47 Impact Resistance To Polypropylene Prepegs

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

COMPOSITES

SCIENCE AND
TECHNOLOGY
Composites Science and Technology 66 (2006) 1682–1693
www.elsevier.com/locate/compscitech

The impact resistance of polypropylene-based fibre–metal laminates


M.R. Abdullah, W.J. Cantwell *

Department of Engineering, The University of Liverpool, Brownlow Hill, Liverpool L69 3GH, UK

Received 14 November 2005; accepted 14 November 2005


Available online 28 December 2005

Abstract

The high velocity impact response of a range of polypropylene-based fibre–metal laminate (FML) structures has been investigated.
Initial tests were conducted on simple FML sandwich structures based on 2024-O and 2024-T3 aluminium alloy skins and a polypro-
pylene fibre reinforced polypropylene (PP/PP) composite core. Here, it was shown that laminates based on the stronger 2024-T3 alloy
offered a superior perforation resistance to those based on the 2024-O system. Tests were also conducted on multi-layered materials in
which the composite plies were dispersed between more than two aluminium sheets. For a given target thickness, the multi-layered lam-
inates offered a superior perforation resistance to the sandwich laminates. The perforation resistances of the various laminates investi-
gated here were compared by determining the specific perforation energy (s.p.e.) of each system. Here, the sandwich FMLs based on the
low density PP/PP core out-performed the multi-layer systems, offering s.p.e.Õs roughly double that exhibited by a similar Kevlar-based
laminate.
A closer examination of the panels highlighted a number of failure mechanisms such as ductile tearing, delamination and fibre failure
in the composite plies as well as permanent plastic deformation, thinning and shear fracture in the metal layers. Finally, the perforation
threshold of all of the FML structures was predicted using the Reid–Wen perforation model. Here, it was found that the predictions
offered by this simple model were in good agreement with the experimental data.
Ó 2005 Elsevier Ltd. All rights reserved.

Keywords: A. Layered structures; B. Impact behaviour; C. Delamination

1. Introduction that many composites offer a relatively poor resistance to


penetration and perforation, failure processes that are fre-
Lightweight composite materials are currently finding quently encountered during impact loading [8–10]. In an
extensive use in a wide range of load-bearing engineering attempt to overcome these limitations, a number of
applications. Previous work has shown that fibre-rein- researchers have investigated the possibility of developing
forced polymers such as carbon fibre reinforced plastic hybrid structures based on, for example, mixtures of glass,
(CFRP) offer outstanding strength and stiffness properties Kevlar and carbon fibres [9,10]. Another interesting possi-
as well as a superior resistance to long-term fatigue cycling bility is to develop metal-composite systems such as fibre–
[1–3]. One of the frequently cited limitations of fibre-rein- metal laminates (FMLs) based on thin layers of metal and
forced plastics is their relatively poor resistance to localised fibre-reinforced composite material. At present, systems
impact loading [4–6]. Indeed, low-velocity impact tests such as GLARE (glass fibre/aluminium), ARALL (aramid
have shown that CFRP suffers a significant reduction in fibre/aluminium) and CALL (carbon fibre/aluminium) are
its load-bearing properties following impact at energies as attracting interest from a wide range of engineering sectors.
low as 2 J [7]. In addition, a number of workers have shown The mechanical properties of epoxy-based fibre–metal lam-
inates have been investigated in a number of studies [11–
13]. Alderliesten [12] conducted a series of fatigue tests
*
Corresponding author. Tel.: +44 151 794 5370; fax: +44 151 794 4675. on GLARE and plain aluminium alloy specimens and
E-mail address: W.Cantwell@liverpool.ac.uk (W.J. Cantwell). found that crack growth rates in fibre–metal systems were

0266-3538/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compscitech.2005.11.008
M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693 1683

