Você está na página 1de 54

1

A Review of Low-velocity Impact on Sandwich Structures


Chai, Gin Boay and Zhu, Shengqing
School of Mechanical & Aerospace Engineering
Division of Aerospace Engineering
Nanyang Technological University, Singapore
Abstract
This contribution is a review of past and current research progress on dynamic response of
composites sandwich structures subjected to low velocity impact. To obtain a comprehensive
overview of the current state-of-the-art published works, impact response on sandwich
structures are broadly classified into two main groups: high-velocity and low-velocity impact;
with the focus on low velocity impact. A summary of some of the commonly used theoretical
solutions for low-velocity impact is provided and described in some details. The response of
low-velocity impact has been determined based on mass ratio which is defined as the ratio of
the impact mass to effective mass of the structural component. As a result, the response under
low-velocity impact was further subdivided into three possible categories, namely, large mass
impact, small mass impact and medium mass impact. For each category of impact, currently
available suitable solutions are discussed in details. To classify the response of low velocity
impact and also to explore the energy absorption characteristics as well, the impact duration is
identified as the key parameter. The review concludes with detail discussions on the damage
mechanisms and failure criteria for sandwich structures subjected to impact loads.
Keywords: Impact damage, Composite sandwich panel, Dynamic response.
1 Introduction
Composite sandwich panels are generally made up of two thin but stiff facesheets or skins
separated by a lightweight and thick but low modulus core. Thus they are relatively light
weight and efficient in supporting transverse loads. Generally, the facesheets are made of light
alloys such as aluminium or fibre-reinforced composites while the core materials used in
structural applications include honeycomb cores made of Nomex, fiberglass reinforced
thermoplastic, aluminium, and foam type cores. The core thicknesses range between 3 and 26
2
mm [1-5]. This combination of materials resulted in a rather competitive structure in
application area where specific strength and stiffness are the main focus. Sandwich structures
are commonly found in many engineering applications such as the aerospace and aircraft
industries, marine industries and wind turbine industries. It is well known that most of the in-
plane membrane and bending forces are supported by the facesheets while the shear loads are
transferred by the core.

Although lightweight, composites and composite sandwich structures are susceptible to impact
damage, which may severely decrease the structural stiffness, stability and load carrying
capacity. Birds or debris from the runaway during aircraft takeoffs and landing are some of
common impactors. However, the damage suffered internally by the panel during impact is
very difficult to inspect [1-8].

The study of the effects of the impact load on plate structures has a long history of more than
200 years while the first 150 years were devoted almost exclusively to empirical studies of the
penetration. Simple semi-empirical or quasi-rational representations of this type of impact
began to appear towards the end of the 19
th
century. The first research work concerning in-
plane compressive loads in sandwich structures was the work by Marguerre in 1944 [8]. From
that point onwards, research on composites and composite sandwich structures boomed and
many books on sandwich structures were published. One of these books was written by Allen
in 1969 [4], which remained as one of the dominant reference in this field until today.

In modern times, composite sandwich panels subjected to impact loads have been widely
examined through the use of experimental techniques, mathematical modeling and numerical
methods. With the increasing use of composite sandwich structures in engineering application
and due to their relatively poor resistance to impact load, the following areas of research have
been and still are being pursued in ascending order [2-3]:
3

Contact analysis between the impactor and the composite surface,
The response of structure after impact,
The failure mechanism by impact,
The residual property after impact,
High-velocity impact analysis,
Impact analysis of soft body impactor.

In the past, composite sandwich structures were mainly restricted to non-critical secondary
structural components in aerospace application. The design of non-critical secondary structure
requires only the determination of the overall global response. The information of impact
response includes the transverse deflection, critical load and maximum load, fundamental
vibration frequencies and their associated mode shapes. The literature on the response of the
sandwich panel subjected to low-velocity impact is vast [2, 9-11]. Their use in recent times is
extended to primary structures in aerospace structures, ship superstructures, high-speed train
carriages and car bodies. Thus, a more accurate model is required to predict the actual
response and this review seems to reveal that there is still a lot of work to be done at this stage.
However the deductions using existing theoretical and numerical models for the impact
analysis of composite sandwich structures will be useful to both analyst and designer [2-3, 12-
14].

