Vous êtes sur la page 1sur 16

The impact resistance of

composite materials---a review


W.J. CANTWELL* andJ. MORTONt
(* Ecole Polytechnique Fdd#rale de Lausanne, Switzerlandl t Virginia Polytechnic
Institute and State University, USA)
In this paper the impact response of continuous fibre-reinforced composites
is reviewed. An attempt is made to draw together much of the work published in the literature and to identify the fundamental parameters determining the impact resistance of continuous fibre-reinforced composite materials.
The effect of varying the properties of the fibre, matrix and interphase are
examined as well as the role of target geometry and loading rate on the
dynamic response of these materials.

Key words: composite materials; impact testing; impact resistance; residual


strength; failure mode; fibre properties; matrix properties; interphase;
laminate stacking sequence; geometry; loading rate
Fibre-reinforced composite materials such as carbon,
glass and Kevlar ~ fibre-reinforced plastics are finding
increasing use in a wide range of both low and high
technology engineering applications. Composites offer
a number of distinct advantages over more conventional engineering materials such as aluminium and
steel. These include higher specific strengths and
stiffnesses, superior corrosion resistance as well as
improved fatigue properties. Coupled with these
improvements in general overall performance is the
fact that the cost of manufacturing components from
fibre-reinforced plastic is often less than with more
conventional metals. This is true not only at the lowcost end of the market occupied by sheet moulding
compounds (SMC) but also in the aerospace sector
where complex load-bearing shapes can be produced in
a limited number of steps, saving time in both joining
and assembly.
Composite materials do, however, suffer some serious
limitations. Perhaps the most significant amongst these
is their response to localized impact loading such as
that imparted by a dropped tool or runway debris. In
recent years many research programmes have been
undertaken in an attempt to better understand the
impact response of these materials 1-9. The majority of
this work has been undertaken on continuous fibre,
high performance composites since these materials are
finding increasing use in the design of a large number of
civil and military aircraft, i.e., in circumstances where
the consequences of impact are likely to be most
serious.
The manner in which composite materials respond to
impact loading and dissipate the incident kinetic energy
of the projectile is very different to that of metals. For

low and intermediate incident energies, metals absorb


energy through elastic and plastic deformation 1.
Although the latter may cause some permanent structural deformation, its consequences on the loadcarrying capability of the component are usually
small 11. At high incident impact energies target perforation may occur and the passage of the impactor will
generally result in petalling, cracking and spalling 12.
Although such damage will degrade the load-bearing
ability of the structure, its effects can generally be
predicted using fracture mechanics principles t2. In
composites, however, the ability to undergo plastic
deformation is extremely limited with the result that
energy is frequently absorbed in creating large areas of
fracture with ensuing reductions in both strength and
stiffness 1.13. Furthermore, the prediction of the postimpact load-bearing capability of a damaged composite
structure is more difficult than for metals since the
damage zone is generally complex in nature and
consequently very difficult to characterize 12. The
problem is further complicated by the lack of existing
standards or established testing techniques for impact
of composite materials. Much of the work published in
the literature has been conducted on purpose-built
machines using convenient specimen geometries. As a
result, direct comparisons between different material
systems is often very difficult and immediate conclusions are sometimes hard to draw.
The objective of this article is to draw together the
pertinent findings of many of the articles published in
the field of impact of fibre-reinforced composites and
thereby present a global view of the state-of-the-art.
Initially, the paper will review briefly the techniques
most frequently employed for assessing the impact

0010-4361/91/050347-16(~) 1991 Butterworth-Heinemann Ltd


COMPOSITES. VOLUME 22. NUMBER 5. SEPTEMBER1991

347

response of composite materials. Following this, the


effect of varying fundamental parameters such as fibre
and matrix properties and stacking sequence on the
dynamic response of these materials will be considered.
Since a substantial amount of the work published in the
literature relates to tests on carbon fibre composites,
the main focus of this report will be on this material.
However, work undertaken on other materials will be
used in order to supplement and substantiate ideas and
theories.

IMPACT TEST TECHNIOUESFOR COMPOSITE


MATERIALS
Ideally, the impact test fixture should be designed to
simulate the loading conditions to which a composite
component is subject in operational service and then
reproduce the failure modes and mechanisms likely to
occur. In simple terms, the impact problem can be
divided into two separate conditions: low velocity
impact by a large mass (dropped tool) and high velocity
impact by a small mass (runway debris, small arms fire,
etc.). The former is generally simulated using a falling
weight or a swinging pendulum and the latter using a
gas gun or some other ballistic launcher. However, as
stated previously, due to a lack of experimental standards a wide variety of testing techniques is presently
being employed in order to assess the dynamic
response of reinforced plastics 1-4, 14-17, making direct
comparison difficult.

Weig

Specimen
Notch

Inertiabar

~I/-- ~pactor

Straingauge--I [

Specimen [ __
Notch

Specimen
or
/I
Straingauges

oe'P"tar
b

In this section the more commonly used techniques will


be presented and discussed as well as the problems
associated with the ensuing data analysis.

Low velocity impact


Machines currently used for simulating the low velocity
impact response of composite materials include the
Charpy and Izod pendulums, the falling weight fixtures
such as the Gardner and drop dart tests as well as
hydraulic machines designed to perform both in-plane
and out-of-plane testing at velocities up to 10 m s-1.
1. Charpy pendulum
Many of the early impact studies on composite materials were undertaken using the Charpy test method
originally developed for testing metals 11'1s'19. The
reason for this choice was the fact that the Charpy
pendulum is both simple to use and can be instrumented, and therefore, in principle, can yield information on the processes of energy absorption and dissipation in composites. The test specimen is generally a
thick beam, sometimes incorporating a notch at its mid
point as shown in Fig. l(a). The specimen is supported
in a horizontal plane and impacted by the swinging
pendulum directly opposite the notch. The energy
dissipated during impact is usually recorded by a dial
on the test apparatus. Further information can be
obtained by instrumenting the impactor with a strain
gauge, thereby enabling the determination of the
variation of the impact force with time. The energy
absorbed during impact can also be determined by
integrating the force/time signal.
348

COMPOSITES. SEPTEMBER 1991

Fig. 1 (a) Charpy impact test; (b) Izod impact test; (c) tensile
Hopkinson-bar test; (d) gas gun

The test set-up suffers a number of disadvantages such


as the fact that the load/time curves often contain high
frequency harmonic oscillations resulting from the
natural response of the impactor, etc. These effects can
generally be filtered once the harmonic frequencies of
the various components have been determined 2. As
stated previously, the test specimen is a short, thick
beam and is not therefore typical of engineering
components. Further, the test is destructive, inducing
failure modes that are not necessarily observed under
low velocity impact loading on operational structures.
Bader and Ellis.21 showed that the Charpy energy of
carbon fibre-reinforced plastic (CFRP) varied with
specimen geometry and doubt, therefore, the applicability of the technique. The Charpy test is only suitable
for ranking the impact performance of continuous fibre

composites and as a first step in determining the


dynamic toughness of these materials.

to geometries such as the DCB since the specimen may


not be deforming in the same mode as a similar
statically-loaded specimen24.

2. Izod test
The Izod impact test is shown schematically in Fig.
l(b). The test set-up and procedure are similar to those
outlined above. In the Izod test specimen is clamped in
the vertical plane as a cantilever beam and impacted by
a swinging pendulum at the unsupported end. The test
suffers similar problems to those reported above and
again is best suited as a tool for ranking the impact
resistance of composite materials.

3. Drop-weight impact tests


Here a weight is allowed to fall from a pre-determined
height to strike the test specimen or plate supported in
the horizontal plane. In general, the impact event does
not cause complete destruction of the test specimen hut
rebounds, enabling a residual energy to be determined
if necessary. The incident velocity of the impactor can
be determined from the equations of motion or by
using optical sensors located just above the target.
Frequently, the impactor is instrumented, enabling the
force/time characteristics to be determined, and may
also contain a displacement transducer in order to
permit the determination of energy dissipation during
the impact event. One of the advantages of this test
with respect to the Charpy and Izod tests is that a wider
range of (est geometries can be tested, thereby
enabling more complex components to be tested.
Although testing is generally undertaken using a
hemi-spherical impactor, it is possible to use other
impactor shapes such as blunt cylinders or sharp points.

Variations on the drop-weight theme include the


Gardner test where a hemi-spherical impactor strikes a
small diameter circular plate and the driven dart test
where a hemi-spherical probe is driven into the
specimen at a pre-determined rate 22.

Intermediate and high rate impact


1. Hopkinson-bar technique
The Hopkinson-bar technique is similar to the previous
test procedure in that it permits the determination of
the variation of basic material properties as a function
of strain rate. Several different types of Hopkinson-bar
are currently employed; these include the punchloaded Hopkinson-bar, the compression bar, the
tensile bar and the Hopkinson-bar shear test z~-27. The
set-up and experimental procedures associated with
these tests has been discussed in detail elsewhere25.
Here, a brief description of the tensile test will be
given. The test set-up for undertaking dynamic tensile
tests is shown in Fig. l(c) ~. Here, specimens waisted
through the thickness are bonded into slots in the
inertia bar and input bar as shown in the diagram. This
configuration is then placed within the weighbar. Strain
rates approaching 1000 s-1 can then be achieved using
gas-driven projectiles to accelerate the weighbar and,
in turn, load the input bar 25. The inertia of the inertia
bar then enables the test coupon to be loaded dynamically. Strain gauges bonded to the input and inertia
bars enable the incident and reflected stress waves to
be analysed and permit the determination of a dynamic
stress/strain curve for the material. The Hopkinson-bar
tensile test is destructive.
Care has to be taken in order to ensure that the interface between the specimen and loading bars is good,
otherwise a shear failure within the gripping section is
likely to occur. Further, in order to minimize stress
concentrations associated with the gripping area,
relatively long specimens (approximately 60 mm) are
required when testing composite materials .

