Sie sind auf Seite 1von 25

International Journal of Impact Engineering 22 (1999) 485—509

Impact damage with compressive preload and post-impact


compression of carbon composite plates
X. Zhang *, G.A.O. Davies, D. Hitchings
Carnfield College of Aeronautics, Cranfield University, Bedford, MK43 0AL, UK
Department of Aeronautics, Imperial College of Science, Technology & Medicine, London, SW7 2BY, UK
Received 1 May 1998; received in revised form 18 December 1998

Abstract

This paper examines the effect on laminated composites of in-plane compression followed by impact
damage, and the coupling between the two, on compression-after-impact (CAI) performance. It is found that
preload can actually raise the CAI strength if the load approaches the initial buckling value, since the plate
loses stiffness and the impact-induced force is reduced and so is the consequent damage. However, as the
preload approaches the CAI strength the induced delamination can propagate catastrophically during the
impact, but at a preload value below the CAI strength. The coupling between impact and preload is
simulated using the equations of motion, and the same dynamic solver is then used to capture the
snap-through when a plate is loaded beyond its initial buckling load. A special ‘dynamic relaxation’ routine is
shown to be robust and reliable when handling these ‘snap-through’ problems which can be formidable
challenges to conventional static incremental loading.  1999 Elsevier Science Ltd. All rights reserved.

Keywords: Delamination; Preload effect; Impact test; Finite element simulation

1. Introduction

In aerospace industry, the topic of impact damage in laminated composite materials continues to
be a vigorous activity, ranging from the low-velocity impact (caused by drop of tools or runway
stones) to the ballistic damage caused by missiles or shell fragments. This paper deals with the
low-velocity impact damage of carbon/epoxy laminated composite structures. There are three
distinct but related parts to this investigation: (i) prediction of damage caused by low-velocity
impact on a wide range of coupons, plates, and realistic stiffened panels; (ii) analysing the effect of

* Corresponding author.

0734-743X/99/$ - see front matter  1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 7 3 4 - 7 4 3 X ( 9 9 ) 0 0 0 0 3 - 2
486 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

in-plane compressive preload on impact damage; (iii) assessing the residual compressive strength of
a panel containing known initial delamination damage caused by low-velocity impact.
The first topic has been reported in the authors’ previous work [1,2], in which both experimental
tests and numerical simulation were carried out over a wide range of small coupons, large plates
and three forms of stiffened panels (i.e. with ¹, I, and top-hat section of stiffeners, respectively). This
wide range of structures, all having different dynamic responses, was chosen to demonstrate that
the maximum impact force, rather than incident energy, could be used as a damage diagnostic over
the complete range of structures. The finite element code used was FE77 which has been developed
at the Aeronautics Department of Imperial College [3]. The main conclusions were: (a) the
maximum impact force is the driving factor for the on-set of delamination; (b) it is necessary to
include in-plane degradation in the FE model to mimic the failure of fibres and matrix during
impact in order to predict the impact force histories accurately. Both the in-plane degradation and
interlaminar delamination can now be predicted [4]. This present paper is thus aimed at the second
and third tasks, i.e. (ii) and (iii) above.
A major weakness of carbon fibre laminated composites is that low-velocity impact, introduced
accidentally during manufacture, operation or maintenance of the aircraft, may result in delamina-
tions between the plies. When a structure is under tensile load the effect of internal delamination on
the residual strength is small. It is in compression that delamination is the worst destabiliser, since
it may become unstable and propagate in direction at right angles to the applied compressive
stress. In the aircraft industry, compression panels are widely used, such as the upper skin panels of
wings. Under in-plane compressive loads the through-thickness forces will develop if the plate is
loaded beyond the initial buckling load and postbuckling rotations increase. These consequent
force components do not have to be large since the through-thickness strengths in shear or tension
may be less than 1/20 of the in-plane strengths. In modern aircraft for example it is not desirable to
allow buckling up to the proof loads (acceptable deformation) but for the design ultimate
(1.5;Proof ) the extra postbuckling strength may be exploited. The presence of impact damage will
of course lower the failure load. For an efficient design of composite aircraft structures, it is
important to be able to simulate the low-velocity impact damage with compressive preload and the
post-impact failure in compression by a numerical model, thus to limit the number of tests at all
structural levels.
It is also known that compression-after-impact and compression-during-impact may not lead to
the same failure load [5], and an explanation is needed, although the number of situations where
a realistic aircraft structure would be loaded to near its ultimate and then receive an impact is
small. (Runway debris hitting a flap or other control surface is one such case.) The problem may
also be described as: the effect of impact damage during service is known to lower the failure loads,
but it is also possible that a preloaded panel will fail during impact at a lower applied compressive
load than that which would cause failure when applied after impact. To examine this possibility,
the investigation has been designed in the following two parts.
Firstly, the problem when a component is under an existing compressive stress at the instant of
impact was investigated. Some work on this problem has been published, such as Refs. [5—7]. In the
work of Tweed et al. [5], low-velocity impact test was carried out on small coupons (100;25 mm)
under either tensile or compressive preloads. The results showed the increasing delamination area
with increasing pre-strain. Since small and narrow coupons were used in this research, it does have
some limitations due to the sensitivity to specimen geometry and end constraint. It is difficult to
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 487

