Sie sind auf Seite 1von 9

Available online at www.sciencedirect.

com
COMPOSITES
SCIENCE AND
TECHNOLOGY
Composites Science and Technology 68 (2008) 1348–1356
www.elsevier.com/locate/compscitech

A model to predict low-velocity impact response


and damage in sandwich composites
C.C. Foo *, G.B. Chai, L.K. Seah
Nanyang Technological University, School of Mechanical and Aerospace Engineering, 50 Nanyang Avenue, Singapore 639798, Singapore

Received 9 September 2007; received in revised form 9 November 2007; accepted 5 December 2007
Available online 15 December 2007

Abstract

Composite sandwich structures are susceptible to low-velocity impact damage, but existing elastic impact models cease to be valid
after the onset of damage. In this paper, a modified energy-balance model coupled with the law of conservation of momentum is
proposed to extend its validity beyond the elastic regime. Three parameters were first derived from the static load–deformation response:
the elastic stiffness, the critical load at the onset of damage, and the damaged stiffness. These parameters were then used in the impact
model to predict transient load and deflection histories for the plate subjected to impact. A three-dimensional finite element model was
also developed to analyse the static indentation problem, and core damage was identified to be one of the damage mechanisms at initial
failure. By accounting for the elastic energies absorbed by the plate up to initial failure in the core, the critical load was found to be
theoretically predictable. Numerical predictions compared well with experimental results.
Ó 2007 Elsevier Ltd. All rights reserved.

Keywords: B. Modelling; C. Failure criterion; C. Finite element analysis (FEA); C. Sandwich structures

1. Introduction [6]. Core materials are expected to substantially affect the


damage initiation characteristics of sandwich panels
Sandwich panels are widely used in lightweight con- because they generally have lower mechanical properties
struction especially in aerospace industries because of their than skins due to their lower density [7]. Hence core dam-
high specific strengths and stiffnesses. However, they are age, which is characterised by a substantial change in the
susceptible to impacts caused by runway debris, hailstones, load–deflection curve, has been identified as the first failure
dropped tools, etc [1]. Although the induced damage may mode in low-velocity impacts of honeycomb sandwich
be barely visible, especially for low-velocity impacts, the structures with a high skin-to-core stiffness ratio [3,5,8,9].
strength and reliability of the structure can be affected. Several analytical solutions have been proposed to pre-
Hence, the behaviour of sandwich structures under impact dict damage initiation [6,10,11], without considering dam-
has received increasing attention. age progression. An indentation model that consists of an
Research work to address impact damage resistance of elastic facesheet resting on an elastic perfectly-plastic foun-
composite sandwich structures, which deals with the dation has been used to model core crushing [10]. The fail-
response and damage caused by impact, is dominated by ure load for facesheet cracking was predicted using the
experimental studies quantifying damage [2–5]. Previous maximum stress and Tsai-Hill criteria [11], but the pre-
studies have indicated that damage initiation depends on dicted failure loads were underestimated. Damage initia-
panel support conditions, projectile nose-shape, and the tion loads for fracture of top facesheet, core shear failure,
material properties and geometry of the facesheet and core and tensile failure of the bottom facesheet had also been
considered and derived separately [6].
*
Corresponding author. Apart from testing and prototyping, finite element (FE)
E-mail address: keithfoo@pmail.ntu.edu.sg (C.C. Foo). simulations that include a progressive damage model and

0266-3538/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compscitech.2007.12.007
C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356 1349