2.
4. Hydraulic test machines
In recent years a number of workers have used hydraulic test machines for assessing the deformation and
failure characteristics of materials at high rates of
strain 23'24. Here, test geometries such as tensile dogbone specimens or double cantilever beam (DCB)type
specimens can be tested over a wide range of strain
rates The strain history of the specimen can be
measured using bonded strain gauges or an optical
transducer. If a strain gauge or any other displacement
measuring device is bonded to the specimen, the strain
rate sensitivity of the adhesive should be considered.
The advantage of this technique is that the test specimens permit the evaluation of basic material properties
such as tensile strength, modulus and interlaminar
fracture toughness without the contact effects associated with falling weight impact. Extreme care has to be
taken in order to ensure that the mass of the load cell
and gripping system are as low as possible since inertial
effects resulting from these components may conceal
the true material response24. Caution should also be
exercised when applying fracture mechanics principles

28

Gas gun impact testing

Impact testing at ballistic rates of strain can be achieved


using a high pressure gas gun such as that shown in Fig.
l(d) ~9. Typically, a gas such as nitrogen is fed to a
chamber located at one end of the barrel. Here the gas
is restrained by a plastic diaphragm. When the gas has
reached a pre-determined value the diaphragm is burst
(by electrical heating or a mechanical puncturing
device), accelerating a projectile down the barrel to
strike a specimen or component supported vertically.
The velocity of the impactor can be determined just
prior to impact using optical sensors3 or by using a
simple break-wire technique 29. Generally, the test is
not completely destructive but frequently results in
large-scale damage and/or target perforation. Until
recently the technique suffered the disadvantage that
very little information could be obtained from the test
itself. However, instrumented gas guns have now been
developed, enabling force/displacement histories to be
measured and the impact event to be analysed in more
detail31,32. Gas guns can be used to test large structures
and are therefore useful for assessing the high velocity
impact response of composite materials.

COMPOSITES. SEPTEMBER 1991

349

Conclusions
At present, no acceptable standard testing procedures
are available for impact testing of composite materials.
Consequently, a wide variety of testing procedures,
specimen geometries and data reduction techniques are
presently being employed. Pendulum techniques such
as the Charpy and Izod tests often require specimen
geometries that are not representative of component
dimensions and so are essentially suitable only for
ranking the impact response of composites. Dropweight rigs and gas guns offer more representative
approaches for assessing the impact response of these
materials. Greater use of instrumented impactors has
led to a deeper understanding of the processes of
energy absorption and dissipation in these materials.

fibre composites are given in Table 1. Generally


speaking, failure modes that involve fracture of the
matrix or interphase region result in low fracture
energies whereas failures involving fibre fracture result
in significantly greater energy dissipation.
The following sections will investigate the role of the
material constituents independently and examine the
influence of varying their properties on the overall
impact resistance of the composite.

Fibre

INFLUENCE OF CONSTITUENT PROPERTIES ON


THE IMPACT RESPONSE OF COMPOSITE
MATERIALS

The role of the fibres in a composite structure is


extremely important since they are responsible for
bearing a significant percentage of the applied load. At
present, many types of fibre are available. In aeronautical applications these include carbon, glass, and
KevlaP~ fibres. Within each of these categories fibres
exhibiting a wide range of mechanical properties are
available. Unfortunately, it is often difficult to separate
the effects of mechanical properties (such as strength
and stiffness) from those arising from geometrical
factors (such as fibre shape and diameter) and interfacial properties (such as the strength of the chemical
bond between fibre and matrix).

In sirqple terms, a fibre-reinforced composite is


composed of three constituents: the fibres, the matrix
and a fine interphase region (sometimes referred to as
the interface) responsible for assuring the bond
between the matrix and fibre. The manner in which the
material deforms and fractures depends upon both the
chemical and mechanical properties of these three
constituents. Detailed analyses using both optical and
scanning electron microscopes have identified a
number of failure mechanisms in fibre-reinforced
composite materials29'3336. These include delamination, intralaminar matrix cracking, longitudinal matrix
splitting, fibre/matrix debonding, fibre pull-out and
fibre fracture. The relative energy absorbing capability
of these fracture modes depends upon the basic properties of the constituents as well as the loading mode.
Typical values measured for a number of continuous

Early studies36, in which the relative performance of a


number of continuous fibre composites was examined,
suggested that the Cliarpy impact resistance of S-glass
and Kevlar~ fibre-reinforced composites was over five
times greater than that of a Modmor II carbon fibrereinforced composite. Beaumont et al. is examined the
Charpy load/time traces of a number of materials and
showed that the curve corresponding to an HMS
carbon composite was extremely brittle, failing catastrophically at maximum load. The E-glass and Kevlar~
49 composites failed in a more progressive manner
indicative of energy dissipation through delamination,
splitting and other failure processes. The authors
quantified the Charpy load/time traces by defining a
ductility index (DI), this being the ratio of the energies
associated with the crack propagation phase (the area
after maximum load) and the initiation phase (the area

More recently, hydraulic machines have been used to


examine the response of composite materials at high
rates of strain. Although somewhat more expensive
than other, more conventional techniques, such
machines are ideally suited for analysing the rate
dependency of basic material properties.

Table 1. Typical values of the energy absorbing capability of various continuous fibre composites for
different failure modes
Failure mode

Material

Splitting

Type II CF/epoxy
AS4/PEEK

0.1-1
3.8

Delamination

T300/epoxy
IM6/PEEK

0.1
2.2

Transverse
fibre fractu re

Treated CF/epoxy
Untreated CF/epoxy
AS4/PEEK

20
60
128

72
72
60

Fibre pull-out

CFIpolyester
CF/bismaleimide

26
800

104
105

Debonding

CF/epoxy

106

350 COMPOSITES. SEPTEMBER 1991

Typical fracture energy


(kJ m-2)

Reference

11
60
102
62

up to maximum load). The resulting ductility indices


for the Kevlar-49, E-glass and HMS carbon/epoxy
composites were 23, 0.4 and 0.0 respectively, indicating
clearly the superior energy absorbing capability of the
Kevlar fibre.
Hancox 37 and Bader and Ellis 21 compared the low
velocity impact response of composites containing type
I and II carbon fibres. In both cases it was shown that
the materials containing type II fibres (higher failure
strength) offered superior impact resistance. Similar
results have been observed following low velocity
impact tests on E- and S-glass fibre composites 15.
In an attempt to better understand the fundamental
parameters controlling the processes of energy absorption and dissipation in composite materials, Chamis et
al. 3s undertook Izod impact tests on a wide range of
systems. They concluded that flexure and interlaminar
shear deformations are dominant energy-absorbing
mechanisms in composites and that the area under the
material's linear stress/strain diagram represents a
useful approach for predicting the impact resistance of
a composite. Essentially, composites with large areas
under the stress/strain curve are more effective energy
absorbers. Fig. 2 presents the experimental data
obtained by Chamis and Sinclair39 plotted as a function
of the energy absorbing capability of the fibre as
determined by the energy under the static tensile
stress/strain curve. An examination of the data suggests
a possible relationship between these two parameters,
with materials containing fibres with a greater strain
energy absorbing capacity offering improved lzod
energies. It appears, therefore, that this technique
forms a useful guide for assessing and evaluating the
impact resistance of composite materials. This technique also explains the results of Hancox 37 and Bader
and Ellis 2~, reported above, who found that composites
based on type II carbon fibres outperformed type I
carbon fibre composites. The former has an elastic
energy absorbing capability some four times greater
than the latter.
Although the above approach is valuable for evaluating
the impact resistance of a composite, a complete
analysis should take into consideration energy dissipation in failure processes such as fibre/matrix debonding
40,

35
I

30 ~25 ~E
2O

Q.

.E

15-10

I t{,
0.0

0.2

I,

0.4

0.6

ii

,,,

0.8

1,

1.0

Normalized strain energy


Fig. 2 Variation of Charpy impact energy with normalized
strain energy absorbing capacity of fibres (data taken from
Ref 39)

1.2

and fibre pull-out. Beaumont 4 gives expressions for


work associated with micro-mechanical fracture
processes such as debonding and fibre pull-out. The
wot'k for debonding is given as:
Wd = ad2o2fld24Ef

(1)

where d = fibre diameter, of = failure strength of the


fibre, ld = length of the debonded zone and Ef = fibre
modulus. The work to pull-out is given as:
(2)

Wp = ~d~Fg2 4

where lc = critical transfer length (-- otd/2x) and x =


constant frictional shear stress.
Beaumont concluded that post-debond fibre sliding is
the primary energy absorbing mechanism in glass fibre
composites, whereas fibre pull-out is responsible for
much of the toughness in a carbon fibre composite. An
examination of Equation (2) indicates a strong dependency of work to pull-out on fibre diameter. In theory,
increasing the diameter of the reinforcing fibre should
result in a composite with an improved resistance to
pull-out and perhaps improved toughness 1L34. Morris
and Smith41 reported results that appear to support a
fibre diameter dependence in continuous fibrereinforced composites. Their conclusions are based,
however, on differences between glass and boron fibre
composites with different fibre volume fractions and
are not, therefore, conclusive.
In recent years, fibre manufacturers have been improving the strain to failure of carbon fibres by reducing
their diameter. Typically, the first generation of carbon
fibres such as T300 and AS4 had diameters of 7--8 ~tm.
More recent fibres such as IM6 have diameters of
approximately 5 tim. By improving the strain to failure
of the fibres in this way the manufacturers have also
improved the strain energy absorbing ability of composites and thereby improved their impact resistance.
Davies et al. 42 showed that the interlaminar fracture
toughness of IMr/polyetheretherketone (PEEK)
(superior strain energy absorbing capability) is superior
to that AS4/PEEK. Curson et al. 4 drew similar conclu#ons following low velocity impact tests on these
materials.
Under conditions of low velocity impact loading, it
appears, therefore, that the improved toughness gained
by increasing the fibre diameter is offset by the reduction in fibre failure strain, resulting in a loss of strain
energy absorbing capability. It may be, however, that
increasing the fibre diameter might lead to improvements in the high velocity impact response of composite materials. Under such loading conditions, the target
response is more localized with energy being dissipated
over an area immediate to the point of impact. Consequently, the overall energy absorbing ability of the
structure is less important and local mechanisms such as
fibre pull-out and fracture become dominant.
As well as influencing the impact resistance of a composite, the properties of the fibre also have a significant
effect upon the residual load-bearing capability of the
material. Composite systems that offer excellent impact
resistance do not necessarily exhibit superior postimpact residual properties. For example, polyethylene