relate the results to realistic structural components. In the work of Sun and Chen [6], it was found
that an initial compressive stress gives rise to a softening effect on the laminate stiffness. However,
the applied compressive preload was well below the critical buckling load, i.e. P"0.75P . Since
 
some aerospace structures are operating near their critical buckling load or even in the post-
buckling region, there is a need to study the effect of a much higher preload. Another study was
carried out by Starnes et al. [7], in which both simultaneous preload and impact and sequential
impact and post-impact compression tests on some T300 CFRP laminates were carried out.
However, the impact velocities were in the range 52—101 m/s which makes it difficult to apply to the
low-velocity impacts. It is known [2] that the transition from failure due to global flexural response
to local through-thickness stress waves occurs at roughly impact velocities of order 20 m/s in
carbon-epoxy composites. Therefore, it was necessary to conduct our own tests representing
low-velocity impact and realistic preload levels, and using structurally relevant plates and panels.
Most importantly, this study aims to simulate the damage process by the finite element method.
A robust dynamic relaxation technique has been developed for this purpose.
Secondly, the usual aerospace compression-after-impact (CAI) tests were conducted to assess the
effect of impact damage on the compressive residual strength. This is a sequential approach of
introducing impact damage first and then applying compression stress to the component until its
final failure. Much work has been carried out in this area, such as [7—9]. These are mainly
experimental investigations and the conclusion has been that low-velocity impact damage can
seriously degrade the laminate compressive strength. In those cases where delamination occurred
near the plate surface, it was found that local ‘blister’ buckling can be formed but it then stabilises
until enough potential energy is available to propagate the delamination further and cause
buckling of the two separate plates. Sometimes if the damage in the plate interior is extensive the
failure in compression is explosive and the order of events is difficult to monitor experimentally. An
analytical programme to simulate the CAI tests is needed in order to understand the mechanism of
delamination propagation. It is currently in progress. Therefore only the experimental results of the
CAI test are reported in this paper.

2. Materials and experimental methods

2.1. Materials and specimens

The material used is T800/924 carbon/epoxy laminated composite. The plates tested have the
quasi-isotropic lay-up of [#45°/!45°/90°/0] , where n"2 or 4 giving the nominal plate
LQ
thickness of 2 or 4 mm, respectively. The unidirectional properties are given in Table 1, and for this
stacking sequence result in an equivalent flexural modulus of 70 GPa.
The dimension of the plates was 270;250 mm which has now become (almost) a European
standard since it allows for a reasonable and realistic amount of impact damage and still has
a reasonable “by-pass” load path which is more representative of realistic full-scale structures. Nine
pairs of strain gauges were arranged across the plate in the in-plane loading direction in order to
measure the compressive strains and to monitor the plate’s buckling and post-buckling behaviour.
Three linear voltage displacement transducers (LVDTs) were also employed to record the plate’s
lateral displacements. The pattern of nine gauges and displacement transducers is sufficient to
488 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Table 1
Unidirectional properties of T800/924

Moduli (GPa) Poisson’s ratio Strengths (MPa)

E E G t p p q q "q
        
168 9.5 5.3 0.3 2700 75 234 85

Fig. 1. Plate specimen and strain-gauge positions.

indicate the mode shape, and the use of pairs on each surface often demonstrates that a local blister
has developed. A test plate is illustrated in Fig. 1.

2.2. Instrumented falling-weight impact rig with preloading facility

A test rig for applying compressive preload was designed and made. The frame and loading jack
were assembled in a horizontal plane under an existing impact rig so that it was possible for the
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 489

Fig. 2. Experimental set-up for drop-weight impact with compressive preload test (a) impactor, (b) load cell, (c) drop
guide, (d) flag, (e) photodiodes, (f) clamping device.

plate to be preloaded and simultaneously subject to a low-velocity impact. The schematic diagram
for this facility is shown in Fig. 2. During the preloading the applied compressive load, lateral
displacements and in-plane strains were recorded by a 96-channel data logger. When the plate was
impacted the dynamic responses in terms of the impact force, deflection of the plate centre and
strain histories were recorded by a transient data logger. Internal delamination damage was
detected by a 3-D image ultrasonic C-scan device.

2.3. Compression-after-impact tests

Since the loading jack limit for the above preload rig was only 225 kN, another machine with
a load capacity of 2500 kN was employed to cover the full range of CAI tests for the thicker plates
including the undamaged plates. The anti-buckling support rig was presented in [10]. The loaded
ends of plate were clamped, and the other two edges were simply supported by blade-guide
restraints as described also in [10].