account for nonlinear material behaviour can also be used results were used to validate the finite element model and
to predict impact response of composite structures [12–14]. analytical model. The facesheets were made from commer-
In numerical impact analyses of honeycomb sandwich cially available FibreduxÒ 913 C-HTA carbon fibre–epoxy
structures [15–17], the multi-cellular core is usually prepregs, and the core consisted of HexWebÒ A 1 Nomex
replaced with an equivalent continuum model and analysed honeycomb. The material properties of the laminate and
in terms of its effective properties. A continuum model may core are listed in Table 1. The test specimens, which mea-
seem a convenient way to represent the real core geometri- sured 100 mm  100 mm each, were hand-laid and cured
cally, but errors exist when it is used to model damage [18]. in-house. Table 2 shows the list of specimen types, with
One reason is that it may be difficult to simulate the exact varying core thickness, cell size and laminate’s orientation.
damage progression due to the discontinuous surfaces of The static indentation tests were conducted using the
the honeycomb core in contact with the facesheets. This Instron 5500 R test system operating under displacement
limitation can be overcome by modelling the honeycomb control at a constant cross-head speed of 0.5 mm/min.
core explicitly using shell elements to obtain more realistic For the drop weight impact tests, the Instron Dynatup
distributions of stresses and strains. 8250 impact testing machine was used. In all tests, the spec-
The spring–mass models [19–21] and energy-balance imen was clamped between two steel plates, each having an
models [2,20,22] are two of the few analytical solutions pro- 76.4 mm circular aperture, with the mid-point of the spec-
posed to study the impact response of composite sandwich imen directly located underneath the indentor/impactor. In
structures due to the complex interaction between the com- the static tests, the two plates were bolted in place, while
posite facesheet and core during deformation and failure. pneumatic clamp plates with a constant clamping force
However, these elastic models cease to be valid after the were used in the impact tests. The clamps secured the spec-
onset of damage and are unable to model damage propaga- imen without causing any buckling of the honeycomb core
tion. Furthermore, the energy-balance model only yields prior to testing. The indentor used in the static tests and the
the maximum impact force but not the entire load–time impactor used in the impact tests had the same diameter of
history. Similarly, analytical elastic indentation models 13.1 mm. The impactor mass was 2.65 kg. More details of
[23,24] are inaccurate after the onset of damage. Modified the test systems may be found in [28].
spring–mass models have been proposed to account for
damage [25,26], but they are heavily based on empirical
correlation; they introduce additional unknowns that have Table 1
to be determined experimentally. Material properties for composite sandwich panel
Most existing models are valid for elastic impact events, Part Property Value
but the impact response of the structure after the onset of Facesheet FibreduxÒ 913C-HTA carbon–epoxy
damage is not well dealt with. In this paper, the energy-bal- Longitudinal stiffness, E11 (GPa) 150
ance model [1] is modified to extend its validity beyond the Transverse stiffness, E22 (GPa) 9.5
Out-of-plane stiffness, E33 (GPa) 9.5
elastic regime. As the quasi-static test can be used to model Poisson’s ratios, m12 and m13 0.263
low-velocity impact events [2,9,27], a three-dimensional FE Poisson’s ratio, m23 0.458
model of the honeycomb sandwich plate and a rigid inden- Shear moduli, G12 and G13 (GPa) 5.43
tor is first developed to analyse the static indentation prob- Shear modulus, G23 (GPa) 3.26
lem. The geometrically correct FE model, which has been Longitudinal tensile strength, Xt (MPa) 1900
Longitudinal compressive strength, Xc (MPa) 1550
used previously to model aluminium sandwich panels Transverse tensile strength, Yt (MPa) 65.5
[28], also includes a progressive damage model for compos- Transverse compressive strength, Yc (MPa)a 140
ites so that accurate predictions of the damage mechanisms Interlaminar shear strength, S (MPa) 101.2
and failure are possible. An analytical model is then pro- Out-of-plane tensile strength, Zt (MPa) 65.5
posed to predict the critical load at the onset of damage Density, q (kg/m3) 1100
by considering the elastic energy absorbed by the plate Core HexWebÒ A1-64-13 Nomex honeycomb
up to damage initiation. Finally, the critical load, together Core height, hc (mm) 15
Cell size, c (mm) 13
with the elastic and damaged stiffnesses, is used in the mod-
Cell wall thickness, t (mm) 0.6
ified energy-balance model to predict the response of the Core density, q (kg/m3) 64
sandwich plate under low-velocity impact. The laws of con- Crush strength, q (MPa) 1.00
servation of energy and impulse-momentum are coupled to Through-thickness shear modulus, Gc (MPa) 55.0
derive the load and deflection histories. Numerical results a
Estimated from Ref. [38].
are validated with experimental data.
Table 2
Sandwich plate specifications
2. Experimental setup
Plate Skin configuration Core cell size (mm) Core height (mm)
An experimental investigation on composite sandwich A [0/90/0/90/0]s 13 15
plates subjected to quasi-static indentation and low-veloc- B [+45/45/0/90/0]s 13 15
C [+45/0/0/90/0/0/45]s 13 15
ity impact loadings was conducted. The experimental
1350 C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356