COMPOSITES. SEPTEMBER 1991

351

fibre composites are capable of absorbing large


amounts of energy before failure but still offer relatively poor compressive properties due to the low
stiffness of the fibres themselves. In most circumstances
a compromise is required. The following section
considers briefly the effect of fibre properties on the
post-impact strength of composite materials.
1. Effect of fibre properties on post-impact
residual strength
In recent years a concerted effort has been made by the
manufacturers of high performance fibres to improve
both the short- and long-term mechanical properties of
advanced composites. Much of this improvement has
been achieved by increasing the strain to failure of the
reinforcing fibres. In the previous section, the energy
absorbing capacity of the fibres was identified as an
important parameter in determining the level of
damage incurred in a composite laminate. Often,
materials that satisfy this condition also offer excellent
residual properties. Cantwell et al. 43 showed that this
was the case for low and high strain carbon fibre
composites. An AS4 carbon fibre composite with a
superior strain energy absorbing capacity than that of
an XAS carbon fibre composite offered superior
residual properties, Fig. 3. This is not always the case,
however. For example, if the stiffness of the fibre is
very low and its strain to failure high, a composite
containing these fibres will be capable of absorbing
large amounts of energy but will exhibit poor residual
compressive properties. In order to overcome this,
hybrid composites are frequently used, combining the
energy absorbing capability of low modulus fibres with
stiffer fibres capable of resisting compressive loads**.
As stated previously, many of the latest generation of
composites are based on fibres with smaller diameters.
Since the compressive strength of a composite depends
upon the stability of the fibres, it would be expected
that smaller diameter fibres would result in a material
with poorer compressive properties. This indeed
appears to be the case 45. However, reducing the fibre
diameter increases the energy absorbing capability of
the composite, resulting in lower levels of damage for a

given incident energy. The reductions in plain compressive properties of the composite appear to be offset by
the reduction in damage area.
2,

Conclusions

The properties of the fibres in a continuous fibre


composite have a significant effect on the impact
resistance and subsequent load-bearing capacity of
components made from such materials. For low
velocity impact loading, the ability of the fibres to store
energy elastically appears to be the fundamental
parameter in determining impact resistance. Kevlar
fibres, which have large areas under their stress/strain
curves, offer excellent impact resistance. The role of
the fibre diameter is not completely clear. A simple
pull-out model suggests that composites with larger
diameter fibres should be inherently tougher.
However, current trends are towards smaller diameter
fibres offering higher strains to failure. Any reduction
in toughness is thereby hidden by the increased energy
absorbing capacity of the fibres.
Matrix
The polymeric matrix in a fibre-reinforced composite
serves to protect, align and stabilize the fibres as well as
assure stress transfer from one fibre to another. In
general, both the stiffness and strength of the matrix
are considerably below those of the reinforcing fibre.
The latter is therefore responsible for carrying most of
the applied load in a composite component. The role of
the matrix is nevertheless critical. For example,
damage to the matrix such as impact-induced delamination can reduce the load-bearing capabifity of the
composite by up to 50% 46. As a result of this relatively
poor behaviour, much work has been undertaken in
recent years in an attempt to identify the fundamental
matrix properties that influence the impact resistance
of composite materials. Since the first generation of
matrix systems for advanced composites lacked toughness, a number of techniques have been developed to
impro~,e the toughness of these materials. These
include:
the use of plasticizing modifiers11;

1200

the addition of rubber particles such as carboxylterminated butadiene-acrylonitrile (CTBN)47--49.;

~- 1000

the addition of thermoplastic particles such as


polyethersulphone (PES)and polyetherimide
e-

~.~ ~ " ' - o

.....

o..

(pE0~52;

Tension

~ 600

a reduction in the cross-linking density of thermosets


such as epoxy resins53'54;

~, 400

the use of thermoplastic matrices such as PEEK; and

n.-

the inclusion of thin, tough layers at ply


interfaces 5,5s-57.

Compression/I'I"

200

Impact energy (J)


Fig. 3 Variation of tensile and compressive residual strength
w i t h impact e n e r g y for l o w and high strain fibres (Ref 43): e,
high strain fibres; o, l o w strain fibres

352

COMPOSITES. SEPTEMBER 1991

10

Early work by Bradshaw et al. 11 showed that adding a


plastidzer to Epikote 828 epoxy resin increased the
Mode I fracture toughness by over two orders of
magnitude. When used as a matrix system in a carbon
fibre composite, increases in toughness did result;
however, in this case the Izod impact energy was

improved by only 25%. This disappointing transfer of


toughness was explained by the fact that the Izod test
induces crack propagation across fibres rather than
between them. Improvements in the interlaminar
toughness and resistance to splitting are hidden,
therefore, by the higher energy fracture energy associated with transverse fibre failure.
Husman et al. 17conducted high velocity impact tests on
a limited number of Modmor II carbon fibre composites. Their data suggested that the impact resistance of
these materials did not depend upon the properties of
the polymeric matrix.
In a more detailed analysis by Williams and Rhodes 49,
the impact resistance of 24 modified and unmodified
carbon fibre/epoxy composites was examined. Their
experimental analysis showed that both the level of
damage incurred as well as the residual compressive
properties of the laminates varied enormously. It was
found that the brittle laminates tended to fail by
extensive delamination whereas the tougher systems
failed in transverse shear near the impact location. The
authors concluded that the tensile performance of the
neat matrix has a significant influence on the impact
behaviour of a composite structure. For improved
impact resistance, the strength of the matrix should
exceed 69 MPa and its strain to failure should be
greater than 4%. Finally, in order to ensure adequate
compressive strength, the shear modulus should be
greater than 3.1 GPa.
Hirschbuehlers7 examined a large number of plain and
interleaved composite systems under a variety of
different loading conditions in an attempt to relate
matrix properties and post-impact compressive
strength. It was found that the residual compressive
properties of the composites increased with the flexural
strain to failure of the pure resin, Fig. 4.
Hunston47 analysed data from three sources in an
attempt to identify a link between matrix properties
and composite fracture toughness. He identified a
definite correlation between the resin Mode I fracture
toughness and composite interlaminar fracture energy
as measured by the DCB specimen, Fig. 5. With brittle

polymers, the resin toughness is fully transferred to the


composite, whereas with tougher polymers the resin
toughness is only partly transferred to the composite.
In the latter it is proposed that the presence of the
fibres restricts the crack-tip plastic zone size, thereby
reducing the positive effect of the tougher matrices.
Masters3s'~ extended this approach by conducting
Mode I, Mode II and compression after impact (CAI)
tests on a number of epoxy and bismaleimide-based
carbon fibre composites. He showed that no correlation
existed between the Mode I interlaminar toughness and
the CAI properties, Fig. 6, whereas a very good agreement was found between the Mode II resistance and
residual compressive strength, Fig. 7. Similar observa
tions have been made by other workers58 . It is
clear
that the matrix in a flexurally loaded composite will be
subjected to a large Mode II component and that the
forward shear properties of the matrix will be important in determining the level of damage incurred. It is
somewhat less clear why the residual compressive
properties of the composite should be Mode II controlled since the failure process is undoubtedly complex
containing a significant Mode I component. Master's
results are impressive nevertheless.
Materials that satisfy the above condition and therefore
I

E
o

E
--

'-

0 O0

/ o

-~ Thermoset
o Experimental

O J

Toughened
thermoset
+ Thermoplastic

~
'~
E
O

Resin fracture energy (kJ m -2 )


Fig. 5 Variation of composite Mode I interlaminar fracture
energy with resin fracture energy (Ref 47)
500 ~-

a.
,'-

350 I
300

.."

250
00

E 450 7
400 ~-

AS4/907
AS4/1808-fi(m E

~ 350
==

C6000/1827-film A

_= 300

2I

AS4/1808-filrn A
AS4/1808
AS4/1808-fitm C
C60O0/1827

150

50

'~

150

100

"o

,
0.0

]M6/1808-film A

[M6/1808

250

~ l I
$
~:

, I , , , I I ~ , I. ,. , , I , , , I , L ,
2.0
4.0
6.0
8.0
10.0
12.0
Neat resin flexural strain to failure (%)

Fig. 4 Variation of residual compressive strength of impactdamaged composites with neat resin flexural failure strain
(Ref 57)

A$4/3502-fi{m A
AS4/3502

50
0
0

100

200

300

400

500

Residual compression strength after impact (MPa)


Fig. 6 Variation of residual compression strength after impact
with Mode I strain energy release rate (Ref 56)

COMPOSITES. SEPTEMBER 1991

353

4000
AS4/1808:film E

E 3500

IM6/1808-film A
AS4/3502.film A

~ 3ooo
_~ 2500

AS4/1808-film C
IM6/3100-film E
IM6(3100-21-film E

2000

eAS4/1808-fiim A

C6000/18274ilm E
AS4/907

A54/1808
....... AS4/1808
l M o / I ~ u ~ C6000/1827
AS4/3502

1500
1000

I M 6 / 3 1 0 0 e I M 6 / 3 1 0 0 - 2 1

0
0

t
100

I
200

i
300

I
400

I
500

Resin compression strength after impact (MPa)