3. Dynamic relaxation technique and its applications

To be able to predict the final compressive failure load after impact, it is necessary to simulate the
dynamic impact and consequent material degradation by in-plane failure or delamination. This
degraded plate can then be subjected to a compression load until failure occurs, which itself may be
dynamic in nature. The initial equilibrium shape can be found by standard Newton—Raphson or
arc length non-linear solution methods. However, these are relatively time consuming and the
490 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

solution technique is not directly compatible with the explicit time marching scheme used in the
dynamic solution for impact simulation.
Another problem, which occurs near and after buckling, is that the panel may switch modes,
often dynamically in the testing machine. This is particularly likely for periodically stiffened panels
where the eigenvalues tend to cluster with 2 or 3 modes having values within 1—2% of each other.
Surprisingly, this mode switching was also found occurring even for our nearly-square plate.
To avoid the complexities of mode switching and indeterminate tangent stiffness, it was decided
to solve even these static problems in a dynamic fashion by computing the response for long
enough time for the damping to have caused the velocity and accelerations to die away to zero.
This approach — usually called ‘dynamic relaxation’ — is well known, but has fallen out of favour for
structures modelled by finite elements, as will be explained. The equation to be solved is
[M]r(#[C]rR #[K]r"[R(t)], (1)
where M is the mass matrix, C the damping matrix, K the stiffness matrix; R the applied loads and
r the displacements. Note that if the acceleration r̈ and the velocity r are both zero then this reduces
to the static equation
[K]r"[R] (2)
which we want. The idea behind dynamic relaxation, particularly for non-linear problems which
need an iterative solution anyway, is to apply the loads R and solve the equations of motion with
an “optimum damping” so that the static solution is obtained as quickly and efficiently as possible.
The details of this new routine are in [10,11] but a brief description is given here.
If Eq. (1) were for a simple single degree of freedom, it is possible to select the damping coefficient
C at its critical value, C"2(mk, so that no oscillations occur but also the system is not so highly
damped that it takes an abnormally long time to relax to the static solution.
For multi-degree of freedom systems we usually prefer proportional damping, i.e.
[C]"a[M]#b[K] (3)
which makes the equations much easier to integrate, but unfortunately in selecting [C] we have
only two constants available, a and b. The a term will be influential at the lower frequencies and
b at the higher. However, in typical structures modelled by finite elements, the natural frequencies
may span several orders of magnitude, and it is not easy to select an optimum. However, since [M]
does not influence the final (static) solution we may choose arbitrarily [M]"[K] and thereby
making all the eigenfrequencies of the undamped system (4) equal to unity:
"[K]![M]u""0. (4)
This completely artificial model bunches all eigenvalues together and enables us to use a single
critical time step and damping factor which does not have to compromise between global
fundamental (low) frequencies and high-frequency responses. We use an explicit solver which has to
form [M]\"[K]\ which is fully populated. If the stiffness matrix changes due to large
deformations or material degradation [K]\ needs to be updated during the course of solution.
This of course will not deliver a realistic dynamic response, but the static solution which we want is
correct. The method has been found to be quite economical compared to implicit solvers with
conjugate gradient and preconditioning — at least for moderate degrees of freedom. Moreover, it
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 491

has been found to be completely robust: convergence to the static buckling mode, even after
a snap-through, is shown to be very rapid.
The dynamic relaxation technique was firstly checked by a shallow arch test case, since “snap
through” is a standard test for buckling codes. A plate postbuckling test case was also carried out.
The results agreed remarkably with two static solutions, using the arc length and Newton—
Raphson. Details are presented in [10,11], but the modelling of the composite plates described in
Section 2 is demonstrated by the following examples.

3.1. Simulation of buckling/postbuckling behaviour of undamaged plates

As described the plates used for FE simulation and experimental test have the dimension of
270;250 mm, but different thickness of either 2 or 4 mm. The four edges are clamped in the
preload rig, which gives a plate size of 200;200 mm for the FE modelling. An initial imperfection
of 10% the plate thickness is used for the FE analysis. The FE model with coarse mesh and the
boundary condition for clamped edges is shown in Fig. 3. Eight-noded Mindlin plate elements were
used including transverse shear deformations which are not negligible in carbon-epoxy plates, even
when plates are thin. The lateral deflection of the plate centre is also indicated in the figure.