3. Finite element model deformations with reasonable accuracy. Surface-based tie


constraints were implemented at the interfaces between
3.1. Sandwich plate with discrete honeycomb core the facesheet and core, which were assumed to be perfectly
bonded. Finally, the 13.1 mm diameter steel spherical
The explicit finite element computer software, ABA- indentor was modelled as a rigid body using a rigid body
QUS/Explicit, was employed for the quasi-static analysis. constraint. It was meshed with four-noded linear tetrahe-
The sandwich plate was modelled as a circular plate, dron continuum elements with a Young’s modulus of
clamped at its top and bottom circumferential edges. A 200 GPa and a Poisson’s ratio of 0.3.
unit regular cell of the honeycomb was first modelled and
then replicated in the 1-direction and 2-direction to pro- 3.2. Progressive damage model
duce many individual cells. Assuming perfect bonding,
these cells were merged together to assemble the honey- A progressive damage model, which integrated stress
comb core, where the axis of cells was oriented perpendic- analysis, failure analysis and material property degrada-
ular to the facesheets (Fig. 1). The core was then meshed tion, was implemented in ABAQUS/Explicit using an user
with shell elements. subroutine (VUMAT). The stress-based Hashin criteria
The cell-wall solid was modelled as an isotropic, elastic [29] were used to detect four failure modes in the matrix
perfectly-plastic material, with the following properties: and fibre under both tension and compression. The Hashin
the modulus Es = 2 GPa, the Poisson’s ratio ms = 0.4, den- criteria in the three-dimensional form are given as
sity qs = 710 kg/m3, and yield strength rys = 30 MPa. A
FE investigation was conducted to simulate the out-of- Fibre tensile fracture; r11 P 0 : r11 ¼ X t ð1Þ
plane compression behaviour of bare honeycombs, so as Fibre compressive fracture; r11 < 0 : r11 ¼ X c ð2Þ
to qualify the material model for the core. At the peak Matrix tensile or shear cracking failure; ðr22 þ r33 Þ P 0 :
compressive load, the top edges of the honeycomb cells 2
started to yield. By modelling the cell-wall solid as an elas- ðr22 þ r33 Þ r2 þ r213 þ r223  r22 r33
2
þ 12 ¼1 ð3Þ
tic–plastic material, core damage was assumed to initiate Yt S2
when yielding occurred. Although the approach here does Matrix compressive failure; ðr22 þ r33 Þ < 0 :
not predict the buckling of the cell-walls, plastic folding "  #
2 2
of the cell-walls was observed as core crushing propagated. 1 Yc ðr22 þ r33 Þ
 1 ðr22 þ r33 Þ þ
Next, the facesheets were meshed with linear reduced- Yc 2S 4S 2
integration solid elements, with each ply in the laminate r212 þ r213 þ r223  r22 r33
represented by a solid element in the through-thickness þ ¼1 ð4Þ
direction. However due to the poor aspect ratio of these S2
elements, a reasonably fine mesh was required to achieve where the rij terms are components of the stress tensor. The
convergence. A graded mesh was thus created, where the notation of the quantities shown above refer to a local
region in the vicinity of the indentor was more finely layer coordinate system. The 1- and 2-axes are parallel
meshed. A mesh convergence study was later carried out and transverse to the fibres, respectively, and the 3-axis
to ensure the mesh refinement in the sandwich structure coincides with the through-thickness direction. The quanti-
was sufficiently fine enough to capture the stresses and ties in the denominators are the strengths in the corre-
sponding directions (Table 1).
For the case of fibre tensile failure, the interaction of the
shear stresses was found to result in conservative predic-
tions, as also observed in Ref. [30]. Hence, the contribution
of the shear stresses to the failure of fibres was ignored, and
the criterion for fibre tensile fracture simplified to the max-
imum stress criterion (Eq. (1)). Some researchers have used
a stress-based criterion to model delamination [12,30,31]
due to its relative ease of implementation in explicit FE
codes. Hou et al. [12] argued that compressive stress con-
strains crack opening, and thus delamination should not
be allowed to occur while an element is under compression
in the through-thickness direction. Moreover, previous
experimental studies found that delamination does not
occur in a narrow band immediately adjacent to the impact
point [31], which suggests that the stress field in that area
does not reach a level to initiate delamination. Thus, it
Fig. 1. FE model of composite sandwich plate with rigid spherical was assumed that interlaminar delamination occurred
indentor. under the following criterion:
C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356 1351

Table 3
Material property degradation rules
Mode of failure Material property degradation rules
Matrix tensile cracking E22 = m12 = 0
Matrix compressive failure E22 = m12 = 0
Fibre tensile failure Material point deleted
(all stiffnesses and stresses set to 0)
Fibre compressive failure Material point deleted
(all stiffnesses and stresses set to 0)
Delamination under tension E33 = G13=G23 = m13 = m23 = 0