Fig. 7 Variation of residual compression strength after impact
w i t h M o d e II strain e n e r g y release rate (Ref 56)

offer superior impact properties include thermoplasticbased composites and interleaved laminates. In recent
years considerable interest has been generated by
carbon fibre-reinforced PEEK (APC2), a semicrystalline thermoplastic compositea'~'59-62. Interlaminar fracture testing and impact loading have shown that
this material offers excellent static and dynamic toughness and is capable of absorbing a considerable amount
of energy whilst incurring only small amounts of
damage4,63,64. Scanning electron micrographs of the
fracture surfaces indicate extensive drawing and plastic
flow65. Another advantage of this material is that its
thermoplastic matrix allows rapid repair using fusion
techniques such as the hot press technique66. Here,
impact damage can be reduced or removed by simply
heating the component to a temperature above the
melting point of the matrix, reforming and cooling. The
high velocity impact response of carbon fibre/PEEK has
received very little attention. Initial testing has suggested that its high velocity impact response is perhaps
relatively poor. Dan-Jumbo et al. 59 showed that beyond
a certain velocity threshold, APC2 experienced a
sudden drop in flexural strength. Similar observations
have been observed by Morton and Godwin63 following
ballistic impact tests on this material. This will be
discussed in more detail in the section on rate effects.
These observations suggest, therefore, that care should
be exercised when attempting to relate static properties
such as interlaminar toughness and strength to characterize tlynamic properties such as impact resistance.
Polymer interleafing involves the use of high toughness
films or layers at ply interfaces in relatively brittle
materials. The inclusion of such layers increases the
laminate's interlaminar fracture toughness56 as well as
reducing the level of damage incurred for a given
incident energy67. The load-bearing properties of
damaged intedeafed composites are significantly
superior to those of conventional epoxy composites67.
Interlayer technology is still in its infancy; however,
early results are very favourable and the technique
offers enormous potential.
1. Effect of matrix properties on post-impact
residual strength

Impact-damaged composites are probably most sensi354 COMPOSITES. SEPTEMBER 1991

tive to compressive loading since impact-generated


delaminations tend to reduce the stability of the
load-beating plies resulting in premature failure
through buckling. In the previous section it was shown
that composites with high Mode II interlaminar
fracture toughnesses offer superior CAI propertiess5,56.
It appears that the majority if not all of this improvement in CAI behaviour results from the lower levels of
damage incurred during impact.
The residual tensile properties of toughened composites do not appear to be significantly better than those
of standard epoxy systems, Fig. 8. This results from the
fact that composites with tougher matrices tend to be
more notch-sensitive due to reduced splitting and
delamination around stress concentrations such as
notches or damage6s.
Toughening composites using elastomeric particles
reduces the level of delamination and therefore
enhances residual compressive properties, Fig. 9.
However, the presence of such inclusions often reduces
the glass transition temperature of the matrix material
which in turn reduces the hot-wet properties of the
composite. In many situations a compromise is therefore necessary.
1400

<.

O - -- -- Tough epoxy
Brittle epoxy

1200'
1000

80o
~

600

4o0

"0 .....

n-

200

0.0

2.0

4.0

6.0

8.0

,IL J ,I,
f ~
10.0
12.0
14.0

Impact energy (J)


Fig. 8

Variation of residual tensile strength with impact energy

for tough and brittle epoxy-based composites (Ref 75)

1000
~
- -

e-

Toughened epoxy
o Brittle epoxy

900,

800,
700
600

s~

400

72
8

200
100
0

2a

4I

6f
8~
Impact energy (J)

110

112

14

Fig. 9 Residual compressive strength of toughened and brittle


epoxy-based composites as a function of impact e n e r g y (Ref 75)

2.

Conclusions

It is clear that matrix properties play a significant role


in determining the impact resistance and subsequent
load-beating capability of a fibre-reinforced plastic. At
present, a significant effort is being made to improve
the post-impact compressive properties of composites.
It appears that materials with high Mode II interlaminar fracture toughnesses offer superior compression
strengths after impact. Care should be exercised,
however, when correlating data obtained under static
loading conditions with results from dynamic tests.

1000
~

800

0 Untreated fibres

"

-=E

--600

'\l

Studies have shown that varying the level of surface


treatment applied to a carbon fibre composite can
change the mode of failure as well as many fundamental mechanical properties H'69-7z. Composites with low
levels of fibre surface treatment fail at relatively low
stresses when loaded transversely to the fibres, leaving
smooth fibres on the fracture surface. Increasing the
level of treatment applied to the fibres increases the
transverse failure stress and failure occurs within the
matrix, i.e., the interphase region is no longer the
weakest link in the composite.
Rogers e t al. 71 showed that improving the fibre/matrix
bond strength in a carbon fibre-reinforced epoxy
resulted in a fourfold increase in the incident impact
energy required to initiate damage. At higher energies
the load-bearing properties of composites with surfacetreated fibres drops dramatically until the perforation
limit is reached72. Bless and Hartman 73 showed that the
perforation threshold energy in a surface-treated
composite is significantly lower than that of a similar
untreated laminate.
This behaviour has been explained by Dorey~'2, who
showed that the transverse fracture energy of a composite, a fundamental parameter for determining
resistance to penetration and perforation, depends
strongly upon the fibre/matrix bond strength. Carbon
fibre-reinforced epoxies with untreated fibres offer
transverse fracture energies as high as 60 kJ m-2 (Ref
72). Transverse failure in composites with high levels of
fibre surface treatment absorbs considerably less
energy, with quoted ~2 transverse fracture energies
being as low as 20 kJ m-2. At energies levels above that
required to achieve perforation, damage in a surfacetreated composite tends to be localized around the
point of impact, often taking the form of a clean hole 71.
The post-perforation residual properties of treated
composites are generally superior to those of untreated
composites as shown in Fig. 10.

I/J

Penetration

..- - O -

Interphase
The strength of the bond between the matrix resin and
the fibre reinforcement is a controlling factor in determining the mechanical performance of most polymer
composites. In general, the surface of the fibres is
treated by an oxidative process in order to improve the
level of adhesion between matrix and fibre. Initially,
this interracial zone was considered as being a twodimensional surface with effectively zero thickness.
However, more recent studies have shown that this
region is in fact three-dimensional, having its own
distinct properties 69.

Surface-treated fibres

(3.

10

12

14

Incident energy (J)


Fig. 10 Residual flexural strength vs. impact energy for ballistically impacted surface-treated and untreated carbon fibre
composites (Ref 72)

McGarry et al. TM attempted to enhance the impact


resistance of T300/MY-720 carbon fibre/epoxy by
coating the surfaces of the fibres with a thin layer of
CTBN rubber. Gardner impact tests showed that
treating the fibres in this way improved the threshold
energy for first damage significantly. Further, above
this threshold the level of damage for a given impact
energy was less in the modified composite.

1. Effect of interphase properties on post-impact


residual strength
Dorey 75 has shown that increasing the strength of the
fibre/matrix bond increases the interlaminar shear
strength (]LSS) of the composite until a plateau is
reached beyond which point no significant increase is
possible, Fig. 11. Over this range of treatment levels
the notched tensile strength falls dramatically. Increasing the ILSS in this way suppresses the formation of
delaminated zones in the region of stress concentrations, rendering the material more notch-sensitive.
Consequently, even though surface treatment of the
fibres reduces the level of damage for a given energy,
the increased notch-sensitivity of the laminate results in
poorer residual tensile properties, Fig. 1 2 7 5 . Conversely, treating the fibres improves the post-impact
compressive properties as shown in the lower part of
Fig. 12. Clearly, the level of surface treatment applied
to the fibres in a multidirectional composite will depend
upon the operational conditions the component will
encounter. In general, a compromise is sought in which
the fibres are given intermediate levels of treatment.

2.

Conclusions

The level of treatment applied to the surface of the


fibres in a composite material has a significant effect
upon both its impact resistance as well as its residual
load-carrying capability. In general, impact on composites with low levels of fibre surface treatment generates
large areas of splitting and delamination with severe
effects on the compressive properties of the material.
Localized impact loading on highly treated fibre
composites results in a smaller, more localized damage
zone, a lower perforation threshold and improved

COMPOSITES. SEPTEMBER 1991 355

'[

Fibre stacking sequence


Composite materials offer a unique advantage in that
properties such as strength and stiffness can be tailored
to meet specific design requirements through a careful
selection of the fibre stacking sequence. Considerable
work has shown that the impact resistance of composite
materials also depends upon the specific order in which
the plies are stacked6'~'76-s. For example, unidirectional composites having all their fibres aligned in one
direction fail by splitting at very low energies and are
therefore highly unsuitable for applications where
impact loading might occur 11. Following impact tests
on a series of (0, +/-45 ) laminates, Dorey 1~as well as
Morton and Godwin 63 showed that composites having
+ / - 4 5 surface plies offered a superior impact resistance and improved residual strengths. It was suggestedsl that + / - 4 5 plies increased the flexibility of the
composite, thereby improving its ability to absorb
energy elastically. Further, placing such plies on the
surface of a composite serves to protect the loadbearing 0 plies against damage induced by the imping,
ing projectilesl. These ideas were supported by Stevanovic et al. s, who conducted instrumented Charpy
tests on a series of multidirectional T300 carbon fibre
composites. They showed that (+/-45 ) composites
were capable of absorbing considerably more energy
than (0,90), (0,+/--45 ) and (00,900,+/-45 ) laminates.

T/T100

1.0

On/OnlO0

0.5

I
100

I
200

A somewhat different approach was adopted by Nolet


and Sandusky in the design of a composite leading edge
for the A10 Thunderbolt 8z. They proposed that the
leading edge be very stiff in order to 'split off' the
impinging bird with little or no loss in strength or
stiffness. They suggested that in order to achieve this
requirement, a high percentage of carbon fibres should
be oriented in the spanwise direction of the aircraft's
leading edge.

I
700

Surface treatment (%)

Fig. 11 Variation of interlaminar shear strength and notched


strength of carbon fibre composites as a function of fibre
surface treatment (Ref 75)

1.0

Treatment level

0%
o . . . . . 5%
" - - - - 100%

0~
~

A
tO
n

The authors s3 showed that damage initiation in a series


of (+/-45 ) laminates subjected to low velocity impact
depended upon the thickness and therefore the stiffness
of the composite. Initial failure in thin, flexible targets
occurs in the lowermost ply as a result of the tensile
component of the flexural stress field. Damage in
thicker, stiffer targets initiates at the top surface due to

Tension

t3
v
.c 0.6
~,+\
\~,\
-~ 0.4
'~,,. ~.
"0

X~O~

o
...

1.6

~'II~
~+

~
"~_
=t .- - . _--.

.. .. .. .