between one and two orders of magnitude lower than in interface. This low-viscosity polymer interlayer offers excel-
aluminium alloy samples. Vlot et al. [14] conducted impact lent flow characteristic and improved interlocking with the
tests on GLARE, a monolithic aluminium alloy and car- surface of the metal. Table 1 gives a summary of the mate-
bon fibre reinforced PEEK. Their results indicated that rials investigated in this research study. The fibre–metal
the damage threshold energy for this multi-layered material laminates were manufactured by stacking the appropriate
was significantly greater than values offered by more tradi- number of composite plies, aluminium sheets and inter-
tional engineering materials. layer films in a picture-frame mould and placing the stack
The first generation of thermosetting-based fibre–metal in a press. The laminates were then heated to 165 °C at a
laminates suffer a number of key limitations including long rate of approximately 10 °C/min before being cooled to
processing cycles, low interlaminar fracture toughness room temperature at a rate of approximately 5 °C/min.
properties as well as difficulties associated with repair. In Once the press had cooled to a temperature below 60 °C,
an attempt to overcome many of these problems, a number the panel was removed from the mould and visually
of novel FMLs based on thermoplastic matrices have been inspected for defects. A schematic illustration of a 3/2
developed and tested [13,15]. Thermoplastic-based fibre– fibre–metal laminate is shown in Fig. 1. Details of the
metal laminates offer a number of advantages including lay-ups and stacking configurations studied in this research
very short processing times, ease of forming, improved programme are given in Table 2.
chemical resistance, excellent repairability and superior A nitrogen gas gun was used to evaluate the high veloc-
interlaminar fracture toughness properties. Extensive test- ity impact response of the FMLs. Fig. 2 shows a schematic
ing on a glass fibre-reinforced polypropylene FML has representation of the gas gun test facility used in this study.
shown that this system offers an excellent resistance to Square plates with an edge length of 100 mm were clamped
low and high velocity impact loading conditions [15]. A in a steel support with a 75 mm square aperture. The rig
subsequent study suggested that FMLs based on a poly- was then bolted to a steel block and a velocimeter was
propylene fibre reinforced polypropylene matrix offer an placed between the end of the barrel and the target to mea-
excellent resistance to localised high velocity impact load- sure the velocity of the projectile (Fig. 3a and b). Here, the
ing [16]. In addition, self-reinforced polypropylene systems velocimeter ensured a minimum accuracy in measuring the
combine the versatility and recyclability of thermoplastics impact velocity of 99.6%. Impact testing was conducted
with the superior mechanical properties offered by fibre- using a 46.7 g steel projectile with a 12.7 mm diameter
reinforced composites. hemispherical head (Fig. 3c). The velocity of the projectile
The aim of this research project is to investigate the was controlled by adjusting the pressure in the main cham-
impact resistance of a lightweight fibre–metal laminate ber of the gas gun shown in Fig. 2.
based on thin layers of an aluminium alloy and a self-rein- Impact testing was conducted over a range of impact
forced polypropylene composite. Particular attention cen- energies until complete perforation of the FML target
tres on investigating the influence of varying the stacking was achieved. The resulting perforation energy was then
sequence on the perforation resistance of these hybrid used to calculate the specific perforation energy of the
laminates. target by normalising the measured perforation energy by
the areal density of the target. In order to achieve a greater
2. Experimental procedure understanding of the impact behaviour of the FMLs, a ser-
ies of impact tests were carried out on the plain aluminium
The fibre–metal laminates investigated in this study were alloy and the PP/PP composite. After testing, the speci-
based on a woven polypropylene fibre reinforced polypro- mens were removed from the gas gun, sectioned, polished
pylene (PP/PP) (Curv from Propex Fabric) and two types and then viewed under an optical microscope in order to
of aluminium alloy, a 0.6 mm thick 2024-O (annealed) elucidate the failure mechanisms during impact. The spec-
alloy and a 0.8 mm 2024-T3 (solution heat treated, cold imens were sectioned through the point of impact using a
worked and naturally aged) alloy. Before laminating, the band-saw and ground using 1200 grit silicon carbide paper.
aluminium and the PP/PP sheets were cleaned with acetone Following impact testing, the maximum out-of-plane
in order to remove any dust, oil or grease. Optimum adhe- displacement in the impacted samples was measured in
sion across the composite metal interface was ensured by order to characterise the impact response of the FML spec-
placing a 60 lm thick modified polypropylene interlayer imens. Here, tangents were drawn to the deformed portions
(Gluco from Gluco Ltd., Leeds, UK) at each common of the specimen, and the displacement, d, was measured

Table 1
Summary of the materials investigated in this research programme
Material Grade Supplier Density (kg/m3)
Aluminium alloy 2024-T3 (0.8 mm thick) Alcan 2770
2024-O (0.6 mm thick) Alcan 2770
Curv Woven polypropylene fibre/polypropylene matrix Propex Fabric 920
Gluco Polypropylene adhesive film 55 g/m2 Gluco Ltd.
1684 M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693

Fig. 1. Schematic illustration of the manufacturing procedure for a 3/2, polypropylene fibre reinforced fibre–metal laminate.