 2
r33 r223 þ r213
Delamination;r33 P 0 : þ ¼1 ð5Þ Fig. 2. Local indentation model.
Zt S2
Once failure was detected in an element, degradation was
son’s ratio of the facesheet respectively. a0 represents the
enforced by assuming that damage within an element had
transverse local indentation at the mid-point of the top
an effect on the elastic properties of that element only.
facesheet, and q is the crushing stress of the core. The initial
The degradation rules are listed in Table 3. In the case of
local stiffness was attained by linearising Eq. (6) using Tay-
fibre failure, failed elements were entirely removed. Due
lor series expansion around some small value of indenta-
to restrictions on engineering properties of composites,
tion, for example a0= 0.1 mm, at the instant of first
the Poisson’s ratios were degraded for compatibility.
dP 
contact, i.e., K loc  da 0
 .When the sandwich plate is
a0 ¼0:1mm
4. Analytical model clamped around its edges, it experiences both local and glo-
bal deformation. Global deformation, w, refers to the
4.1. Elastic response of sandwich plate under quasi-static bending and shearing of the entire plate. An interaction
indentation between the local and global deformations exists in reality
[19]. As the core crushes during local deformation, its
An important first step in the impact analysis of sand- height decreases and the global bending and shearing stiff-
wich panels is to determine the elastic stiffnesses of the nesses of the sandwich plate reduce. However, this effect
sandwich plate prior to damage. One approach is to decou- can be ignored since the impact damage is small and local-
ple the local and global responses and ignore any interac- ised around the impactor for low-velocity impacts. In addi-
tion between the two, so that these stiffnesses can be tion, as the core is much thicker than the facesheets, the
determined separately. A rigidly supported sandwich plate membrane stretching of the facesheets is assumed to be
undergoes only local deformation of the top facesheet, a, negligible. The membrane stiffness of the core is also negli-
which consists of both indentation of the top facesheet gible [22]. Thus, the load sustained by the plate is related to
and core crushing. The Hertzian contact law has been used this global deflection [20] by
for isotropic homogenous linear elastic bodies when the
P ¼ K glo w0 ð7Þ
indentation is much smaller than the thickness of the plate
[1]. However, for a sandwich plate, where the facesheets are where Kglo = KbKs/(Kb + Ks) is the effective global stiffness
stiff and the core is flexible, core crushing becomes signifi- due to bending stiffness, Kb, and shear stiffness, Ks Assum-
cant and the Hertzian contact law may not be appropriate ing an uniform pressure acting over the central part of the
here. panel in the contact area with radius Rc, the shear stiffness
The principle of minimum potential energy has been of the panel, Ks, can be expressed as [22]:
used by several investigators to determine the stiffnesses 2
4pGc ðhf þ hc Þ
of composite panels [19,22]. Zhou and Stronge [22] mod- Ks ¼ ð8Þ
elled the quasi-static indentation of a circular panel resting hc ð1 þ 2lnðRp =Rc ÞÞ
on a rigid foundation (Fig. 2). By accounting for the bend- where Gc is the through-thickness shear modulus of the
ing and membrane stretching energies in the top facesheet, core; hc is the thickness of the core; Rp and Rc are the outer
as well as the work done to crush the perfectly-plastic core and contact radii of the plate, respectively. An initial value
and the work done by the contact load, they derived the of Rc, which depends on the indentation, is required to
localised load–displacement relation: obtain Ks Further numerical calculations showed that the
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
" #ffi effective shear stiffness is insensitive to this assumed contact
u
u
16p t ð7505 þ 4250mf  2791mf Þa0 2 2
P¼ Df qa0 1 þ ð6Þ radius, as also observed in [22], and hence it may be justi-
3 17640h2f fiable to substitute the mean value for the function, ln(Rp/
Rc), for the range of a0 from 0 to Rind (radius of indentor).
1
and Df ¼ Ef h3f ½12ð1  v2f Þ where Df, Ef, hf, and vf are the The mean value was calculated to be 2.2. The bending stiff-
bending stiffness, Young’s modulus, thickness, and Pois- ness of the clamped circular panel, Kb, is given as [20]:
1352 C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356

16pD may exist for honeycombs that have relatively thick walls.
Kb ¼ ð9Þ
ð1  v2f ÞR2p The area under the load–displacement curve up to the peak
load is the energy absorbed by the core at initial failure, UPk
where D ¼ Ef ðhf þ hc Þ2 hf =2 þ Ef h3f =6 is the bending stiff- Consider a honeycomb core that is discretized into springs
ness of the panel. Consequently, the elastic stiffness of the located at each vertical edge of the hexagonal cells. By exam-
plate before the onset of damage, K0, is then, ining the number of vertical edges in the core loaded in the
1 1 1 flatwise compression tests, Nv, the energy absorbed by each
¼ þ ð10Þ
K 0 K loc K glo vertical edge at onset of failure is, U Pk Pk
edge ¼ U =N v .In this
study, flatwise compression tests on 9 bare honeycombs
4.2. Initiation of damage cores with 3 different core sizes (9, 33 and 60 cells) were first
conducted, as previously described [34]. 3 samples for each
At the onset of damage, there is a substantial drop in the core size were tested, and U Pk edge was found to be 0.050 ±
global structural stiffness due to core failure [3,5,8,9]. Thus, 0.003 J (mean and standard deviation). With this, the energy
a method was developed to predict the critical load at fail- absorbed by the core up to initial damage under indentation
ure initiation by considering the elastic energy absorbed by loading, U th core , was estimated by summing the energy
the plate up to the onset of core damage. absorbed by each damaged vertical edge in the core, while
Under indentation loading, the core in the sandwich neglecting the remaining vertical edges, i.e. U th core ¼
P Pk
plate is subjected to both shear and out-of-plane compres- U edge . This is reasonable since the initial damage is
sive stresses [32]. However Aminanda et al. [33] examined expected to be highly localised in a few cells near the
the behaviour of Nomex honeycomb core under compres- indentor due to a lack of deformation in the neighbouring
sion, and found that the compression load is supported cells.
mainly by these vertical edges in the hexagonal cell. Next, the elastic strain energy of the top facesheet at the
Neglecting the influence of shear loading, they then mod- onset of damage, U th TFS , was calculated. The top facesheet
elled the honeycomb core as a grid of nonlinear springs was assumed to be elastic at the onset of core damage.
located exactly at the vertical edges, and accurately pre- Additionally, the interaction between the core and the
dicted the indentation of sandwich structures with metallic top facesheet, as well as the deformation of the bottom
skins and honeycomb cores. Hence, shear loading of the facesheet, was ignored. For an undamaged facesheet, its
core was also neglected here. elastic bending energy is expressed as [22],
Fig. 3 shows a typical load vs. displacement curve
Z Z " 2 2  2 #
obtained from the compression tests on bare honeycombs. Df 2p a oa 1 oa 2v oa o2 a
Ub ¼ þ 2 þ r dr dh
Initially, the compressive load increases linearly due to the 2 0 0 or2 r or r or or2
elastic bending of the cell walls until a critical load, PPk, is
reached. After the peak load, a sharp drop to a plateau load ð11Þ
is observed. This load drop corresponds to the onset of plas- The boundary condition of the localised indentation
tic deformation of the vertical edges [33], and core failure is area is assumed to be clamped. For a clamped circular
assumed to initiate at this point. This failure mechanism plate, the profile of the local indentation is approximated
[22] by
 2
r2
aðrÞ ¼ a0 1  2 ð12Þ
a
where a0 and a are the transverse deflection and the radius
of region of local indentation on the top facesheet, respec-
tively (see Fig. 2). Eq. (11) can then be simplified as
32pDf a20
Ub ¼ ð13Þ
3a2
Zhou and Stronge [22] showed that the radius of the region
of local indentation on the top facesheet, a, can be given
as
sffiffiffiffiffiffiffiffi
3P
a¼ ð14Þ
2pq