3.1.1. 2 mm thick plates


For these thin plates, the predicted initial buckling load is 29 kN. Since “snap through” is
a standard test for buckling codes, this case is demonstrated first. The applied load should be high
enough to take the plate well into the postbuckling region, therefore a load which is twice the initial
buckling load, P"2P , is applied instantaneously at time zero. Fig. 4 shows the plate’s central
 
deflection, d, versus number of iterations. Initially, the plate tended to buckle to its fundamental
shape of one half-wave length but eventually snapped through into a mode dominated by the
second two-half-waves mode, in which the central deflection is very small: d"1.23 mm. The
solution converged after about 2200 iterations, and the computing time is about 2 min using an
IBM RS workstation.
If the history of a loaded plate’s behaviour is needed, then FE77 also allows a series of loads in
terms of P/P to be applied in a single computer job. The solution iterates to converge very
 
efficiently at each load level before applying the next load level. An example showing the lateral

Fig. 3. Finite element model used for plate buckling/postbuckling simulation.


492 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 4. A dynamic relaxation solution (2 mm thick plate) showing lateral deflection of the plate’s centre at P"2P .
 

Fig. 5. Numerical: displacement history of the plate centre (2 mm thick plate) (predicted P "29 kN).
 

displacement history of plate centre under several load steps is presented in Fig. 5. The maximum
load applied is 5P , well into the post-buckling region. It shows that below 1.5P the plate
   
buckles in a fundamental half-wave length curve, Fig. 5b; the mode changes at 1.5P and the plate
 
snaps through into a two half-wave shape and no further changes as shown in Fig. 5c. The tests also
show that these thin plates buckle at strains well below the matrix and fibre failure thresholds.
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 493

Fig. 6. Experimental: post-buckled shapes at different load levels (2 mm thick plate) (measured P "30 kN).
 

Fig. 6 is an experimental test recording the plate out-of-plane displacement by three LVDTs (as
shown in Fig. 1) for the same plate as above. It confirms that the plate indeed buckled at about
29 kN as predicted, and the mode change occurred at 44 kN (about 1.5P ). This example has
 
demonstrated that the numerical simulation can handle the plate buckling problem well into the
postbuckling region.

3.1.2. 4 mm thick plates


This thicker plate has been chosen so that the buckling stress and material strength are much
closer than the previous application. We therefore expected a strong coupling between the impact
dynamic response, the destabilising compressive loading, and material failure. The experimentally
measured P is about 175—180 kN compared to the FE77 predicted P "180 kN. Since the
   
load-cell capacity was only 225 kN (50,000 lb), the maximum compressive preload that could be
applied for this thicker plate is about 220 kN (+1.25P ). Fig. 7 presents a finite element
 
prediction of the plate centre deflection under a series of applied loads. It shows that this thicker
plate will also snap through at a load value of P'1.2P .
 

3.2. Simulation of impact-with-preload

The FE code simulates the process of impact-with-preload in a sequential fashion using one
dynamic routine: firstly the dynamic relaxation routine is used to achieve a static solution under
the applied preload as described in Section 3.1; then the (calculated) compressive stresses
and the buckled shape is held as the initial condition of the plate before receiving the low-velocity
impact. The results of FE simulation and experimental tests will be presented and discussed in
Section 4.
494 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 7. FE simulation of plate buckling—postbuckling behaviour (4 mm thick plate).

4. Impact with compressive preload: experimental versus FE simulation

It is known that impact damage by delamination depends directly on the maximum impact force
induced whatever the incident energy and plate size [2]. Will this force be affected by the presence
of a preload in the vicinity of the buckling load? Near the initial buckling load, a plate will have
very little resistance to a normal force until the displacements become large enough for the
membrane action to become significant. To investigate this problem, low-velocity impact tests with
compressive preload were conducted. The test plan is shown in Fig. 8.

4.1. Thinner plates (2 mm)

Fig. 8 indicates that at P/P "1—1.5 the plate buckles into a single-wave-length curve, either
 
upward or downward depending on the plate’s initial imperfection. Because these plates buckle at
strains well below the matrix and fibre failure thresholds, the preload effect on impact damage is
little. When the compressive preload was increased beyond 1.5P the buckling mode changes and
 
the plate snaps through into a two half-wave length curve as shown in Fig. 8. This post-buckled
shape unloads the pre-stress in the plate centre where it is impacted. The test results have shown
that the effect of preload on the impact damage of these thin plates is small [10].
The impact force histories and delamination damages of the 13 J impact tests at different
preloading levels are shown in Fig. 9, in which the experimental results are compared with the FE
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 495

Fig. 8. Test plan for impact-with-preload.

solutions. It shows that for the thin plates the maximum impact forces, all around 3000 N, are
virtually unaffected by the preloading and the plate’s initial buckling shape. The load history is
initially affected by whether the initial buckle is ‘up’ or ‘down’, but the former snaps through
immediately and most of the force history takes place when the large deflection membrane stresses
are tensile. Consequently, the delamination damage areas are almost the same. The FE77 simula-
tion agrees well with the test results and it has captured the highly truncated force history, due to
in-plane degradation. The Chang—Chang composite damage model for in-plane fibre breaking and
matrix fracture [12,13] is used in the FE model with the material properties quoted in Table 1. The
C-scans all display a similar elongated shape at 45° to the loading direction. This delamination is
confined to the back face and is due to the bending strains which cause matrix cracking which in
turn causes delamination. The outer lamina fibres are at 45° in these specimens and offer no
resistance to matrix cracking.
These examples have demonstrated that the dynamic response may depend on initial buckled
shape and amount of preload, but the peak impact force and hence the impact damage were hardly
affected for these thin plates. Moreover, the post-impact compression tests [10] have shown that
these plates failed in regions where the bending strains were large, and in the postbuckled state this
is no longer in the middle of the plate. Hence a conventional coupon impact test followed by CAI
evaluation would be irrelevant. Therefore the CAI test results are not included in this paper.