In this paper, the focus will be directed on the factors influencing the response, classification
of response as well as the failure modes of composite sandwich structures when subjected to
low-velocity impact. Up until now, experimental, theoretical and numerical methods were
applied to investigate the impact response of sandwich panels [2, 3, 15-18]. Among the
theoretical methods are the spring-mass model, the modal superposition method and the
energy-balanced method [2-3, 11, 18]. However over the last few decades, finite element
4
method of analysis has evolved into a powerful technique with the ability to simulate the
impact responses for various geometries and different impact load conditions [17]. However
the current easily available computational power is one of the reasons for the exponential
advancement of the use of finite element method in non-linear impact analysis.

At this stage in time, a structural designer or engineer faced an uphill challenge in selecting
the appropriate method of analysis to cater for the right impact response. It is hope that this
contribution will help to reduce some of difficulties faced in the selection process. From the
deduction of the influence factors on response, the response of composite sandwich structures
is divided into large mass impact, small mass impact and medium mass impact. For each
category of impact, currently available suitable solutions are discussed and compared in
details. Failure modes of sandwich panel in particular are also discussed and summarized
towards the end of this review paper.

2 Primitive Impact Factors and Composite Sandwich Parameters

The vast amount of literatures on the impact of sandwich structures can be grouped into the
impact response, damage resistance, damage tolerance and perforation. Pioneer workers made
great effort on the response of sandwich panel subjected to quasi-static and low-velocity
impact load. These theories were extended to dynamic response and were improved to deal
with the high-velocity impact [19-24]. Two key indices have been identified in the area of
foreign body impact on composites and sandwich structures; damage resistance and damage
tolerance. Damage resistance is a measure the ability of sandwich structures to resist damage
when subjected to foreign body impact and damage tolerance is a measure of the ability of
sandwich structures to perform satisfactory with impact damage and without any reparation.

With regards to research on the low-velocity impact load response of sandwich panel,
understanding in certain areas has matured. For example, most researchers adopted the general
5
approach to model the local indentation of a composite plate using a smooth impactor[6-7].
And it is widely agreed that the response is dependent on the following parameters, with
reference to Figure 1 [3, 7, 9-11, 25-27]:
1) Geometry parameters such as skin thickness
f
h , thickness of the core
c
h , support span
p
R , the core type, impactor or indentor radius
i
R , and the radius of contact
c
R .
2) Material properties such as Young's Modulus of the facesheet E , Shear Modulus of the
core
c
G and Youngs Modulus of the impactor
i
E .
3) Impact velocity
0
v .
4) Mass of the impactor M and mass of the sandwich structure
p
M .

This review begins with discussion on the contact analysis, the deflection of the impact
response and several existing theoretical and numerical models. Damage mechanisms of
composite sandwich panel and the various modes of failure will be subsequently discussed
towards the end of this contribution.

In order to optimise the stiffness of sandwich panels, the skins are generally made of strong
materials such as light metals or composites. The skins material should be lightweight but high
in modulus and failure stress. The mechanical data of the skins are elastic modulus E
f
,
Poissons ratio and the thickness of the skin
f
h . The core is usually made of lightweight
material with relative low modulus and low strength. It is glued together with skins in both
sides, which produces complex interaction between them during deformation. The weakness
of the relatively soft core may be the cause of damage initiation in the sandwich panel [9, 11,
28-30] but debonding of the interface between the core and the skins may also be the source of
initial damage.

6
Early research work on honeycomb sandwich structures treated multi-cellular core as an
equivalent continuum with effective engineering properties. These methods ignored the
discontinuous surfaces between the core and skins, and thus failed to reveal the exact damage
progression. The existing impact models, such as the spring-mass and energy-balance models,
assume elastic behaviour for the core, thus they ceased to be valid after the onset of damage.
The finite element simulation can model geometrically correct core using shell elements to
obtain a more realistic distribution of stress and strains. The mechanical data of the core in use
are shear modulus
c
G , Poissons ratio
c
and height of the core
c
h .