1,2

1.0

Compression

0.0

O3

Fig. 12 Residual strength of treated and untreated carbon fibre


composites vs. impact energy (Ref 75)

compressive properties. However, the increased notch


sensitivity associated with fibre surface treatment
results in a reduction in the post-impact tensile strength
of the material. The level of treatment applied to the
surface of the fibres will depend, therefore, upon the
desired application.
C O M P O S I T E S . S E P T E M B E R 1991

o.8

10

Impact energy (J)

356

Lower [ Upper
surf~rface

'1,4
.....

0.2

............
".. . . . . . .

0.6
0.4

0,2
0.0

,,l..,n

, , , l . . , l , , , J , , , l . , , l , , , f . , , l , , ,

4
5
6
7
Target thickness (mm)

Fig. 13 Low velocity impact energy to initiate damage vs.


target thickness for (+/-45 ) CFRP composites (Ref 29)

10

the contact stress field. This is shown in Fig. 13. This


curve clarifies Dorey's sl claim that increasing the
flexural stiffness of a target, for example, by placing
fibres on the surface of a laminate, can enhance its
impact resistance. This is true for the range of stiffnesses where initial failure occurs at the top surface of
the component. In more flexible targets, however,
reducing the flexural stiffness may precipitate failure at
a lower incident energy.
A detailed study by Hong and Liu 6 identified fundamental aspects in the development of damage in glass
fibre-reinforced plastic (GFRP) subjected to high
velocity impact loading. They showed that increasing
the angle q in a (~,q,0~) laminate resulted in greater
delamination-type damage for a given incident energy,
Fig. 14. Increasing q in this way also had the effect of
reducing the first damage threshold energy. The
authors als0 showed that for a given energy, increasing
the thickness of the GFRP target resulted in an increase
in delaminated area. This increase in damage area may
result from the reduction in the target's energy absorbing capability as proposed by DoreySL Lius4 extended
this work by developing a simple model for predicting
the likely delamination sites in a number of different
composites. It was suggested that delamination in
multi-angle composites is more likely to occur at
interfaces where the mismatch in bending stiffness is
greatest, for example, between +/--45 plies. Liu
showed experimentally that the level of delamination in
a glass/epoxy composite increased as angle q in a (0,q)
laminate increased, i.e., as the bending stiffness
mismatch increased This evidence suggests that if
delamination needs to be suppressed, laminates with
sudden large changes in fibre direction should be
avoided.
Other techniques to reduce impact-induced delamination include the use of woven fabrics46"85,
hybridization 16'a6-88 (for example, carbon fibres with
Kevlar fibres) and three-dimensional stitchings9'9. The
first of these techniques involves replacing the unidirectional +/-45 plies in a multidirectional composite by a
+/-45 woven fabric. The three-dimensional nature of
the fabric helps suppress the formation of delaminated
zones at this critical interface.

4I
35

A
cq

30

o
=

[ 05/90s/05
[ 0s/605/05
[ 05/455]0s
[ 05/30s/05
[ 05/15~/0~

[3

25

]
]
]
]
]

oo

++0~]

20
15
r~

+(

10

lid

C3
0

10

15

20
25
30
Impact energy (J)

35

40

45

Fig. 14 D e l a m i n a t e d area vs. i m p a c t energy for i m p a c t e d


(0s*,05",05") GFRP l a m i n a t e s (Ref 6)

50

The impact resistance of carbon fibre composites can


be enhanced considerably by incorporating plies of
lower modulus fibres 16,s6-s8. In order to assure compatibility, the matrix resin is usually the same in the two or
more constituent materials. Hancox and Wellsa7
showed that the Izod impact energy of an HT-S carbon
fibre composite could be increased by 500% through
hybridization with E-glass fibres. As well as reducing
the basic price of the composite, the addition of the
glass fibres was found to change the mode of fracture
from a clean break to a delamination-type noncatastrophic failure Similar conclusions were drawn by
Helfinstine 16following Charpy impact tests on KevlarT300 carbon fibre hybrids. However, the absolute
magnitude of the increases was less impressive than
that reported by the previous workers for GFRP.
Dorey e t al. ~ assessed the high velocity impact
response of a number carbon-Kevlar hybrid laminates.
They showed that the addition of the lower modulus
Kevlar fibres increased the threshold energy for the
onset of damage by up to four times.
At present many workers are assessing the feasibility of
using weaving and braiding techniques in order to
improve the damage tolerance of fibre-reinforced
polymer composites s9-93. Su 93 conducted Mode I
delamination tests on both stitched and non-stitched
AS4 carbon fibre/J1 (a semi-crystalline thermoplastic).
His results showed that stitching with Kevlar fibre
resulted in a 100% increase in interlaminar fracture
toughness. Instrumented drop-weight impact tests on a
number of 2- and 3-D composites indicated that the
latter offered a superior impact resistance, the presence
of the third dimension reinforcement served to inhibit
the propagation of delaminated zones. Fabricating 3-D
structures is clearly more expensive and time consuming than constructing with conventional 2-D prepregs.
In order to reduce these costs, it has been proposed
that a selective procedure be adopted, that is, 3-D
reinforcement be used at critical ply interfaces or at
component edges 93.

1. Effect of fibre stacking sequence on


post-impact residua/ strength
In this section the role that the fibre stacking sequence
plays in determining the residual properties of impactdamaged composites is discussed As outlined above,
much of the work published in the literature concerns
the residual compressive properties of damaged
composites since this is considered to be the most
critical form of loading condition Certain conflicts may
exist, however, when considering the optimum fibre
stackin.8 sequence for residual compressive strength.
DoreyTM suggested that for improved impact resistance
the +/--45 fibres should be located on the outermost
surface of the composite. This may not be an ideal
stacking sequence for stability in compression Here,
stiffer laminates, for example, those with surface 0
fibres, are better suited to in-plane compressive
63
loading. Nevertheless, Morton and Godwin
have
shown that an APC2 ( ~ , +/-45)2s laminate offers
inferior properties to those of a (+/-45,~,+/--45,0)s
plate, Fig. 15.

COMPOSITES, SEPTEMBER 1991 357

1400

they split and fail at low energies. The mismatch in


bending stiffness between two plies appears to have a
significant effect upon the level of damage incurred at
that interface. Damage appears to be greatest where
ply orientation changes of 90 occur. This suggests that
for containment of damage laminates with abrupt
changes in fibre direction should be avoided. Other
ways to suppress damage include the use of woven
fabrics, the use of hybrid composites or stitching at
desired locations.

45 outside

1200 I
/ ~ 0 outside
I000 r

==
800 "

L
i
6OO

E
oo

/k
4O0

Jk

Geometry

r~
200 i
0
5

10

15

Incident impact energy (J)

Fig. 15 Effect of placing 45 plies on the outer surface of a


16-ply carbon fibre/PEEK laminate (Ref 63!
1000

(3 Non-woven
Mixed-woven

.~
8o0
Qe-

g, coo
e--

A
W

.o_ 400
e~
E
o
~

200

3
4
Impact energy (J}

Fig. 16 Effect of replacing the + / - 4 5 plies in a 16-ply (0,+/-

45) CFRP composite with a woven fabric (Ref 46)

In the previous section it was stated that the use of


woven +/--45" fabrics in (0", +/--45 ) laminates serves to
reduce the overall level of delamination under impact
loading. The subsequent residual strengths of the
mixed-woven composites were superior to those of the
standard material manufactured from unidirectional
plies, Fig. 16.
Similar improvements in residual strength have been
noted following impact on stitched carbon fibre composites 92. Compression after impact tests on a number
of AS4/3501--6 laminates92 showed that stitched laminates offered residual strengths up to 100% greater
than their unstitched counterparts. One of the disadvantages of this process is that the undamaged compressive strength of the material is reduced by up to
20% 92.

2. Conclusions
The impact resistance of a multidirectional laminate is
strongly dependent upon the specific orientation of the
plies. Unidirectional laminates should be avoided since
358

COMPOSITES. SEPTEMBER 1991

Geometry is a fundamental parameter in determining


the impact response of a composite componenta,rt A5,63,75,83.Low velocity impact tests on CFRP
have shown that the mode of failure in a simple beam
may vary depending upon its span-to-depth ratio. Short
thick specimens tend to fail in an interlaminar mode
whereas as long thin beams failed in a flexure 11.
Broutman and Rotem 15 showed that increasing the size
of a GFRP beam increased its energy absorbing capability under low velocity impact conditions. However,
doubling the size of the beam did not result in an
equivalent increase in energy absorption. Similar tests
o n CFRP 83 have shown that both the low velocity first
damage threshold energy and perforation limit of a
CFRP (0,-I-/---45 ) beam increased linearly with increasing beam length. These results indicated that the
target's ability to absorb energy in elastic deformations
determines its low velocity impact resistance. However,
increasing the volume of the target does not necessarily
increase its impact resistance, for example, small beams
may be capable of absorbing greater energy than a
large circular plate s3. Further, the process of failure in
a large component often differs from that in a simple
laboratory specimen. Clearly, care needs to be taken
when using small, simple specimens to characterize the
impact response of larger, more complex structures.

High velocity impact tests on CFRP indicated that the


areal geometry of the target is less important at high
rates of strain 93. Ultrasonic C-scans of impacted
specimens showed that the level of damage in a small,
50 mm long beam was the same as that in a 150 mm
coupon, Fig. 17. This suggests that high velocity impact
loading by a light projectile induces a localized form of
target response in which much of the incident energy of
the projectile is dissipated over a small zone immediate
to the point of impact. Tests on large plates have
substantiated this claim 93 and it appears that under
certain conditions small simple coupons can be used to
characterize the high velocity/low mass impact response
of composite structures.
Few workers have undertaken tests on full-size engineering components primarily as a result of the high
costs involved. Gause et al. 94 showed that the curvature
of the test component or structure influences the level
of damage incurred during impact loading. It appears
that negative curvatures inhibit delamination growth.
Madan and Sutton 95 conducted low velocity impact
tests on a number of stiffened panels designed for use
in advanced wing structures. Their results indicated
that the impact resistance of the structure was poor if

Impact energy = 1.8 J

Length = 50 mm

n D
Length = 75 mm

Length = 100 mm

Impact energy = 7.0 J

Length = 50 mm

Length = 75 mm

Length = 100 mm

U U
Length = 150 mm

Length = 150 mm

Fig, 17 Effect of beam l e n g t h o n d a m a g e d e v e l o p m e n t in a


16-ply (0,+1-45 ) CFRP composite subjected t o h i g h v e l o c i t y
i m p a c t l o a d i n g (Ref 29)

the flanges were either too stiff or too flexible. This is


clearly related to the structural stiffness effects shown
previously in Fig. 13. For optimum impact resistance a
compromise is again required.
2.