where P = K0Xind, and Xind is the displacement of the


Fig. 3. A typical load–displacement curve for compressive tests on bare indentor. Similarly for a clamped circular plate, the mem-
honeycombs. brane stretching energy can be determined as
C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356 1353
Z a 
N r er N h eh where Mimp denotes the mass of the impactor, Vimp is the
U m ¼ 2p þ r dr
0 2 2 impact velocity and V(t) refers to the velocity of the impac-
Z a tor at time t. The contact load is also a function of time,
pEf hf 2 
¼ 2
er þ e2h þ 2vf er eh r dr ð15Þ i.e., P = P(t). By the law of conservation of impulse–
1  vf 0
momentum,
2
where radial strain, er ¼ du dr
þ 12 ðda
dr
Þ , and circumferential Z t
u
strain, eh ¼ r . For vf = 0.3, Eq. (15) can be simplified to M imp ðV imp  V ðtÞÞ ¼ P dt ð22Þ
! 0
a40 Using Eqs. (21) and (22), the load and velocity histories
U m ¼ 2:59pDf 2 2 ð16Þ
a hf were solved. Subsequently, the deflection of the impactor
was obtained by integrating the velocity history. Once the
Finally, the number of damaged edges in the core, the local load reached the critical load (Pc), the elastic stiffness, K0,
indentation (a0 = Xind  w0), and the radius of local inden- was degraded to the damaged stiffness, Kdam, to account
tation (Eq. (14)) at the onset of damage were determined for damage. The integral on the right-hand side of Eq.
with the aid of the FE model described in the previous sec- (22) was approximated by the area under the load-time
tion. Consequently, the threshold energy at the onset of curve using the trapezoidal rule. No unloading was
damage, Uth, is the sum of the elastic energies of the core considered.
and top facesheet up to the onset of damage,
th 5. Results and discussion
U th ¼ U th th th
core þ U TFS ¼ U core þ ðU b þ U m Þ ð17Þ
Assuming that the load varies linearly with deformations 5.1. Load–displacement curves for quasi-static indentation
[35], i.e., P = Kloca0 = Kglow0, the energies used to deform
the plate can be quantified and expressed as a function of The FE model was validated by comparing the load–dis-
the load [4,22]. The energy due to local indentation in the placement plots of the sandwich panels from tests and FE
contact region is analysis (Fig. 4). The comparisons show that there exists a
Z a0 Z a0 good agreement between the experimental and predicted
P2 results in terms of overall trend. Furthermore, three
Uc ¼ P da0 ¼ K loc a0 da0 ¼ ð18Þ
0 0 2K loc parameters could be used to describe the deformation
Similarly, the energy due to global deformation in the form stages: the elastic stiffness (K0) to describe the elastic behav-
of bending and shear deformations is iour, the critical load (Pc) for damage initiation, and the
Z w0 Z w0 damaged stiffness (Kdam) for failure propagation (Fig. 4).
P2 Initially, the panels deformed in a linear elastic manner
U bs ¼ P dw0 ¼ K glo w0 dw0 ¼ ð19Þ
0 0 2K glo until a critical load was reached, where there was a sudden

Therefore, the total work done on the plate by the indentor


is
 
P2 1 1 P2
U Total ¼ U c þ U bs ¼ þ ¼ ð20Þ
2 K loc K glo 2  K0
The above equation implies that the load sustained by the
plate at any instant depends on the instantaneous work
done on the plate and its elastic structural stiffness. As a re-
sult of Eq. (20), the critical impact force at the onset of
damage [26] is related to the stiffness of ffi the structure and
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the threshold energy: P c ¼ 2  K 0 :U th .