Table 2 from the point of intersection of these tangents to the ini-


Details of the lay-ups and stacking configurations investigated in this tial, undeformed shape, as shown in Fig. 4. Although this
study
method is somewhat crude, it does yield useful information
Lay-up Stacking configuration Average Areal on the energy-absorbing capacity of the specimens.
thickness density
(mm) (kg/m2)
3. The perforation model
Al 2024-O and PP/PP composite
O2/1 Al/PP/Al 1.75 4.06
O3/2 Al/PP/Al/PP/Al 2.85 6.28 The perforation threshold energy for the FMLs was pre-
O4/3 Al/PP/Al/PP/Al/PP/Al 3.88 8.48 dicted using the perforation model developed by Reid and
O5/4 Al/PP/Al/PP/Al/PP/Al/PP/Al 4.89 10.69 Wen [17]. Here, it is assumed that when a laminate is per-
Al 2024-T3 and PP/PP composite forated by a projectile at high velocities, the mean pressure
T32/1 Al/PP/Al 2.06 4.97 (r) applied to the projectile by the target can be separated
T33/2 Al/PP/Al/PP/Al 3.28 7.62 into two components: one associated with the cohesive sta-
T34/3 Al/PP/Al/PP/Al/PP/Al 4.53 10.31 tic resistive pressure due to elastic-plastic deformations
T35/4 Al/PP/Al/PP/Al/PP/Al/PP/Al 5.73 12.96
within the laminate (rs), and a second associated with the

Fig. 2. Schematic illustration of the test arrangement for high velocity impact testing.
M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693 1685

Fig. 3. (a) Schematic of the specimen clamp in plan view, (b) specimen clamp in side view, and (c) hemispherical-ended projectile.

pD2 T r
Ep ¼ ; ð4Þ
4
where T is the thickness of the laminate and D is the diam-
eter of the projectile.
Substituting Eq. (3) into Eq. (4) gives:
 rffiffiffiffiffi 
pD2 T qt
Ep ¼ 1þC V b re . ð5Þ
4 re
Knowing that Ep ¼ 12mV 2b , the perforation velocity can be
written as [17]
Fig. 4. Method employed for measuring the permanent displacement of pffiffiffiffiffiffiffiffiffi " sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi#
the target. pC qt re D2 T 8m
Vb ¼ 1þ 1þ 2 . ð6Þ
4m pC qt D2 T
dynamic resistive pressure (rd) associated with loading rate
This equation was used to predict the perforation threshold
effects [17]. This can be written as
of the FMLs examined in this study.
r ¼ rs þ rd . ð1Þ
It is also assumed that the cohesive static resistive pressure 4. Results and discussion
is equal to the static linear elastic compression limit (re) in
the through-thickness direction, i.e., rs = re, and that the 4.1. Density of the FMLs
dynamic resistive pressure (rd) can be written as
1=2 The density of a fibre–metal laminate is clearly related to
rd ¼ Cðqt =re Þ V i re ; ð2Þ
the densities and relative proportions of its constituent
where qt and Vi represent the density of the laminate and materials, the aluminium alloy and the self-reinforced poly-
the initial projectile velocity, respectively. The term C is a propylene in the present case. Prior to testing, the densities
constant having a value of 1.5 for a hemispherical projectile of all of the laminates investigated in this study were deter-
[17]. mined by measuring and weighing the as-manufactured
Eq. (1) can therefore be written in the following form: panels. Fig. 5 shows the variation of density of the 2/1
 rffiffiffiffiffi  fibre–metal laminates examined in this study as a function
qt
r¼ 1þC V i re . ð3Þ of laminate thickness. The figure shows that increasing the
re
laminate thickness by increasing the thickness of the com-
Reid and Wen showed that the perforation energy of a posite core has the effect of reducing the structures density.
composite laminate when struck transversely by a rigid It is interesting to note that the densities of laminates based
projectile at high velocity is given as [17] on the Al 2024-O alloy were slightly lower than those based
1686 M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693

are up to 55% lower than those that associated with the


plain aluminium alloy. The densities of both types of
FML shown in Fig. 5 are significantly lower than those
reported for CARE (carbon fibre-reinforced epoxy/alumin-
ium), ARALL (aramid fibre-reinforced epoxy/aluminium)
and an aluminium–lithium FML system based on a car-
bon–fibre composite [18].