Conclusions

In order to design components and structures for


impact resistance, geometrical effects need to be fully
understood. In the case of low velocity impact loading,
the size of the specimen or component is a critical
parameter in determining its dynamic response. Here
again the response of the target as well as the amount
of damage incurred is related to the target's ability to
store energy elastically. As a result of the lower level of
transverse constraint, beams tend to be capable of
absorbing more energy than larger structures such as
circular plates. Care has to be taken, therefore, when
using simple beam-like specimens to evaluate the
dynamic response of more complex structures.
High velocity impact loading by a light projectile
induces a very localized form of target response,
resulting in much of the incident energy being dissipated in a very small volume. Here, the areal geometry
of the target is less important and simple coupons can
often be used to characterize the response of the
full-scale structure.
Rate
The rate at which the structure is loaded affects both

the material's behaviour as well as the structural


response of the target. The latter has already been
discussed in some detail above. In summary, it appears
that the low velocity impact resistance of a composite is
strongly dependent upon its ability to absorb energy
elastically. Consequently, the strain energy absorbing
capability of the fibres as well as the geometrical
configuration of the target are of great importance. At
very high rates of strain the structure responds in a
local mode and the strain energy absorbing of the fibres
and structure is less important. Here, the magnitude of
the energy dissipated in mechanisms such as delamination, debonding and pull-out may become important.
In recent years more and more attention has been given
to determining the rate dependence of the ultimate
properties of composites and their
constituentsz~'24`26'28,~-gs. Harding and
co-workers25'27'2s'96 have examined the strain rate
sensitivity of Kevlar, GFRP and C'FRP. They have shown
that carbon fibre composites are rate-insensitive when
tested in fibre-dominated modes, whereas GFm' and
Kevlar composites exhibit a distinct rate-dependent
behaviour with modulus and tensile strength increasing
with rate. Other workers have assessed the rate dependence of matrix-dominated modes of failure such as
interlaminar fracture23'99. Such tests are particularly
useful since delamination (a matrix-dominated mode of
failure) is particularly detrimental to the compressive
strength of a laminated composite ~'11. DCB tests on
carbon fibre/epoxy composites t2 have shown that the
Mode I interlaminar toughness does not vary with
strain rate, Fig. 18. Similar tests on carbon fibre/PEEK
(APC2), a thermoplastic matrix composite, have
identified a distinct rate dependence :3, Fig. 18. Over a
wide range of strain rates the Mode I fracture toughness remains invariant of strain rate. However, beyond
a certain threshold, the toughness drops dramatically to
approximately 20% of its original value. This gives
cause for concern, suggesting that the impact resistance
of this material may be poor at high rates of strain.
Indeed, high velocity impact tests on APC2 have
suggested that beyond a certain threshold velocity a
change in failure mode o c c u r s 63 and the material
experiences a sudden drop in mechanical performance59.
5O00
CF/PEEK

D Unstable
Stable

4000
CF/epoxy
E"
I

3000

ca

D=
mo

o
o o

oo

ta
o

o o

[]

oo

oo

1000

10-6

10-s

10-4

+,,
10-3

10-2

10-1

10 o

(s -~ )
Fig. 18 Variation o f Gic w i t h s t r a i n rate f o r c a r b o n fibre/PEEK
and a carbon fibre/epoxy composite (Refs 24 a n d 102)

COMPOSITES. SEPTEMBER 1991 359

These results suggest that care should be taken when


using data obtained from static tests in order to characterize the dynamic response of composite materials.
Brittle thermosets such as epoxy resins do not appear
to exhibit a significant rate-dependent behaviour
whereas tougher systems such as the latest generation
of thermoplastic-based composites do 12. Clearly,
simple static tests would fail to identify such strain rate
sensitivity and may rank the materials incorrectly.

the target response. Low velocity impact loading by


a heavy object induces an overall target response,
whereas high velocity impact by a light projectile
induces a localized mode of target deformation
resulting in energy being dissipated over a small
region immediate to the point of impact. In the
latest generation of tough composites matrixdominated modes of fracture appear to show a
distinct rate dependency and care should be taken
when using static tests to characterize dynamic
behaviour.

SUMMAR Y AND CONCLUSIONS


Concerns expressed regarding the impact resistance of
fibre-reinforced composites are without doubt wellfounded. Low energy impacts are capable of generating
large areas of delamination, resulting in significant
reductions in residual strength. In recent years a
concerted effort has been made by the materials
manufacturers to improve the impact resistance and
damage tolerance characteristics of continuous fibre
composites. In this review some of the fundamental
parameters governing the impact response and subsequent load-bearing properties of components manufactured from these materials have been identified. The
following conclusions may be drawn from the present
review.
1) The strain energy absorbing capacity of the fibres is
one of the most important parameters in determining the impact resistance of a composite structure.
Fibres that have a large area under the stress/strain
curve tend to be better suited to energy absorbing
applications.
2) The Mode II (forward shear) properties of the
matrix appear to determine the level of damage
incurred during impact and therefore the residual
compressive properties of the composite. For good
compression strength after impact the Mode II
interlaminar fracture toughness should be high.
3) The strength of the fibre/matrix interphase region
can be adapted to the required application. For
example, if a projectile has to be stopped and
residual properties are not important, then the
interphase region should be weak to encourage
failure through gross splitting and delamination. If
damage containment is required, then the level of
treatment applied to the surface of the fibres should
be greater.
4) The fibre stacking sequence determines both the
elastic energy absorbing capability of the composite
as well as the failure mode. For damage containment, laminates with abrupt changes in fibre
orientation (for example, one with +/-45 interfaces) should be avoided.
5) Geometrical effects are significant under conditions
of low velocity impact loading. Varying the
geometry changes the target's ability to store energy
and therefore its impact resistance. Large targets
are not necessarily better energy absorbers than
small coupons. Care should be taken, therefore,
when interpreting data from tests on laboratory-size
specimens.
6) Varying the impact velocity and therefore the strain
rate affects both the material's properties as well as
360

COMPOSITES. SEPTEMBER 1991

ACKNOWLEDGEMENTS
The authors would like to acknowledge the financial
support of the National Science Foundation Science
and Technology Center for High Performance
Polymers, Adhesives and their Composites at Virginia
Tech. This paper was originally submitted for 'Bonding
and Repair of Composites II', Zurich, Switzerland,
March 1991 (this conference was cancelled).
REFERENCES
1 Mmders, P.W. and Harris, W.C. 'A parametric study of composite performance in compression-after-impact testing'
SAMPE Journal 22 (1986) pp 47-51
2 Almy, M. 'Post damage capabilityofcarbon fibre reinforced
matrices' Proc lnt Conf on Polymersfor Composites (The
Plastics and Rubber Iost, London, UK, 1987) pp 11.1-11.10
3 Stellbrink, K.K.U. 'Improved impact damage tolerance' Proc
European Syrup on DamageDevelopmentand FailureProcesses
in CompositeMateriah edited by I. Verpoest and M. Wevers
(Leuven, Belgium, 1987)
4 Curson, A.D., Leach, D.C. and Moore, D.R. 'Impact failure
mechanisms in carbon fiber/PEEK composites' J Thermoplastic
Composite Mater3 (1990) pp 24-31
5 Redmk, S. and Sna, C.T. 'Optimal use of adhesive layers in
reducing impact damage in composite laminates' in Composite
Structures, Vol 2, DamageAssessmentand MaterialEvaluation
(Elsevier Applied Science Publ, 1987) pp 2.18-2.31
6 Hong, S. and IAu, D. 'On the relationship between impact
energy and delamination area' Exptl Mech 13 (1989) pp 115-120
7 Boll, D.J., Bracero, W.D., Weidaer, J.C. and Man'i, W.J. 'A
microscopy study of impact damage on epoxy-matrix carbon
fibre composites' ProcInt Conf on Post FailureAnalysis Techniquesfor FiberReinforced Composites, OH, USA, 1985paper 8
8 Takeda, N., Sierakewski, R.L., Ross, C.A. and Malvera, L.E.
'Delamination-crack propagation in ballistically impacted
glass/epoxy composite laminates' Expil Mech 4 (1980) pp 19-25
9 Takeda, N., Sierakowski, R.L. and Malvern, L.E. 'Transverse
cracks in glass/epoxy cross-ply laminates impacted by projectiles'
J MaterSci 16 (1981) pp 2D08-2011
10 Shadbelt, P.J., Cerraa, R.S.J. aad Ruiz, C. 'A preliminary
investigation of plate perforation by projectiles in the subordnance range' Report No 1372/81 (Univ of Oxford, UK, 1981)
11 Bradshaw, F.J., Derey, G. and Skley, G.R. 'Impact resistance of
carbon reinforced plastics' RAE TR 72240(MOD, 1972)
12 Avery, J.G. DesignManualfor Impact Damage TolerantAircraft
Structures, AGARDograph No 238 (NATO, 1981)
13 Retem, A. 'Residual flexural strength of FRP composite specimens subjected to transverse impact loa"drag'SAMPEJourna124
No 2 (1988) pp 19-25
14 Rea, C.A. and Sierakowski, R.L. 'Studies on the impact resistance of composite plates' Composites4 (1973) pp 157-161
15 Broatman, L.J. gad Retem, A. 'Impact strength and toughness
of fiber composite materials' in Foreign ObjectImpact Damage
to Composites, ASTM STP 568 (American Society for Testing
and Materials, 1975) pp 114--133
16 Helflmtim~,J.D. 'Charpy impact of unidirectional graphite/
aramid/epoxy hybrid composites' in CompositeMaterial~:
Testingand Design (Fourth Conference),ASTM STP 617
(American Society for Testing and Materials, 1977) pp 375-388