The crushing capacity of a bare honeycomb core is illustrated in Figure 2 [3, 6, 31-33]. The
compressive stress of the core increases linearly with increasing compressive strains in the
initial stage until the stress reaches the critical value
max
o and then decreases sharply. After
that, the stress in the core approximately remains constant at
0
q with strain increasing. At the
end stage, the core is crushed to a state of densification and therefore the stress increases
sharply with increasing compression.

3 Contact Response

Hertz law is a well known theory for contact analysis of isotropic bodies[34]. For most
sandwich structures in practice, the penalty of having a low modulus core would means that
the local deformation of the core dominates the total transverse deflection. Moreover, the
thicknesses of the skins and the height of the core are quite different while the modulus varies
from the skin to the core, which could not be accounted for in the Hertz law. It has been
shown that the traditional Hertz law has to be modified to cater for sandwich structure with a
lightweight core. The modified Hertz law has included the contact characteristics of
composites especially when the upper skin is made of composites [3, 11, 35].

7
According to Hertz law[34, 36], the contact area of a spherical shaped impactor in contact
with a flat plate is circular and as shown in Figure 3 the distribution of contact pressure p is :
2
0
1 ( )
c
r
p p
R
= (1)
where
0
p is the maximum contact pressure in the centre of the contact zone, and are the
radial position of an point in contact zone and the contact radius, respectively. The force
indentation law is expressed as
q
h
P k o = (2)
where P ando are the contact force and indentation; q is a constant and k
h
is the contact
stiffness for composites dependent on two parameters R
-
and E
-
as [37]
* *
4
3
h
k E R = ;
*
1 2
1 1 1
R R R
= + ;
1 2
*
1 2
1 1 1
E E E

= + (3)
where
1
R ,
2
R are the curvature of the two bodies and
1
E ,
2
E ,
1
,
2


are their Young Moduli
and Poissons ratios, respectively. For the contact between the impactor and the sandwich
panel, the radius of the contact area is given by [11, 38],
c i
R Ro = (4)
The profile of the local indentation of the upper face-sheet is given by [9, 11],
( )
2
2
0 2
(1 ) r
a
r
o o = (5)
where ( ) r o and
0
o are the indentation of the point and the maximum indentation and a is the
radius of the region of local indentation. It should be noted that Eqn. (5) is only valid for
quasi-static model and needs to be modified for high velocity impact, which will be discussed
in some details later.

Extensive experimental results indicate that the modified Hertz Law yields good results in the
initial stage of impact but the contact area increases dramatically and then deviates
8
significantly for larger size of indentations [9-10, 26, 28, 39-43]. Eqn. (2) demonstrated the
relationship of Hertz law but its application is limited to elastic contact analysis. Olsson [36]
improved the contact law for contact of spherical impactor with a specific type of plate and
deduced that in eqn. (2);
3
2
q = for composite plate and 1 q = for sandwich panel. These
values were extensively validated via experiments on certain type of sandwich panels [27, 30].
It must be cautioned that the coefficient
h
k

in eqn (2) for practical sandwich panels should be
determined and thoroughly validated using an extensive experimental study [10, 26-27, 30,
38].
4 Contact and Impact Duration
The impact response has been known to be highly dependent on the contact duration in an
impact event. The duration of impact then depends on the time for stress wave to travel to one
boundary. However, impact duration is rather difficult to quantify and as such the mass
criteria was developed. With extensive experimental investigations [44-45] and sophisticated
mathematical modelling, explicit expressions for impact duration has been published [18, 26,
46]. But none of the work cited so far can be used for all types of impact. Ref. [7] proposed an
expression to roughly estimate impact duration,
c
bs
M
t
K
t = (6)
where M is the mass of the impactor and
bs
K is the combined bending-shear stiffness of the
target. Impact duration was also estimated from the empirical relationships between the
contact force and the indentation as [18],
0
1.7
c
t
v
o
~ (7)
Olsson [26, 47] published a theoretical model for the impact response for wave-controlled
impact but this model could not explain why the impact duration goes to infinity in some cases
9
[28, 47]. Assuming an elastic body is impacted by spherical projectile, the impact duration can
be expressed as [45]
1 2
5
2
0
3.2145( )
c
h
M
t
v k
= (8)
The impact duration will be discussed further in this review.
5 Deflection of Sandwich Panel Subjected to Impact Loading
The deformation of a sandwich panel subjected to impact load can be divided into two parts :
f w o = + (9)
where w the global deflection ando is the indentation. For low-velocity impact, the global
deflection w depends on the boundary conditions and highly dependent on the impact duration.
If the impact duration is long enough, global deflection will be in the phase with the motion of
the impactor and thus dominates deformation. On the other hand, if the impact duration is too
short, indentation will dominate the deformation. This is especially true for high-velocity
impact and small mass impact.