4.2. Impact failure mechanisms

The failure processes in the impact-loaded FMLs were


initially investigated by examining the front and rear sur-
faces of the impact-damaged samples. Initial fracture in
the 2024-O 2/1 FML following a 33.6 J impact is shown
Fig. 5. The variation of density with laminate thickness for the 2/1 FMLs.
in the low magnification micrographs presented in
Fig. 6a. Here, damage takes the form of a large top surface
on the Al 2024-T3 alloy. This difference is due to the fact dent and a localised crack parallel to the rolling direction
that the two types of aluminium alloy employed in this of the rearmost aluminium layer. Similar failure mecha-
study offered differing thicknesses (the 2024-T3 system nisms have been observed in GLARE laminates tested
was 0.2 mm thicker than the 2024-O alloy). Therefore, under low-velocity impact conditions [19]. With increasing
for a given overall laminate thickness, the 2024-O FML energy, the length of the rear surface crack increased as did
contained a larger volume fraction of the low-density com- the size of the top surface dent. At the perforation thresh-
posite, thereby lowering the density of the structure. The old, the passage of the projectile through the target pro-
solid lines in Fig. 5 correspond to the predictions of a sim- duced a relatively clean hole with a diameter similar to
ple rule of mixtures approach based on the volume frac- that of the steel projectile, Fig. 6b. Clearly, the perforation
tions of composite and metal in the laminates (the process involves significant local ductility in both the alu-
presence of the polypropylene interlayer was ignored). minium and composite plies as well as fracture of the con-
Clearly, the simple rule of mixtures approach predicts the stituent materials. The mechanisms of damage initiation in
densities of the laminates with some accuracy. It is worth the thicker 2024-O 4/3 laminate were similar to those
noting that the FML configurations offer densities that observed in its thinner 2/1 counterpart with initial failure

Fig. 6. Low magnification optical micrographs of the impact-damaged 2024-O 2/1 FMLs: (i) impacted surface and (ii) rear surface.
M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693 1687

occurring in the rear surface layer parallel to the rolling The failure processes in the impacted specimens were
direction of the aluminium, Fig. 7a. At higher energies, a further investigated by sectioning a number of samples
large dent and then a circular crack develop around the transversely through the centre of impact after testing.
impactor and the uppermost composite and metal plies The cross-sections were then polished and examined under
begin to be pushed through the rear surface opening, an optical microscope. The subsequent development of
Fig. 7b. Finally, at the perforation threshold, the projectile damage with increasing impact energy in the 2/1 and 5/4
passes through the target leaving a relatively clean hole and FML configurations is presented in Figs. 9 and 10, respec-
limited petalling at the rear surface. Fig. 8a shows damage tively. In the 2024-O 2/1 laminate at lower impact energies,
in the 5/4 T3 laminate where initial failure again took the the lower surface aluminium fractures and the upper
form of top surface denting coupled with a split in the low- surface ply exhibits localised thinning around the point of
ermost aluminium ply parallel to the rolling direction. At impact, Fig. 9a. It is believed that this thinning process
the perforation threshold, a circular hole was visible in results from membrane stretching and subsequent yielding
the top surface ply and localised petalling of the rear alu- in the thin aluminium plies during the impact process. Fol-
minium layer. Typically, there were three dominant petals lowing an impact energy of 46.2 J, Fig. 9b, fracture of
in these perforated samples with minor ones propagating upper and lower aluminium plies is observed as well as
between them. In the 4/3 and 5/4 laminates, the projectile localised fracture in the composite core of this sandwich
frequently remained embedded in the target at impact ener- structure. Finally, at the perforation threshold, the projec-
gies just below the perforation threshold. A careful exami- tile passes through the target, a process that involves
nation of the impact chamber indicated that the significant plastic deformation in the zone of material
penetration and perforation processes produced very little immediate to the impact location Fig. 9c. Similar failure
debris. In most cases, only a small number of composite processes were observed in the corresponding 2024-T3 lam-
fragments and a dish-shaped plug were observed. inates, Fig. 9d–f although evidence of shear fracture in the

Fig. 7. Low magnification optical micrographs of the impact-damaged 2024-O 4/3 FMLs: (i) impacted surface and (ii) rear surface.
1688 M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693

Fig. 8. Low magnification optical micrographs of the impact-damaged 2024-T3 5/4 FMLs: (i) impacted surface and (ii) rear surface.

Fig. 10. Low magnification optical micrographs of polished sections of 5/


4 FMLs subjected to various impact energies: (i) 2024-O 5/4 and (ii) 2024-
T3 5/4.

aluminium layers was observed at intermediate energies.


The processes of damage initiation and subsequent target
perforation were similar in the 3/2 and 4/3 laminates. Once
again thinning of the aluminium plies was observed around
the point of impact in the 2024-O laminates at intermediate
Fig. 9. Low magnification optical micrographs of polished sections of 2/1 impact energies whereas shear fracture was more pro-
FMLs subjected to various impact energies: (i) 2024-O and (ii) 2024-T3. nounced in the T3 systems. Typical examples of localised
M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693 1689