17 Husmn, G.E., Whitney, J.M. and Halpin, J.C. 'Residual


strength characterization of laminated composites subjected to
impact loading' in Foreign Object Impact Damage to Composites
op.cit, pp 92-113
18 Beaumont, P.W.R., Riewald, P.G. and Zweben, C. 'Methods for
improving the impact resistance of composite materials' ibid.
pp 134-158
19 Novak, R.C. and de Crescente, M.A. in Composite Materials:
Testing and Design (Second Conference), ASTM STP 497
(American Society for Testing and Materials, 1972) pp 311-323
20 Cheresh, M.C. and McMichael, S. 'Instrumented impact test
data interpretation' in Instrumented Impact Testing of Plastics
and Composite Materials, ASTM STP 936 edited by S.L.
Kessler, G.C. Adams, S.B. Driscoll and D.R. Ireland
(American Society for Testing and Materials, 1987) pp 9-23
21 Bader, M.G. and Ellis, R.M. 'The effect of notches and
specimen geometry on the pendulum impact strength of uniaxial
CFRP' Composites 5 (1974) pp 253-258
22 Kakarala, S.N. and Roche, J.L. 'Experimental comparison of
several impact test methods' in Instrumented Impact Testing of
Plastics and Composite Materials op. cit. pp 144-162
23 GilllespieJr, J.W., Carlsson, L.A. and Smlley, A.J. 'Rate
dependent Mode I interlaminar crack growth mechanisms in
graphite/epoxy and graphite/PEEK' Composites Sci and Technol
28 (1987) pp 1-15
24 Beguefin, P. and Barbezat, M. 'Caract6risation m6canique des
polym6res et composites ~tl'aide d'une machine d'essais rapides'
Proc 5th Journ~e Nationale D YMA T, Bordeaux, France, 1989
25 Harding, J. 'Impact damage in composite materials' Sci and
Engng of Composite Mater I (1989) pp 41--68
26 AmiJima, S. and Fujli, T. 'Compressive strength and fracture
characteristics of fiber composites under impact loading' in
Progress in Science and Engineering of Composites, Proc 4th lnt
Confon Composite Mater, 1982 edited by T. Hayashi, K.
Kawata and S. Umekawa (Okasan & Co Ltd, Tokyo) pp 399413
27 Bai, Y. and Harding, J. 'Fracture initiation in glass reinforced
plastics under impact compression' in Structural Impact and
Crashworthiness, Vol2 edited by J. Morton (Elsevier Applied
Science Publishers, 1984) pp 482--493
28 Harding, J. and Welsh, L. 'A tensile testing technique for
fibre.reinforced composites at impact rates of strain' J Mater Sci
18 (1983) pp 1810-1826
29 Cantwell, W.J. 'Impact damage in carbon fibre composites' PhD
thesis (Univ of London, UK, 1985)
30 Rhodes, M.D., winlams, J.G. and Starnes Jr, J.H. 'LOwvelocity impact damage in graphite-fiber reinforced epoxy
laminates' Proc 341h Ann Tech Conf, Society of the Plastics lnst,
1979section 20-D pp 1-10
31 Morton, J. manuscript in preparation
32 Degrleck, J. and Dechaene, R. 'Real time recording of transverse
impact experiments on composite laminates' in Composites
Evaluation, Proc 2nd lnt Conf on Testing, Evaluation and
Quality Control of Composites (TEQ C 87) (Butterworths,
London, UK, 1987) pp 61--68
33 Elber, W. 'Failure mechanics in low velocity impact on thin
composite plates' NASA Technical Paper 2152 (1983)
34 Vedula, M. and Koezak, M.J. 'Impact resistance of cross-plied
polyphenylene sulfide composites' Proc Fourth Japan-US Conf
on Composite Materials (Technomic Publ Co Inc, 1988) pp 7281
35 Clark, G. 'Modelling of impact damage in composite laminates'
Composites 20 (1989) pp 209-214
36 Adams, D.F. and Miller, A.K. 'An analysis of the impact
behavior of hybrid composite materials' Mater Sci and Engng 19
(1975) pp 245-260
37 Hancox, N.L. 'Izod impact testing of carbon fibre reinforced
plastics' Composites 2 (1971) pp 41-45
38 Chamls, C.C., Hanson, M.P. and Seraflni, T.T. 'Impact resistance of unidirectional fiber composites' in Composite Materials:
Testing and Design (Second Conference) op.cit, pp 324-349
39 Chamis, C.C. and Sinclair, J.H. 'Impact resistance of fiber
composites: energy absorbing mechanisms and environmental
effects' in Recent Advances in Composites in the United States
and Japan, ASTM STP 864 edited by J.R. Vinson and M. Taya
(American Society for Testing and Materials, 1985) pp 326-345
40 Beaumont, P.W.R. 'Fracture mechanisms in fibrous composites'
in Fracture Mechanics, Current Status, Future Prospects edited
by R.A. Smith (Pergamon Press, 1979) pp 211-233

41 Morris, A.W.H. and Smith, R.S. 'Some aspects of the evaluation


of the impact behaviour of low temperature fibre composites'
Fibre Sci and Technoi 3 (1971) pp 219-242
42 Davies, P., Cant'well, W.J., Richard, H., MouHn, C. and
Kanseh, H.H. 'Interlaminar fracture testing of carbon fibre/
PEEK composites validity and applications' in Developments in
the Science and Technology of Composite Materials, Proc
ECCM3, Bordeaux, 1989 edited by A.R. Bunseli, P. Lamicq
and A. Massiah (Elsevier Applied Sci Publ) pp 747-755
43 Cantwell, W.J., Curtis, P.T. and Morton, J. 'An assessment of
the impact performance of CFRP with high strain carbon fibres'
Composites Sci and Techno125 (1986) pp 133-148
44 De~-ey,G., Skley, G.R. and H u t e h t ~ , J. 'Impact properties of
carbon fibre/Kevlar 49 hybrid composites' Composites 9 (1978)
pp 25-32
45 Lesser, D. and Leach, D. 'Compressive properties of thermoplastic matrix composites' Proc 341h lnt SAMPE Syrup, May 1989
pp 1464-1473
46 Cantweli, W.J., Curtis, P.T. and Morton, J. 'Low velocity
impact damage tolerance in CFRP laminates containing woven
and non-woven layers' Composites 14 (1983) pp 301-305
47 Humtoa, D.L. 'Composite interlaminar fracture: effect of matrix
fracture energy' Composites Tech Review 6 (1984) pp 176-180
48 Beseem, W.D., Cottingham, R.L., Jones, R.L. and Peyser, P.
'The fracture of epoxy and elastomer modified epoxy polymers
in bulk and as adhesives' J Appl Polym Sci 19 (1975) pp 25452562
49 Williams, J.G. and Rhodes, M.D. 'Effect of resin on impact
damage tolerance of graphite/epoxy laminates' in Composite
Materials: Testing and Design (Sixth Conference), ASTM
STP 787 edited by I.M. Daniel (American Society for Testing
and Materials, 1982) pp 450--480
50 Budmall, C.B. and Partridge, I.K. 'Phase separation in epoxy
resins containing polyethersulphone' Polymer 24 (1983) pp 639644
51 Recker, H.G., AUspaeh, T., Alstadt, V., Folda, T., Heckman,
W., lttemann, P., Teach, H. and Weber, T. 'Highly damage
tolerant carbon fiber epoxy composites for primary aircraft
applications' SAMPE Quarterly 21 No 1 (1989) pp 46-51
52 l)lamant, J. and Moulton, R.J. 'Development of resins for
damage tolerant composites w a systematic approach' SAMPE
Quarterly 16 No 1 (1984) pp 13-21
53 Bravene:, L.D., Filippov, A.G. and Dewhirst, K.C. 'New
concepts in damage to tolerant composites-- lightly crosslinked
thermosets' Proc 341h lnt SAMPE Syrup, May 1989pp 714--725
54 Sykes, G.F. and Steakley, D.M. 'Impact penetration studies of
graphite/epoxy laminates' Proc 12th Nat SAMPE Tech Conf,
Oct 1980 pp 482-493
55 Evam, R.E. and Masters, J.E. 'A new generation of epoxy
composites for primary structural applications: materials and
mechanics' in Toughened Composites, ASTM STP 937 edited by
N. Johnston (American Society for Testing and Materials, 1987)
pp 413-436
56 Masters, J.E. 'Correlation of impact and interleaf delamination
resistance in intefleafed laminates' Proc Sixth lnt Conf on
Composite Mater~Second European Conf on Composite Mater
edited by F.L. Matthews, N.C.R. Buskell, J.M. Hodgkinson and
J. Morton (Elsevier Appfied Science Publ, 1987) pp 3.96-3.107
57 Hlrschbuehler, K.R. 'A comparison of several mechanical tests
used to evaluate the toughness of composites' in Toughened
Composites op.cit, pp 61-73
58 Tobukure, K., Odagiri, N., Ite, Y. and Nishimura, T. 'High
impact resistance CFRP from thermosetting system of Torayca
Tg00H/3900 prepreg' in New Generation Materials and Processes
edited by F. Saporiti, W. Merati and L. Peroni (SAMPE, 1988)
pp 293-306
59 Dan-Jumbo, E., IAx~ood, A.R. and San, C.T. 'Impact damage
characteristicsof bismaleimides and thermoplastic composite
laminates' in Composite Materials: Fatigue and Fracture, 2nd
Vol, ASTM STP 1012 edited by P.A. Lagace (American Society
for Testing and Materials, 1989) pp 356-372
60 Leach, D.C. and Moore, D.R. 'Toughness of aromatic polymer
composites reinforced with carbon fibres' Composites Sci and
Techno125 (1985) pp 131-161
61 Davies, P., Cantwell, W., Moulin, C. and Kansch, H.H. Composites Sci and Techno136 (1989) pp 153-166
62 Davies, P., CamtweU,W.J. and Kamch, H.H. 'Measurement of
initiation values of GIc in IM6/PEEK composites' Composites
Sci and Techno135 (1989) pp 301-313