4.2. Thicker plates (4 mm)

For the thicker plates the buckling stress and material compression strength were much closer
than for the thinner plates. Therefore, a strong coupling effect between the impact dynamic
response, the destabilising compressive loading, and material failure was expected. The aim was to
see if this coupling resulted in a failure during impact at an applied stress lower than a compres-
sion-after-impact test on a plate subjected to the same incident impact energy. The results are
presented and discussed in the following section.
496 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 9. Force histories and damages of impact-with-preload tests of 2 mm plates.


X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 497

4.2.1. Experimental
For the 4 mm plates, the critical buckling load P predicted by the FE model is 180 kN. The
 
experimentally measured value is very close, between 175—180 kN. Since the load-cell capacity is
225 kN, the maximum compressive preload that can be applied for this thicker plate is about
220 kN (+1.25P ). The threshold impact energy to cause internal delamination damage is about
 
8 J. Thus the tests were designed for three incident energy levels of 12, 20 and 32 J, and also at
several preload levels. Eight plates have been tested: three plates were impacted without compres-
sive preload and at 12, 20 and 32 J, respectively; the rest of the plates were preloaded in
compression according to the test plan and subsequently impacted while the preload was held. The
ultrasonic C-scan pictures for all tests are presented in Fig. 10. These images show the damage
extent against different preload ratios at each specified impact energy level. The test results are
summarised in Table 2. The preload is given in terms of the initial buckling load and the mean
strain is the average of strain gauges. Please note that the damage areas presented in Table 2 are
internal delaminations (caused by interlaminar shear force). The back-face splitting and associated
delamination (caused by bending strains) are not included. This is because that the CAI strength is
sensitive only to internal delaminations. The CAI test results for these plates are also shown in the
same table but will be discussed in Section 5.
(a) Low and intermediate impact energy levels: Considering first the lower impact energy 12 J tests,
the damage at no preload is the usual circular envelope due to interlaminar shear-driven failure
near the plate mid plane expected when the force threshold has just been exceeded. The damage at
P"P is actually slightly less but so is the maximum force of 5000 N as indicated in Table 2. This
 
reflects the loss of stiffness as P approaches P . The big difference occurs when the applied load
 
was increased to 1.2P : the damage extent revealed by the C-scan is now extensive and in addition
 
to the usual back face delamination there is also much more interior delamination elongated in
a direction transverse to the applied compressive load. This is similar to the type of unstable
delamination that takes place when a damaged plate is subjected to a CAI test, and suggests that
the impact dynamics may have been on the point of precipitating a compressive failure, due to the
extra compressive stresses inducing compression degradation in the front face region and conse-
quent unstable local buckling of the delaminated sublaminate in the front face region. (Some 3-D
C-scan pictures will be shown only for the 20 J impact group.)
The ANDSCAN 3-D images were also used to improve these results as the conventional C-scan
maps only give the envelope of the total damage. Fig. 11 shows one of the tests (plate C4-4). The
higher energy produces higher forces (Table 2) as expected but little difference in the maximum
force values for P"0 and P"P , repeating the lower energy test behaviour. The total damage at
 
P"P is however greater and different in character. Comparing Fig. 10d and e we see the back
 
surface splitting developing, due to the higher bending strains, along the local fibre direction
(i.e. !45° in this case). There is also a transverse delamination which appeared in the previous
tests (12 J) at the maximum value of compression preload. This is confirmed in the 3-D AndScan
(Fig. 11), so it looks as if the induced bending compression stress are on the point of initiat-
ing a compression failure. This is exactly what happened when the preload was increased to
P"1.2P , but energy still 21 J, as indicated in Table 2. The maximum force did not even reach
 
5000 N so we have a clear case of failure during impact at a lower compressive load than that which
would cause failure when applied to the plate after the impact energy of 20 J caused the damage.
The delamination in Fig. 10f has propagated to the plate boundary and the photo Fig. 12 shows
498 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 10. Damage maps of the impact-with-preload CFRP plates (4 mm thick).