thinning and shear fracture of the aluminium plies are 4.3. Residual deformation in the impacted FMLs
shown in Fig. 11. A closer inspection showed that delami-
nation and ductile tearing within the composite plies were The impact response of the FMLs was further investi-
more pronounced in the T3 systems than in their 2024-O gated by measuring the residual displacement in the
counterparts. Interestingly, Hagenbeek stated that trans- impact-loaded structures. As expected, the residual perma-
verse shear failure such as that observed in the T3 system nent displacement in the targets increased with increasing
can act as a precursor to delamination in FMLs [19]. Fail- incident impact energy, reaching a plateau as the perfora-
ure in the 5/4 laminates tended to initiate in the upper part tion threshold was approached.
of the target immediate to the point of impact, Fig. 10. Fig. 12 shows plots of the out-of-plane displacement at
Here, the passage of the projectile through the target pre- the perforation threshold for the 2024-T3 and 2024-O.
cipitated failure of the aluminium alloy immediately under Interestingly, the critical out-of-plane deformation, dmax,
the point of impact. As previously observed, localised thin- for the 2024-O laminates is roughly constant for all of
ning of the aluminium was observed in the 2024-O FMLs the laminates, averaging approximately 12 mm regardless
whereas transverse shear fracture was apparent in the T3 of the plate configuration or target thickness. Similarly,
laminates. Once again, delamination and ductile tearing the average of dmax for the 2024-T3 system is also roughly
within the composite layers was more pronounced in the constant, averaging approximately 10 mm over the range
T3 system that in its 2024-O counterpart. of laminates investigated. It is believed that the lower value
of dmax for the T3 FMLs results from the increased thick-
ness of these layers, the lower strain to failure of this alloy
as well as the change in failure mode (thinning to shear)
during impact. The fact that the maximum rear surface
deformation of the targets remains roughly constant as
the thickness of the target is increased suggests that the alu-
minium layers may deform independently of each other
during the perforation process. Here, the low modulus of
the self-reinforced composite allows the aluminium plies
to deform in a quasi-independent manner when loaded in
flexure. As observed previously, the out of plane displace-
ment of these plies can lead to localised membrane stretch-
ing which in turn leads to enhanced energy absorption.

Fig. 11. Optical micrographs showing thinning of the aluminium plies in a


2024-O laminate and shear fracture of the aluminium plies in a 2024-T3 Fig. 12. Maximum permanent out-of-plane displacement at the perfora-
laminate. tion threshold for the 2024- O and 2024-T3 laminates.
1690 M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693

This was investigated further by conducting tests on alu- two laminate types is most pronounced in the thinner lam-
minium samples that were bonded using an unreinforced inates, where the perforation resistance of the 2024-T3
polypropylene adhesive. The perforation energies of these FML is almost double that of its 2024-O counterparts. This
systems were not as impressive as their composite counter- difference reduces as the laminate thickness is increased. It
parts indicating that the composite makes a positive contri- should be noted that the 2024-T3 aluminium is thicker than
bution to the perforation process rather than simply acting that of its 2024-O counterpart and this is likely complicate
as a low modulus adhesive holding the aluminium sheets comparisons between the two systems. However, tensile
together. Indeed, the room temperature Izod fracture tests have shown, however, that the higher strength 2024-
energy of the composite is very high (4750 J/m) [20] sug- T3 alloy absorbs considerably more energy up to fracture
gesting that considerable energy is also absorbed in fractur- (i.e., a larger area under the stress–strain curve) than the
ing these layers during the perforation process. This will be 2024-O alloy and this appears to be translated into a supe-
investigated further in the following section where the vol- rior energy absorbing capacity under impact loading.
ume fraction of composite in the FMLs is varied. Included in Fig. 13 are perforation data resulting from tests
on the plain composite material. Clearly, the impact resis-
4.4. Perforation resistance of the FMLs tances of both types of FML are superior to that offered by
the composite material.
The impact response of the fibre–metal laminates was The solid lines in Fig. 13 represent the predictions of the
initially investigated by determining the perforation thresh- Reid–Wen model. From the figure, it is clear that the
old of each system investigated during the course of this model predicts the perforation resistances of the two types
study. Here, test specimens were subjected to increasing of FML as well as the plain composite with a high degree
impact energies until complete perforation of the target of success. Here, values of the static linear elastic compres-
was achieved. Fig. 13 compares the perforation resistances sion limit (re) of 131, 192 and 247 MPa were obtained for
of the 2/1 FMLs based on the 2024-T3 and 2024-O alloys. the plain composite, the 2024-O and the 2024-T3 systems,
In these laminates, the thickness of the composite core respectively, Table 3. As expected the computed value of
sandwiched between the outer aluminium plies was re was higher for the FMLs than the plain composite
increased from approximately 0.5–6.6 mm. In Fig. 13, the and re was higher in the stronger T3 FMLs than in the
two points on the y-axis correspond to laminates based 2024-O laminates.
on two aluminium plies bonded by a single layer of unrein- Fig. 14a and b compares the perforation resistances of
forced polymer adhesive. Therefore, these points effectively the multi-layered systems. Here, impact tests were under-
correspond to laminates with a composite core thickness taken on 2/1, 3/2, 4/3 and 5/4 laminates as well as on an
equal to zero. Clearly, the perforation resistance of both supplementary 10/9 system based on the 2024-O alloy. As
types of FML increases with increasing target thickness. previously observed, in terms of perforation energy, the
For example, the perforation resistance of the 2024-T3 sys- FMLs based on the stronger aluminium alloy out-perform
tem with a 6.5 mm composite core was more that three their 2024-O counterparts, Fig. 14a. Again, the properties
times greater than that of the thinnest system. It is also evi- of the aluminium alloy have the greater influence (in per-
dent that the 2024-T3 fibre–metal laminates out-perform centage terms) in the thinnest FMLs where the perforation
their 2024-O counterparts. The difference between the resistance of the T3 FML was approximately double that
associated with the corresponding 2024-O system.
Fig. 14b presents the data shown in Fig. 14a normalised
by the thickness of the target. The data again support the
conclusion that the T3 laminates (particularly at low target
thicknesses) outperform their 2024-O counterparts. It is