4.3. Impact analytical model

The energy-balance model was modified to derive the


load and deflection histories for the sandwich plate under
impact. By the law of conservation of energy, assuming
no other energy losses, the total work done by the contact
load on the plate at any instant t is equal to the change in
kinetic energy of the impactor at that instant,
Fig. 4. Comparison of predicted and experimental load–displacement
P2 1
U c þ U bs ¼ ¼ M imp ðV 2imp  V ðtÞ2 Þ ð21Þ curves for Plates A, B, and C under quasi-static indentation (Solid lines for
2  K0 2 experimental results; dashed lines for FE results).
1354 C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356

Table 4
Comparison of predicted and experimentala values for elastic stiffnesses, critical loads, and damaged stiffnesses for Plates A, B and C
Plate K0 (kN/mm) Pc (kN) Kdam (kN/mm)
Predicted Experimental Predicted Experimental Predicted Experimental
A 1.34 1.37 ± 0.08 1.22 1.04 ± 0.06 0.523 0.675 ± 0.035
B 1.30 1.26 ± 0.06 1.17 1.09 ± 0.07 0.710 0.711 ± 0.060
C 1.69 1.94 ± 0.05 1.58 1.46 ± 0.09 0.704 0.863 ± 0.026
a
Experimental data given in mean ± standard deviation, with three specimens for Plate A and C, and two specimens for Plate B.

drop in slope. This indicates the onset of initial damage and Core damage then initiated at 1.18 kN, with yielding
a subsequent loss of global stiffness. Both the critical loads first occurring in the honeycomb cell that was located clos-
and the initial slopes are higher for Plate C (Table 4), which est to the indentor, at its six vertical edges and cellular walls
is consistent with earlier findings suggesting that an almost simultaneously. At this instant, delamination
increase in skin thickness increases the critical load [3,7]. quickly advanced toward the inner plies near the mid-
This may be explained by the increase in flexural rigidity plane, where the maximum shear stresses were found. This
and local contact stiffness for sandwich panels with thicker was followed quickly by fibre compressive failure in the top
facesheets. Consequently, the core becomes more shielded ply underneath the indentor and subsequently, the load–
from the contact stresses, and the onset of damage in the displacement curve exhibited a loss in the slope. As the
core happens at a greater load. The critical loads for all load increased, matrix cracks propagated extensively near
the plates are between 30% and 35% of the respective ulti- the bottom plies and also at the top surface where the face-
mate loads. sheet experienced further stretching due to the relatively
As the load increased, a number of minor load drops soft core. Fibre breakage also occurred at the bottom ply
occurred without significantly affecting the global stiffness and propagated in the inner plies. Around the maximum
of the specimen. A steepened slope towards the ultimate load of 3.5 kN, extensive fibre failure had occurred through
load was observed for Plate B, which was due likely to the depth in the localised region underneath the indentor.
the membrane stretching of the top facesheet. Membrane Core crushing was then observed to spread to seven cells.
effects were likely to be significant due to the small speci- Similar trends in behaviour were observed for Plates A
men size with a clamped boundary. Ultimate failure was and C.
characterised by a sudden major load drop, which always Although matrix failure and delamination at the top ply
coincided with an audible crack observed in the tests. This developed at an early stage of the loading, the loss in stiff-
suggests the cracking of the top facesheet, as also reported nesses for the damaged elements was not serious enough to
in [5]. affect the global stiffness of the structure. As the load
increased, the remaining undamaged stiffer elements had
5.2. Damage characteristics of sandwich panels under quasi- to carry a greater portion of the additional load. In addi-
static indentation tion, elements which had sustained matrix failure could
not carry any more load in the matrix, which led to an
In order to understand the nature of damage, specifi- increase in stress level in the fibre direction. This process
cally the initial damage, the onset of failure modes in the continued until core failure, delamination near the mid-
sandwich plate were identified and located using the FE plane of the top skin, and fibre compressive failure in a very
progressive failure model. For Plate B, the bottom plies small region contacting the indentor were observed to
of the top facesheet was subjected to high in-plane tensile occur either simultaneously or in a very quick succession
stresses since bending dominated in the initial stage of for the three plates. This abrupt loss in stiffnesses, com-
loading. Due to low transverse lamina strength, the first bined with the removal of elements due to fibre failure,
ply failure was matrix cracking in the bottom ply (ply 10) contributed to the decrease of global stiffness at the critical
of the top skin at 365 N, following which matrix crushing load. This result seems consistent with earlier findings that
initiated in ply 2 due to the high local contact stress. As reported a simultaneous occurrence of core crush and
the load increased, matrix crushing propagated through delamination at the onset of initial failure [7]. Moreover,
the depth of the skin in a localised region. A small delam- this also infers that the critical load can be predicted by
ination first occurred on the top ply around the indent determining the initiation of core failure.
point at approximately 785 N, which is similar to the
observation reported in [12]. However in a separate diag- 5.3. Results from the analytical model
nostic experimental test, where loading was terminated
once the initial threshold load was exceeded, no delamina- Table 4 lists the predicted elastic stiffnesses, critical loads
tion was detected on the top surface of the specimen. No and damaged stiffnesses for the three plates, compared with
effect on the load–displacement curve was observed up to the experimental values. The elastic stiffnesses and the crit-
this point. ical loads are derived from the equations detailed in Sec-
C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356 1355