COMPOSITES. SEPTEMBER 1991 361

63 Morton, J. and Gedwtn, E.W. 'Impact response of tough carbon


fibre composites' Composite Struct 13 (1989) pp 1-19
64 Moere, D.R. and Predlger, R.S. 'A study of low energy impact
of continuous carbon fibre reinforced composites' presented at
the 8th Ann NRCC/IMRI Syrup, Composites 87, Quebec,
Canada, Nov 1987
65 Davies, P. 'Delamination behaviour of thermoplastic matrix
composites' PhD thesis (University of Compiegne, France, 1987)
(in French)
66 CantweU, W.J., Davies, P., Jar, P.-Y., Bourban, P.-E. and
Kau~h, H.H. 'Joining and repair of carbon fibre PEEK composites' in Plastics-Metals-Ceramics edited by H.L. Horufeld
(SAMPE, 1990)pp 411--426
67 Evans, R.E., Masters, J.E. and Courter, J.L. 'Toughened
interleafed composites' in Advanced Composites (American
Society for Metals, 1985) pp 249--257
68 Griffin, C.F. 'Damage tolerance of toughened resin graphite
composites' in Toughened Composites op.cit, pp 23-33
69 Lchmmm, S., Megerdiglan, C. and Papalla, R. 'Carbon fiber/
resin matrix interphase: effect of carbon fiber surface treatment
on composite performance' SAMPE Quarterly 16 No 3 (1985)
pp 7-13
70 Drzal, L.T. and Rich, M.J. 'Effect of graphite fiber/epoxy matrix
adhesion on composite fracture behavior' in Recent Advances in
Composites in the United States and Japan op. cir. pp 16--26
71 Rogers, K.F., Sidey, G.R. and Kinstan-Lee, D.M. 'Ballistic
impact resistance of carbon-fibre laminates' Composites 2 (1971)
pp 237-241
72 Dorey, G. 'Relationship between impact resistance and fracture
toughness in advanced composite materials' in Effect of Service
Environment on Composite Materials, A GARD CP 288 (1980)
73 Bless, S.J. and Hartman, D.R. 'Ballistic penetration of S-2 glass
laminates' Proc 2lst Int SAMPE Con[, Sept 1989 pp 852-866
74 McGarry, F.J., Manden, J.F. and Knwamoto, J. 'Impact resistance of rubber modified carbon fiber composite' in Advanced
Composites op.cit, pp 195-205
75 Dorey, G. 'Damage tolerance and damage assessment in
advanced composites' in Advanced Composites edited by I.K.
Partridge (Elsevier Applied Science Publ, 1989) chapter 11
76 Dorey, G. 'Fracture bebaviour and residual strength of carbon
fibre composites subjected to impact loads' in Failure Modes of
Composite Materials with Organic Matrices and Their Consequences in Design, AGARD CP 163 (1975) paper 8
77 Ikewicz, L.B., i)est, E.F. and Coggeshall, R.L. 'A model for
compression after impact strength evaluation' Proc 21st lnt
SAMPE Tech Con[, Sept 1989pp 130-140
78 Kant, C.Y. and Walker, J.V. 'Toughened composites selection
criteria' in Toughened Composites op.cit, pp 9-22
79 Demuts, E. and Sharpe, P. 'Tougher advanced composite
structures' 28th Structures, Structural Dynamics and Materials
Con[, Monterey, CA, USA, April I987 (AIAAIASMEIASCEI
AHS) pp 385-393
80 Stevanovle, M., Kostic, M., Stecenko, T. and Briski, D. 'Impact
behaviour of CFRP composites of different stacking geometry'
in Composites Evaluation, Proc TEQ C 87 op.cit, pp 78-83
81 Dorey, G. 'Fracture of composites and damage tolerance' in
Practical Considerations of Design, Fabrication and Tests for
Composite Materials, A GARD Lecture Series 124 (1982)
82 Nolet, S.C. and Sandusky, P.M. 'Impact resistant hybrid composite for aircraft leading edges' SAMPE Quarterly 17 No 4
(1986) pp 46-53
83 CantweH, W.J. and Morton, J. 'Geometrical effects in the low
velocity impact response of CFRP' Composite Struct 12 (1989)
pp 35-59
84 Liu, D. 'Impact-induced delamination - - a view of bending
stiffness mismatching' J Composite Mater 22 (1988) pp 674--692
85 Vednla, M. and Keczak, M.J. 'Impact resistance of cross-plied
polyphenylene sulfide composites' J Thermoplastic Composite
Mater2 (1989) pp 154-163
86 Malllek, P.K. and Broutman, L.J. 'Static and impact properties
of laminated hybrid composites' J Testing and Evaluation 5
(1977) pp 190-200
87 Hancex, N.L. and Wells, H. 'Izod impact properties of carbon
fibre/glass fibre sandwich structures' Composites 4 (1973)
pp 26-29
88 Adams, D.F. and Perry, J.L. 'Static and impact behaviour of
graphite/epoxy composite laminates containing third-phase
reinforcement materials' J Testing and Evaluation $ (1977)
pp 114-123

362

COMPOSITES. SEPTEMBER 1991

89 Ko, F.K. and Hartnmn, D. 'Impact behavior of 2-D and 3-D


glass/epoxycomposites' SAMPE Journal 22 No 4 (1986) pp 263O
90 Herrlek, J.W. 'Impact resistant multidimensional composites'
Proc 12th Nat SAMPE Tech Con[, Oct 1980 pp 845-856
91 Madan, R.C. ' G - - a measure of damage tolerance of composites' Proc 20th lnt SAMPE Tech Con[, Sept 1988 pp 403--412
92 Dow, M.B, and Smith, D.L. 'Damage-tolerant composite
materials produced by stitching carbon fabrics' 21st Int SAMPE
Tech Con[, Sept 1989
93 Su, K.B. 'Delamination resistance of stitched thermoplastic
matrix composite laminates' in Advances in Thermoplastic
Matrix Composite Materials, ASTM STP 1044 edited by G.M.
Newaz (American Society for Testing and Materials, 1989)
pp 279--300
93 Cantwell, W.J. 'The influence of target geometry on the high
velocity impact response of CFRP' Composite Struct 10 (1988)
pp 247-265
94 Ganse, L.W., Resenfeld, M.S. and Vining, R.E. 'Effect of
impact damage on the XFV-12A composite wing box' Proc 25th
Nat SAMPE Conf, 1980 pp 679--690
95 Madan, R.C. and Sutton, J.O. 'Design, testing and damage
tolerance study of bonded stiffened composite wing cover
panels' 29th Structures, Structural Dynamics and Materials Con[,
Williamsburg, VA, USA, April 1988 (AIAA/ASME/ASCE/
AHS) pp 623--630
96 Welsh, L.M. and Harding, J. 'Effect of strain-rate on the tensile
failure of woven-reinforced polyester resin composites' Report
OUEL 1578/85 (Univ of Oxford, UK, 1985)
97 Daniel, I.M., Yaniv, G. and Anser, J.W. 'Rate effects on delamination fracture toughness of graphite/epoxy composites' in
Composite Structures, Vol2 op. cit. pp 2.258-2.272
98 Karger-Kocsis, J. and Frledrkh, K. 'Temperature and strainrate effects on the fracture toughness of PEEK and its short
glass-fibre reinforced composite' Polymer27 (1986) pp 17531760
99 Mall, S., Law, G.E. and Katea~i=n, M. 'Loading rate effect on
interlaminar fracture toughness of a thermoplastic composite' J
Composite Mater 21 (1987) pp 569-579
100 Starnes, J.H., Rhodes, M.D. and Williams, J.G. 'Effect of
impact damage and holes on the compressive strength of a
graphite/epoxy laminate' in Nondestructive Evaluation and Flaw
Criticalityfor Composite Materials, ASTM STP 696 edited by
R.B. Pipes (American Society for Testing and Materials, 1979)
pp 145-171
I01 Sarma Avva, V. and Padmanabha, H.L. 'Compressive residual
strength prediction in fiber-reinforced laminated composites
subjected to impact loads' in Advances in Fracture Research,
Proc 6th lnt Conf on Fracture edited by S.R. Valluri, D.M.R.
Taplin, P. Rama Rao, J.F. Knott and R. Dubey (1984) pp 28972907
102 Barbezat, M. 'The influence of loading rate on the behaviour of
epoxy composites' PhD thesis (Ecole Polytechnique F~d~rale de
Lausanne, Switzerland, 1990) (in French)
103 Smihy, A.J. 'Rate sensitivity of interlaminar fracture toughness
in composite materials' M.S. thesis (Univ of Delaware, USA.
19S5)
104 Harris, B., Beaumont, P.W.R. and Monennin de Ferran, E.
'Strength and fracture toughness of carbon fibre polyester
composites' J Mater Sci6 (1971) pp 238-251
105 Bal~lyopadhyay, S., Gellert, E.P., SHva, V.M. and Underweod,
J.H. 'Microscopic aspects of failure and fracture toughness in
advanced composite materials' J Composite Mater 23 (1989)
pp 1216-1231
106 Kirk, J.N., Mnnro, M. and Beaumont, P.W.R. 'The fracture
energy of hybrid carbon fibre and glass fibre composites' J Mater
Sci 13 (1978) pp 2197-2204

AUTHORS
W . J . C a n t w e l l , w h o is with t h e L a b o r a t o i r e d e P o l y m e r e s , E c o l e P o l y t e c h n i c F t d t r a l e d e L a u s a n n e , 1007
L a u s a n n e , S w i t z e r l a n d , is c u r r e n t l y a visiting scientist
at V i r g i n i a P o l y t e c h n i c I n s t i t u t e a n d S t a t e U n i v e r s i t y .
J. M o r t o n is w i t h t h e D e p a r t m e n t o f E n g i n e e r i n g
Science and Mechanics, Virginia Polytechnic Institute
a n d S t a t e U n i v e r s i t y , B l a c k s b u r g , V A 24061--0219,
USA.

Vous aimerez peut-être aussi