Table 3
Summary of the static linear elastic compression limit (re) data for the
families of systems examined here
Material system Static linear elastic
compression limit,
re (MPa)
Plain composite 131
2024-O 2/1 laminates 192
2024-T3 2/1 laminates 247
2024-O multi-layer laminates 224
2024-T3 multi-layer laminates 338
Fig. 13. The variation of the perforation energy with thickness of the (0.8 mm plies)
composite core for the 2/1 FMLs. Included in the figure are the 2024-T3 multi-layer laminates 208
perforation thresholds of the plain composite. The solid lines represent (0.3 mm plies)
the predictions of the perforation model.
M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693 1691

Fig. 15. Comparison of the perforation resistances of 2024-T3 laminates


based on thin (0.3 mm) aluminium plies (filled circles) and thick (0.8 mm)
aluminium plies (open circles).

FMLs. Given the fact that, for a given target thickness,


the perforation resistance of the aluminium is superior to
that of the plain composite, it is not surprising that lami-
nates based on thicker metal plies offer a superior impact
resistance. If the data in Fig. 15 are re-plotted as a function
of the thickness of aluminium within the composite, the
two sets of data appear to collapse onto one line suggesting
that ply thickness has a secondary effect on the perforation
characteristics of these laminates. Included in Fig. 15 are
the predictions for the perforation thresholds of the lami-
nates based on 0.3 mm thick plies. Here, it is evident that
Fig. 14. (a) Comparison of the perforation resistances of the multi-layered once again the model successfully predicts the perforation
2024-O and 2024-T3 FMLs. (b) Comparison of the normalised perfora-
resistance of these FMLs over the range of thicknesses con-
tion resistances of the multi-layered 2024-O and 2024-T3 FMLs.
sidered. It is worth noting that the 6.7 mm thick system in
this figure corresponds to a 10/9 laminate, a case where the
interesting to note that the data in Fig. 14b indicate that a model over-estimated the perforation threshold in the
laminate thickness of approximately 5 mm represents the thicker 0.8 mm FMLs, Fig. 14a.
optimum configuration. At values above this threshold, In the final part of this study, the impact resistance of
the normalised perforation energy begins to fall. the self-reinforced FML system was compared to values
The solid lines in Fig. 14a correspond to the predictions determined from high velocity tests on similar materials.
of the Reid–Wen model discussed previously. The resulting One of the difficulties in comparing data from tests on dif-
values of re are presented in Table 3 where it is again clear ferent materials is that the target thicknesses and densities
that the T3 system offers the higher value of re. Once again, vary from system to system. Previous work to compare the
the perforation model predicts the impact resistances of the impact resistance of fibre–metal laminates has involved the
laminates with some success. Clearly, the model over-esti- determination of the specific perforation resistance or what
mates the perforation resistance of the 10/9 system, this is sometimes referred to as an impact energy per unit
may be due to the use of an inappropriate value of re for weight [19]. This type of normalising procedure has been
predicting the perforation resistance of this laminate. It is used to compare the ballistic impact resistance of fragment
possible that the parameter, re, exhibits a thickness depen- barriers and other types of arresting systems [21]. The per-
dency, whereby it decreases with increasing target thick- foration resistances of the various hybrid structures inves-
ness. Further work is needed to fully investigate this. tigated in this research programme were therefore
Fig. 15 compares the perforation resistances of T3 lam- compared by determining the specific perforation energy
inates based on thin (0.3 mm) and thick (0.8 mm) alumin- of each laminate (calculated by normalising the measured
ium plies. The figure clearly shows that for a given target perforation energy by the areal density of the specimen).
thickness, laminates with thicker aluminium plies offer a Fig. 16 presents the specific perforation energies of the
superior impact resistance to those based on thin plies. This FMLs investigated in this research study. Here, the FMLs
results from the fact that, for a given target thickness, the are grouped according to the alloy type and stacking con-
volume fraction of aluminium in the FMLs based on figuration (2/1 or multi-layer laminate). Included in the fig-
0.8 mm thick aluminium is greater than in the 0.3 mm ure is the associated specific perforation energy of the
1692 M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693