tion 4 whereas the damaged stiffnesses are estimated by lin- gies of 3.6 J and 1.8 J respectively. In the load–time curves,
earising the slope of the load–displacement curves (Fig. 4). the load initially increases linearly up to the critical load
The predicted results are in good agreement with the exper- and then drops suddenly. This is followed by a subsequent
iments. In all cases, the number of damaged vertical edges increase to the maximum load at a lower stiffness. The pre-
in the core was six with the initial damage limited to the cell dicted results are comparable with test data, in terms of
nearest to the indentor, thus the elastic energy of the core peak load and overall profile. This implies that the quasi-
at the onset of initial damage was 0.30 J. The elastic energy static assumption adopted for the energy-balance model
absorbed by the facesheet in Plate C was the highest at is valid in our case here.
0.435 J, due to its thicker facesheet and higher flexural This approach demonstrates the capability of the simple
rigidity. Discrete modelling of the honeycomb core with modified energy-balance model to reproduce the low-veloc-
shell elements allowed the visualisation of core damage, ity impact response of a sandwich plate by using just three
and thus accurate prediction of failure initiation was parameters (K0, Pc and Kdam) derived from the static tests.
possible. For a low-velocity impact to be considered quasi-static,
Next, the impact model in Section 4.3 was used to derive the suggested upper limit for velocity ranges from 10 to
the impact response of the composite sandwich plates. 100 m/s, and a range of 10 to 20 m/s has been suggested
Figs. 5 and 6 depict the load–time histories and load-deflec- for typical composite materials [36]. Likewise, an impac-
tion histories for Plates A and B subjected to impact ener- tor-plate mass ratio of 2 is sufficient to ensure quasi-static
impact response; however, an impactor that has a mass
at least 10 times greater than its target is recommended
[37].

6. Conclusion

An approach is described to predict the low-velocity


impact response of composite sandwich plates after the
onset of damage. The behaviour of composite honeycomb
sandwich plates subjected to quasi-static loading with a
hemispherical indentor was first investigated experimen-
tally and numerically. A three-dimensional progressive
damage model was developed and incorporated into the
commercial FE software, ABAQUS, to analyse the static
indentation problem. Comparison with experimental
load–displacement curves demonstrated that the numerical
model was capable of modelling the behaviour of compos-
ite sandwich plates under quasi-static loading. In addition,
the load–displacement curve exhibited three characteristics,
Fig. 5. Load–time and load–deflection histories for Plate A under 3.6 J namely, the elastic stiffness, the critical load at the onset of
impact. initial damage, and the stiffness after damage. At the onset
of initial damage, the plates exhibited a simultaneous
occurrence of core yielding, onset of delamination in the
top skin, and fibre fracture in a small region contacting
the indentor. Further loading resulted in continued core
crush, propagation of delamination and fibre fracture,
which culminated in top-skin failure.
Next, an analytical approach has been suggested to pre-
dict the critical load at the onset of initial damage. The crit-
ical load was theoretically predictable by accounting for
the elastic energies absorbed by the core and the top face-
sheet up to the point of core damage. Since the energy
absorbed by the facesheet increases with increasing flexural
rigidity, this implies that the critical loads would be greater
for plates with thicker facesheets.
Finally, the energy-balance model was modified to
extend its validity beyond the elastic regime. The laws of
conservation of energy and momentum were coupled to
Fig. 6. Load–time and load–deflection histories for Plate B under 1.8 J predict the low-velocity impact response of sandwich plates
impact. using the stiffnesses and critical loads derived from the
1356 C.C. Foo et al. / Composites Science and Technology 68 (2008) 1348–1356

