F O R M U L A FOR CARBOXYL-
T E R M I N A T E D B U T A D I E N E - A C R Y L O N I T R I L E (CTBN) RUBBER

HOOC--~(CH 2 - - C H = C H - - C H 2)~--(CH 2 - - C H ) y ] - C O O H
[_ CN J~
where on average x = 5, y = 1 and m = 10, giving a molecular weight of
3320 g m o l - 1.

You might also like