Based on Kirchhoffs plate theory for orthotropic composite plate without damping, the
central deflection of a plate subjected to a central concentrated load is given by [28]
( )
0
1
( )
8
t
w t P d
D
t t

=
}
(10)
where the mass per unit area of the plate, 2
f f c c
h h = + and D is the flexural rigidity of the
plate. Neglecting the effect of gravity, the acceleration of the impactor is ( ) / P t M and the
displacement of impactor can be written as
( )
( )
0
1
( ) ( )
0 0; (0) .
t
o
o
f t v t t P d
M
f f v
t t t

' = =

}
(11)
Combining eqns. (9), (10) and (11) gives the indentation as :
10
0 0
1 1
( ) ( ) ( ) ( )
8
t t
o
t v t t P d P d
M D
o t t t t t

=
} }
(12)
Differentiating eqn. (12) twice with respect to time, the governing equation for indentation
with contact force yields:
( )
2
2
1
0
8
0 0, (0)
o
d dP P
dt dt M D
v
o

o o

+ + =

' = =

(13)
The relationship of contact force and the indentation will be discussed further in the next part.
6 Classification of Impact Response
This section will begin with the classification of the various impact responses and the
conditions that govern these responses as illustrated in Figure 4. From extensive experimental
and numerical studies, impact velocity is known to be an important factor in assessing the type
of impact response. High-velocity impact usually brings about larger transverse deflection,
more damage and even perforation. Hence the type of impact response is cursorily divided
into low-velocity and high-velocity impact according to their impact velocity.

Presently, there are two criteria that are used to distinguish low-velocity impact from high-
velocity impact. One is based on the structural deformation and damage [43, 48] while the
other on the structural response [3]. Stronge [48] elaborated that the amount of plastic
deformation (or damage) near the contact zone governed the difference between low-velocity
impact from high-velocity impact. At low-velocity impact, the plastic deformation is localised
around the contact area. In contrast, high-velocity impact leads to a larger area of deformation
or damage around the area of contact. Note that, for some cases, these two criteria gave the
same impact response.

Early literature review by Bland [25] showed that the impact mass should be taken into
consideration jointly with the impact velocity. Furthermore, early works also showed that the
11
impact response and damage may be quite different for impact of the same velocity and
different impact masses [2-3, 11, 26, 49]. Besides impact velocity and impact mass, impact
duration is also a key parameter that distinguishes high-velocity impact response from low-
velocity impact response. Based on the impact duration, t , the impact response can be divided
into three classifications [26, 43] :
1) When the impact duration
/
~
E
h
t
c
is of the order of the transition time for the
stress wave to travel through the thickness, the response is dominated by 3-D wave
propagation. The analysis for this kind of response usually involved the consideration
of stress wave propagation.
2) When the impact duration
1
~
/
p
R
t
E
is significantly longer than the time taken for
the flexural and shear stress wave to travel to the boundary. This response is illustrated
in Figure 5(b) and is usually analyzed using Modal Superposition Method.
3) When the impact duration
1
/
p
R
t
E
>>

is much larger than the time needed for the
stress wave to reach the boundary, the lowest vibration mode of the impactor-plate
system governs the impact response as shown in Figure 5(a). This kind of response can
be modelled as a quasi-static analysis and the energy balance model is commonly used
for this type of response.
In above expressions,
1
/ E is the stress wave velocity, and the stress wave velocity for
composites is [41, 43] :
12 21
1 12 21 32 23 13 31 21 32 13
(1 )
(1 2 )
f
composite
E
v