a transverse compression buckle (in the 90° direction) of the upper region of the laminate even
though the top surface is a 45° ply.
Fig. 11 indicates the top face delamination extending at right angles to the loading direction,
characteristic of unstable blister growth, and shows that CAI failure is imminent at preload 180 kN
and when preload increases to 216 kN it clearly fails at impact — indicating a possible failure when
preloaded at between 180 and 216 kN, and probably near 180. This is significantly different to CAI
with no preload at 226 kN (see CAI tests in Section 5).
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 499

Table 2
Impact and CAI test results for the large plates (nominal thickness "4 mm)

Test Thickness Preload at impact Impact test results CAI strength

No. (mm) P/P Mean strain Energy F Damage Area Force


  
(Microstrain) (J) (N) (mm) (kN)

C4-12 4.35 0.0 0 12 5400 550 250


C4-1 4.29 1.0 !3100 12 5000 460 270
C4-2 4.30 1.2 !3800 12 5000 470 275
C4-3 4.37 0.0 0 20 6200 780 225
C4-4 4.31 1.0 !3200 20 6000 750 230
C4-5 4.35 1.2 !3800 21 4900 Failed on impact with preload" 200
C4-6 4.30 0.0 0 32 6450 940 185
C4-8 4.31 1.0 !3200 33 6350 1430 180

C4-7 4.31 N/A N/A 0 0 0 340

This is the area of the envelope of internal delamination determined by 3-D C-Scan. It does not include the back-face
splitting by bending and the additional surface delamination caused by preloading.

Fig. 11. 3D C-scan pictures of impacted plate C4-4, P/P "1.0, incident energy"20 J (2D scan previously shown in
 
Fig. 10e).
500 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 12. Photograph of plate C4-5: it failed during impact (P/P "1.2, 21 J).
 

(b) Highest impact energy level: For the highest incident energy (32 J), plate C4-6 was impacted
without preload while C4-8 subjected a preload of P"P . The maximum impact force for the
 
first impact is 6450 N and the damage is a combination of delamination and back face matrix
failure (fibre splitting). The delamination area is about 940 mm. The back-face damage extent is
130 mm. At P"P the maximum force of 6350 N is slightly lower but the damage is significantly
 
greater and different in character as we found in the 20 J impact tests. The delamination
area is about 1430 mm. The back face damage extent is 180 mm. There is also an internal
delamination perpendicular to the preload direction with the total length of about 80 mm,
see Fig. 10h.

4.2.2. Finite element simulation


To demonstrate that the code FE77 can mimic the dynamic response with preload effect, using
the novel dynamic relaxation approach described in Section 3, two test cases (both under 20 J
impact) are demonstrated here. The Chang—Chang composite damage model, which has tensile,
compressive and shear in-plane failure based on unidirectional tests [12,13], has been incorporated
into the FE77 code to simulate the in-plane degradation during impact. The code looks at the
stresses at all laminar levels in the laminate, for each element, during the impact event. If failure
threshold is achieved then that layer is given zero stiffness thereafter. Firstly, for the impact test
without preload (case C4-3), Fig. 13 shows a comparison of the predicted and recorded impact
force histories. The clearly truncated top of the force response indicates that in-plane damage has
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 501

Fig. 13. FE simulation of test C4-3 (P/P "0, incident energy"20 J).
 

occurred. The predicted force with in-plane degradation gives the same maximum force of 6000 N
as the experiment. In the same figure it shows that without considering degradation the maximum
force would be overestimated by 15%. At this energy level the damage is quite extensive, involving
both fibre and matrix failure. The C-scan picture shows a total damage area of 780 mm, mostly
interior delamination. The predicted in-plane damage area is about 1000 mm as shown in
Figs. 13b and c. The damage area is calculated by simply counting the Gauss points which have
failed due to either fibre breakage or matrix fracture. If a 2;2 Gauss integration rule is used each
Gauss point represents a quarter of the total finite element area. The computing time is about one
hour for this 4 ms impact event using an IBM RS workstation.
For the same impact energy but with preload of P"P (case C4-4), the force histories are
 
presented in Fig. 14. Again the predicted force response agrees with the test result. The
502 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 14. FE simulation of test C4-4 (P/P "1.0, incident energy"20 J).
 

3-D ultrasonic scan indicates that the damage includes both interior delamination and the back
surface bending failure. The predicted in-plane damage is dominated by the back surface
matrix fracture in tension, with the damage area of 950 mm compared with the test value of
760 mm.
It has been found that the modelling of damage degradation is very sensitive to the strengths of
fibre and matrix used in the code. For this T800/924 material, the manufacturer recommended
strength values have been used in our simulations, except that a lower value for the matrix tensile
strength of 75 MPa was used instead of 93 MPa, and this was confirmed by a separate coupon test
presented in [10].
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 503

5. Compression-after-impact (CAI) tests (4 mm plates)

In the test programme, the plates were tested in compression to failure in a panel testing machine
with a capacity of 2500 kN. This made possible the CAI testing of plates with no or little damage,
which were beyond the capacity of the preloading device under the impact test rig. A fresh plate
(No. C4-7) without impact damage was tested first to check the critical buckling load and
compressive strength. Initial buckling was defined by the mean strain versus load response, and the
strain reversal point is taken as the onset of buckling. The compressive strength of the fresh plate
should provide a baseline for the following CAI tests, in which those plates that did not fail
catastrophically by impact were loaded in compression to their final failure to find their residual
compressive strength. The test results are summarised in Table 2.