Fig. 16. Comparison of the specific perforation energies of the 2024-O and
2024-T3 (based on 0.8 mm thick aluminium plies). The dimensions in
brackets correspond to the thicknesses of the 2/1 laminates.

Fig. 17. Comparison of the specific perforation energy of the 2024-T3 3/2
aluminium alloy used for each family of laminates. The FML with previously published data.
specific perforation energies of the 2/1 laminates based
on the 2024-O alloy increase rapidly as the laminate thick-
ness is increased. The s.p.e. of the thinnest 2/1 laminate is note that the s.p.e. of this self-reinforced system is more
just above that of the plain aluminium alloy. However, as than 250% higher than that measured on a comparable
the thickness of the low-density composite core increases, Kevlar-based FML. When the data are presented in terms
so does the specific perforation energy. The value corre- of perforation energy rather than specific perforation
sponding to the thickest 2/1 laminate represented the high- energy, the self-reinforced systems continue to exhibit an
est value of s.p.e. recorded in this study. Interestingly, the excellent resistance to impact although the improvements
s.p.e. values for the multi-layer 2024-O laminates were less relative to the other systems are somewhat reduced due
impressive. Indeed, the s.p.e. of the 10/9 laminate was to the fact the PP/PP system offers a low areal density.
lower than that of the corresponding 5/4 system. Although As previously stated, it is believed that the low modulus
the incorporation of increased amounts of aluminium composite plies allow the metal layers to deform indepen-
increased the absolute perforation resistances of the FMLs, dently during the impact process facilitating energy absorp-
its relatively high density reduced their specific perforation tion through membrane stretching and gross plastic
resistances. The third grouping of laminates in Fig. 16 deformation. The other laminates in Fig. 17 are based on
shows that the s.p.e. values of T3 2/1 laminates increases stiffer fibres that may restrict this deformation process in
rapidly with increasing core thickness. It is interesting to the FMLs, thereby reducing their perforation resistance.
note the performance of the thickest of these laminates The evidence in Fig. 17 suggests that FMLs based on com-
was similar to that exhibited by the corresponding O lam- binations of thin layers of aluminium and self-reinforced
inates, a fact that emphasises the predominating contribu- polypropylene offer significant potential for use in the
tion of the thick composite cores in these laminates. In design of impact resistant structures.
contrast, the multi-layer T3 laminates out-perform their
O counterparts, these being laminates in which the volume 5. Conclusions
fraction of aluminium was high in all cases. Once again,
increasing the thickness of these multi-layer systems even- The high velocity impact performance of a range of
tually results in a reduction in the specific impact perfor- fibre–metal laminates based on a self-reinforced polypro-
mance of these systems. The data in Fig. 16 show that pylene has been studied. It has been shown that these
the specific perforation energies of the thicker FMLs hybrid laminate systems offer potential for use in light-
greatly exceed the values associated with the aluminium weight energy-absorbing structures. Impact tests have
alloy. shown that multi-layer FMLs based on the stronger
Fig. 17 compares the specific perforation energy of the 2024-T3 alloy offer a superior perforation resistance to
2024-T3 3/2 FML with previously published data following those based on a 2024-O alloy. The superior perforation
impact tests on a range of woven thermoplastic-matrix resistance of the 2024-T3 system over the 2024-O FMLs
composites (glass fibre reinforced polypropylene, glass/ is likely to be due to the superior energy-absorbing behav-
PEI, a glass fibre PA6,6, and a Kevlar PA 6,6 composite) iour of the aluminium alloy in the former. A detailed exam-
and a thermosetting-matrix laminate (a woven glass fibre/ ination of the perforation zone indicated that significant
phenolic) [16]. The superiority of the self-reinforcing sys- thinning of the aluminium plies had occurred during the
tem examined here is clearly evident. It is interesting to perforation process. It is postulated that the low modulus
M.R. Abdullah, W.J. Cantwell / Composites Science and Technology 66 (2006) 1682–1693 1693