quasi-static analysis. The load–time history and load– [15] Meo M, Morris AJ, Vignjevic R, Marengo G. Numerical simulation
deflection history up to maximum load were well predicted of low-velocity impact on an aircraft sandwich panel. Compos Struct
2003;62:353–60.
in terms of peak load and overall profile, which infer that [16] Besant T, Davies GAO, Hitchings D. Finite element modelling of
the quasi-static assumption for low-velocity impacts is low-velocity impact of composite sandwich panels. Compos Part A –
valid here. The simple modified energy-balance model is Appl S 2001;32:1189–96.
shown to be capable of predicting the low-velocity impact [17] Aktay L, Johnson AF, Holzapfel M. Prediction of impact damage on
response of a sandwich plate by using the three character- sandwich composite panels. Comp Mater Sci 2005;32(3–4):252–60.
[18] Horrigan DPW, Aitken RR. Finite element analysis of impact
istics derived from the static tests. damaged honeycomb sandwich; 1998. http://www.lusas.com.
[19] Fatt MSH, Park KS. Dynamic models for low-velocity impact
Acknowledgements damage of composite sandwich panels – part a: deformation. Compos
Struct 2001;52:335–51.
We are grateful for the funding received from Agency [20] Shivakumar KN, Elber W, IIIg W. Prediction of impact force and
duration due to low-velocity impact on circular composite laminates.
for Science, Technology and Research (A*STAR) of Singa- J Appl Mech 1985;52:674–80.
pore and to the School of Mechanical and Aerospace Engi- [21] Malekzadeh K, Khalili MR, Mittal RK. Response of composite
neering, Nanyang Technological University of Singapore sandwich panels with transversely flexible core to low-velocity
for the use of the computing and laboratory facilities. transverse impact: a new dynamic model. Int J Impact Eng 2007;34:
543–52.
[22] Zhou DW, Stronge WJ. Low velocity impact denting of HSSA
References lightweight sandwich panel. Int J Mech Sci 2006;48:1031–45.
[23] Anderson T, Madenci E. Graphite/epoxy foam sandwich panels
[1] Abrate S. Impact on composite structures. Cambridge University under quasi-static indentation. Eng Fract Mech 2000;67:329–44.
Press; 1998. [24] Sburlati R. The contact behaviour between a foam core sandwich
[2] Hazizan MA, Cantwell WJ. The low velocity impact response of an plate and a rigid indentor. Compos Part B – Eng 2002;33:
aluminium honeycomb sandwich structure. Compos Part B – Eng 325–32.
2003;34:679–87. [25] Anderson T. An investigation of SDOF models for large mass impact
[3] Herup EJ, Palazotto AN. Low-velocity impact damage initiation in on sandwich composites. Compos Part B – Eng 2005;36(2):135–42.
graphite/epoxy/Nomex honeycomb sandwich plates. Compos Sci [26] Feraboli P. Modified SDOF models for improved representation of
Technol 1997;57:1581–98. the impact response of composite plates. J Compos Mater 2006;40:
[4] Wen HM, Reddy TY, Reid SR, Soden PD. Indentation, penetration 2235–55.
and perforation of composite laminates and sandwich panels under [27] Schubel PM, Luo J, Daniel IM. Low velocity impact behaviour of
quasi-static and projectile loading. Key Eng Mat 1998;141–143:501–2. composite sandwich panels. Compos Part A – Appl S 2005;36(10):
[5] Dear JP, Lee H, Brown SA. Impact damage processes in composite 1389–96.
sheet and sandwich honeycomb materials. Int J Impact Eng [28] Foo CC, Chai GB, Seah LK. Quasi-static and low-velocity impact
2005;32:130–54. failure of aluminium honeycomb sandwich panels. Proc IMechE Part
[6] Fatt MSH, Park KS. Dynamic models for low-velocity impact L: J Materials: Design and Applications 2006;220:53–66.
damage of composite sandwich panels – part b: damage initiation. [29] Hashin Z. Failure criteria for unidirectional fibre composites. J Appl
Compos Struct 2001;52:353–64. Mech 1980;47:329–34.
[7] Zhou G, Hill M, Loughlan J, Hookham N. Damage characteristics of [30] Tserpesa KI, Labeasb G, Papanikosb P, Kermanidisa Th. Strength
composite honeycomb sandwich panels in bending under quasi-static prediction of bolted joints in graphite/epoxy composite laminates.
loading. J Sandwich Struct Mater 2006;8:55–90. Compos Part B – Eng 2002;33:521–9.
[8] Karger L, Baaran J, Tessmer J. Rapid simulation of impacts on [31] Luo RK, Green ER, Morrison CJ. Impact damage analysis of
composite sandwich panels inducing barely visible damage. Compos composite plates. Int J Impact Eng 1999;22:435–47.
Struct 2007;79(4):527–34. [32] Petras A, Sutcliffe MPF. Indentation failure analysis of sandwich
[9] Feraboli P. Damage resistance characteristics of thick-core honey- beams. Compos Struct 2000;50:311–8.
comb composite panels. In: Fourty seventh AIAA/ASME/ASCE/ [33] Aminanda Y, Castanie B, Barrau JJ, Thevenet P. Experimental
AHS/ASC structures, structural dynamics, and Materials Conference. analysis and modelling of the crushing of honeycomb cores. Appl
Number 2006–2169;2006. Compos Mater 2005;12:213–27.
[10] Mines RAW, Worall CM, Gibson AG. Low velocity perforation [34] Foo CC, Chai GB, Seah LK. Mechanical properties of Nomex
behaviour of polymer composite sandwich panels. Int J Impact Eng material and Nomex honeycomb structure. Compos Struct
1998;21:855–79. 2007;80:588–94.
[11] Turk MH, Fatt MSH. Localised damage response of composite [35] Olsson R. Closed form prediction of peak load and delamination
sandwich plates. Compos Part B – Eng 1999;30:157–65. onset under small mass impact. Compos Struct 2003;59:341–9.
[12] Hou JP, Petrinic N, Ruiz C, Hallett SR. Prediction of impact damage [36] Olsson R. Mass criterion for wave controlled impact response of
in composite plates. Compos Sci Technol 2000;60:273–81. composite plates. Compos Part A – Appl S 2000;31:879–87.
[13] Kelly G, Hallstrom S. Strength and failure mechanisms of composite [37] Feraboli P. Some recommendations for characterization of composite
laminates subject to localised transverse loading. Compos Struct panels by means of drop tower impact testing. J Aircraft 2006;43(6):
2005;69:301–4. 1710–8.
[14] Huang CH, Lee YJ. Experiments and simulation of the static contact [38] Trappe V, Harbich KW. Intralaminar fatigue behaviour of carbon
crush of composite laminated plates. Compos Struct 2003;61:265–70. fibre reinforced plastics. Int J Fatigue 2006;28:1187–96.

Das könnte Ihnen auch gefallen