5.1. Undamaged plate (control plate)

The control plate (C4-7) failed at 340 kN. The residual shape of the plate after failure is
illustrated in Fig. 15. The photograph indicates a classical compression failure: side 2 failed by fibre
compressive failure (note the sharp micro-buckling edges at right angles to the fibre) and side
1 failed in tension.

5.2. Low-energy impacted plates (12 J )

Plates C4-12, C4-1 and C4-2 had mostly the usual internal delamination damages having
a circular envelope, but plate C4-2 (which had the maximum preload of 1.2P ) also had additional
 
delamination transverse to the compressive preload similar to the damage map of C4-4 as shown in
Fig. 11. We see that applying the preload reduces the stiffness of the plate and leads to a reduction
in damage area and increase in CAI strength. For these plates the final failure occurred between
250—275 kN, substantially lower than 340 kN for an undamaged plate. This means 19—26%
reduction in residual compressive strength.

5.3. Intermediate energy impacted plates (20 J )

In this group the preload of P again lowers the damage area and leads to a slight increase in
 
CAI strength, but increasing the preload to near the CAI strength results in catastrophic lateral
propagation of the induced delamination. The CAI strength for these plates is between
225—230 kN, representing 32—34% reduction in compression strength.
To compare the test results of plate C4-4 (failed during CAI) and C4-5 (failed under impacting
while preloaded in compression), 11 strain gauges were arranged across the plate C4-4 as shown in
Fig. 16. There are six gauge points and the gauges are in pairs (except No. 11), with those numbers
in the brackets being on the other side of the plate. Gauges 5, 6 and 11 were placed near impact
damage zone to record lateral damage propagation, if any. The compressive preload-induced
delamination may propagate under the compressive loading and this should be picked up by these
gauges. The surface strain distributions are shown in Fig. 17a—c. Firstly the strains near the vertical
supported edges are uniform (gauge pairs 1 & 2, 9 & 10), recording the compressive strain of 0.7%
504 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 15. Photograph of plate C4-7 after failure in compression (control plate without impact damage).

(1 & 9) at final failure. Gauges 1, 3, 7 and 9 are in compression all the time, with gauge 3 measuring
the final failure strain of 0.95%, which is the highest compressive strain of all. Notice that there is
a distinct change in the pattern of the surface strains. Gauges 3 and 7 on the impact side increase in
compression after global buckling and the gauges on the opposite face (4 & 8) swing back into
tension as the buckling mode develops. This development is earlier than the undamaged control
plate due to the residual eccentricity caused by impact. However, at 215 kN the compression
in 7 suddenly decreases indicating a local blister which puts gauges on both sides of the plate
(gauges 7 & 8) into tension. The surface strain in gauge 11 showed a more idiosyncratic behaviour
but it was placed very close to the damaged surface. When a local blister forms both outer surfaces
bend in tension and there is a load shedding as the mean strains indicate for gauge points 7 & 8,
5 & 6, as shown in Fig. 17d. The bending strains will also decrease abruptly since both surfaces
are moving into tension. It is unfortunately not possible to record the true bending strains in
both halves of the blister since this would need gauges on the inside surfaces! The local
blister propagation occurs in test C4-4 because the total damage area of 1100 mm is greater than
was in C4-1 and in our simulation of local blister development [10] it was found that a critical
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 505

Fig. 16. Strain-gauge positions for CAI test (Plate C4-4).

blister size was necessary for the mode to change from global to local. Fig. 18 sketches our
interpretation.

5.4. Highest energy impacted plates (32 J )

The highest impact energy of 32 J produced much larger delamination and back face bending
failure. In addition, plate C4-8 (which had the preload of P"P ) also had internal delamination
 
transverse to the compressive preload. The transverse delamination is interpreted as an unstable
blister which did not go completely unstable and propagate to the edges of the plate, but is clearly
on the point of so doing. We would expect this plate to have a correspondingly low CAI strength.
Indeed the CAI failure load for both plates C4-8 and C4-6 is between 180—185 kN, about 46%
lower than that of undamaged plate. The photograph of CAI damaged plate (C4-6) is shown in
Fig. 19. Both sides of the plate have clearly local blister buckling and growth transverse to the
load direction. This confirms the damage mode change illustrated in Fig. 18. Note that the control
test plate did not have mode change before its final failure (Fig. 15).
Finally, the usually presented relationship of residual strength versus impact energy for the 4 mm
thick plates is shown in Fig. 20. In this case the reduction in compression strength is between
18—46% depending on the impact energy and amount of preloading. It should be noted that one of
the preloaded plate, No. C4-5 under P"1.2P , failed during impact at a lower applied load than
 
that which would cause failure when applied after impact. The catastrophic failure-during-impact
occurred when both the preloading and incident energy were sufficiently high. Thus, there is
a coupling effect between the impact energy, the destabilising compressive loading, and material
failure.
506 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 17. Surface and mean strain distributions of plate C4-4 (P"P , 20 J).
 