composite plies in the FMLs allow the aluminium layers to [4] Prichard JC, Hogg PJ. The role of impact damage in post-impact
deform independently, absorbing significant energy compression testing. Composites 1990;21:503–11.
[5] Dorey G, Bishop SM, Curtis PT. On the impact performance of
through localised membrane stretching. In addition, the carbon–fibre laminates with epoxy and PEEK matrices. Compos Sci
high strain to failure of the polypropylene fibres with the Technol 1985;23:221–37.
composite allow large amounts of energy to be absorbed [6] Zhou G. Prediction of impact damage thresholds of glass–fiber-
in the failure process thereby enhancing the perforation reinforced laminates. Compos Struct 1995;31:185–93.
resistance of these layered structures. [7] Cantwell WJ, Curtis P, Morton J. An assessment of the impact
performance of CFRP reinforced with high strain carbon fibres.
It was observed that the highest specific perforation Compos Sci Technol 1986;25:133–48.
energy was offered by a simple sandwich construction [8] Thanomsilp C, Hogg PJ. Penetration impact resistance of hybrid
based on a thick composite core and thin outer aluminium composites based on commingled yarn fabrics. Compos Sci Technol
plies. An examination of the failed laminates showed that 2003;63:467–82.
membrane stretching, plastic deformation and shear frac- [9] Abrate S. Impact on laminated composite materials. Appl Mech Rev
1991;44:155–90.
ture in the aluminium layers as well as plastic drawing, [10] Dorey G, Sidey SGR, Hutchings J. Properties of carbon fiber-Kevlar
delamination and ductile tearing in the composite plies 49 fiber hybrid composites. Composites 1978;9:25–32.
were the primary energy-absorbing mechanisms in these [11] Krishnakumar S. The synthesis of metals and composites. Mater
laminates. The perforation resistance of these polypropyl- Manuf Process 1994;9:295–354.
ene fibre-based FML structures can be predicted using a [12] Alderliesten RC. Fatigure. In: Vlot A, Gunnick JW, editors. Fibre–
metal laminates: an introduction. Dordrecht: Kluwer Academic
simple model previously used to predict the impact Publishers; 2001 [chapter 11].
response of composite materials. [13] Reyes-Villanueva G, PhD thesis, University of Liverpool, 2002.
[14] Vlot AD, Kroon E, La Rocca G. Impact response of fiber metal
Acknowledgements laminates. Key Eng Mater 1998;141–143:235–76.
[15] Reyes-Villanueva G, Cantwell WJ. The mechanical properties of
The authors acknowledge the financial support of the fibre–metal laminates based on glass fibre reinforced PP. Compos Sci
Technol 2000;60:1085–94.
Public Service Department (Malaysia) and Universiti Tek- [16] Cantwell WJ, Wade G, Guillen GF, Reyes-Villanueva G, Jones N,
nologi Malaysia. The authors are also grateful to Derek Ri- Compston P. The high velocity impact response of novel fiber–metal
ley of Propex Fabric for supplying the self-reinforced laminates. In: Proceedings of the ASME conference, New York; 2001.
composite (Curv) and David Robinson and Professor Tony [17] Reid SR, Wen HM. Perforation of FRP laminates and sandwich
Johnson of Gluco Ltd., Leeds, UK for supplying the inter- panels subjected to missile impact. In: Reid SR, Zhou G, editors.
Impact behaviour of fibre-reinforced composite materials and struc-
layer adhesive (Gluco). The authors are also grateful to tures. Cambridge: Woodhead Publishing; 2000. p. 239–79.
Mr. Peter Smith for his help in conducting the impact tests. [18] Afaghi-Khatabi A, Ye L, Mai YW. Hybrids and sandwiches. In:
Kelly A, Zweben C, editors. Comprehensive composite materials.
References Amsterdam: Elsevier; 2000. p. 249–90.
[19] Hagenbeek M. Impact properties. In: Vlot A, Gunnick JW, editors.
[1] Peters ST, editor. Handbook of composites. London: Chapman & Fibre–metal laminates: an introduction. Dordrecht: Kluwer Aca-
Hall; 1998. demic Publishers; 2001 [chapter 27].
[2] Matthews FL, Rawlings RD. Composites structures and materials. [20] Available from: www.curvonline.com .
Boca Raton (FL): CRC Press; 1999. [21] Shockey DA, Erlich DC, Simmons JW. Full-scale tests of lightweight
[3] Kelly A, editor. Concise encyclopaedia of composite materials fragment barriers on commercial aircraft, final report DOT/FAA/
(revised). Amsterdam: Elsevier; 1994. AR-99/71; 1999.

You might also like