6. Conclusions

In assessing the nature of impact damage, compressive loading, and the coupling between
the two, there are three characteristic non-dimensional parameters. Firstly, the ratio of
P/P which can affect the dynamic response and impact force when this ratio is of order one (a
 
state of neutral equilibrium). Secondly, the ratio of applied compression strain to the material’s
failure, which can affect the amount of in-plane damage during the impact-induced strains
of a preloaded plate. Finally, the ratio of impactor energy to its threshold value (or better
the maximum impact force equivalent) dictates whether internal delamination will occur, and
this form of damage has the potential to seriously degrade the CAI strength. It was found
that the sensitivity to preload effects was only apparent when these parameters were all of
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 507

Fig. 18. Damage mode change in CAI test of plate C4-4.

Fig. 19. Photograph of plate C4-6 after failure in compression (impacted with the incident energy of 32 J).

order unity. Thus:


E The thinner plates (2 mm) buckle at low strains so the second parameter is much less than one.
The reduced postbuckling stiffness attenuates the force so the third parameter is reduced. The
CAI strength actually increases with preload.
508 X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509

Fig. 20. Residual compressive strength versus incident kinetic energy for carbon/epoxy plates (t"4 mm): experimental
results and fitted line.

E For thicker plates (4 mm) the second parameter now approaches unity with the first and
unstable blister propagation nearly occurs at 20 J, P/P "1. When this figure increases to 1.2,
 
the propagation occurs during impact but at a lower applied pre-load than the CAI failure (see
Fig. 20 for example).
Finally, we have found that to predict compression-after-impact strength and to predict the
amount of impact damage, it is necessary to solve the problem dynamically. It was found that even
for static buckling well into the postbuckling regime it was safer to treat the problem dynamically,
and if no dynamic effects were present a new form of ‘dynamic relaxation’ was more robust than
conventional non-linear incremental strategies.

Acknowledgements

The authors would like to acknowledge the support of DTI/DERA under the CARAD scheme,
contract number FRN1c/532.

References

[1] Davies GAO, Zhang X, Edlund A. Predicting damage in composite aircraft structures due to low velocity impact.
Aerotech’94, Birmingham, UK, January 1994.
[2] Davies GAO, Zhang X. Impact damage prediction in carbon composite structures. Int J Impact Engng 16:
1995;149—170.
[3] Hitchings D. FE77 general purpose modular finite element system for static and dynamic, linear and non-linear
analysis. Dept Aeronautics, Imperial College, 1994.
X. Zhang et al. / International Journal of Impact Engineering 22 (1999) 485—509 509

[4] Davies GAO, Zhang X, Hitchings D. Modelling impact damage in laminated composites. Presented at NAFEMS
World Congress’97, Stuttgart, Germany, 9—11 April 1997, Proc published by NAFEMS Ltd, 1997:1216—31.
[5] Tweed JH, Lee RJ, McCarthy JC. Impact evaluation of stressed composites. European Conf on Composites Testing
and Standardisation, Amsterdam, 8—10 September. 1992:339—46.
[6] Sun CT, Chen JK, On the impact of initially stressed composite laminates. J Compos Mater 19:1985;490—504.
[7] Starnes Jr JH, Rhodes MD, Williams JG. Effect of impact damage and holes on the compressive strength of
a graphite/epoxy laminate. Non-destructive evaluation and flaw criticality for composite materials. ASTM STP
696, 1979:145—71.
[8] Ishai O, Shragai A. Effect of impact loading on damage and residual compressive strength of CFRP laminated
beams. Compos Struct 1990;14:319—37.
[9] Guild FJ, Hogg PJ, Prichard JC. A model for the reduction in compression strength of continuous fibre composites
after impact damage. Composites 1993;24:333—9.
[10] Zhang X, Hitchings D, Davies GAO. Damage tolerance of realistic stiffened compression panels. Final Report,
Contract FRN1c/F32, Department of Aeronautics, Imperial College, June 1996.
[11] Hitchings D. A new dynamic relaxation technique. In preparation.
[12] Chang FK, Chang KY. A progressive damage model for laminated composites containing stress concentrations.
J Compos Mater 1987;21:834—55.
[13] Choi HY, Chang FK. A model for predicting damage in graphite/epoxy laminated composites resulting from
low-velocity point impact. J Compos Mater 1992;26:2134—69.

Das könnte Ihnen auch gefallen