=

(14)
where
f
E ,
1
,
ij
are the elastic modulus, density and Poissons ratio of the face-sheet. A note
for consideration is that the impact duration t is difficult to quantify due to the high
nonlinearity of the response, especially for the first two classes of impact.
12

For the response of the first class of impact, the wave velocity is usually more than 1000 m/s
for most solid material [13, 19-20, 22, 24, 32, 50]. Typically the thickness of the panels
generally varies from 6 mm to 20 mm which means the transition time for the stress wave to
travel is less than 20 s. The class of impact would often results in perforation or visible
damage and thus will not be discussed further in this review.

The response of the third class of impact has been extensively explored as it can be analyzed
using a quasi-static method [9, 11, 33, 51-52] and will be reviewed in some details later. The
analysis for the response of small mass impact is also available [10, 26, 28] but still has some
room for improvement [41, 53].
6.1 Stiffnesses of a circular sandwich panel
The stiffness of the target is a crucial parameter and is needed to complete the analysis of an
impact event. The stiffness determines the effective mass of panel which in turn affects the
dynamic impact response. In this section, the various stiffnesses of the target panel are
reviewed and analyzed. For a circular isotropic or quasi-isotropic plate, Table 1 lists the
available expressions for the bending stiffness, shear stiffness and membrane stiffness [10-11,
46]. The flexural rigidity of an orthotropic plate is given as
11 21
(1 )
2
D D A
D
+
=
(15)
where
12 66
11 21
2 D D
A
D D
+
= and
ij
D are the laminate stiffness. The combined stiffness
bs
K of plate
in transverse loading condition can be determined as :
1
1 1
1
b s
bs b
b s
b s
K K
K K
K K
K K

= = =
+
+
+

(16)
13
where is defined as the ratio of the shear stiffness to bending stiffness. The expressions for
the bending and shear stiffnesses are compiled in Table 2 :

( )
2
1 (3 )
4 1 2ln( / )
p
s c
b f p c c f
R
K G
K E R R h h

+
= =
+

(17)
For thin isotropic plates, deformation due to shear effect can be ignored since the bending
stiffness is far less that shear stiffness [54-55]. For sandwich panel with soft core, the shear to
bending stiffness ratio is dominated by the ratio of thickness to span and is less affected by
the moduli and thicknesses of the core and facesheet. When the span is much larger than the
thickness, the combined stiffness
bs
K in eqn. (17) is essentially dominated by the bending
stiffness
b
K .

For thin isotropic plate or Kirchhoff plate with span to thickness ratio that exceeds 10, its
shear stiffness is much larger than its bending stiffness. However for sandwich structures with
a low modulus core, the shear-bending stiffness is affected by both the span to thickness ratio
and the low modulus of the core. The variation of the shear to bending stiffness ratio for a
circular panel with fixed supports is numerically plotted with 0.3 = in Figure 6. The figure
shows that the shear stiffness is more than 10 times of the bending stiffness when span to
height ratio exceeds 25, which means that the panel can be treated as a Kirchhoff plate.

The contact stiffness
h
k for composite laminates can be determined using the well known
semi-empirical relation of eqn. (2) [35, 37]. However, to date, the contact stiffness
h
k for
sandwich panel is derived purely from experimental data [47, 53].
6.2 Large mass impact
The impact duration is used as a basis to judge whether the impact response is boundary-
controlled [3, 11, 52]. Even for low velocity impact, the impact duration is measured in
14
microsecond or millisecond and is therefore rather difficult to quantify using experiment.
Neither can it be predicted exactly using theoretical method. As mentioned earlier, for low-
velocity impact, velocity is not a key parameter to gauge impact response. Thus the impact
response are classified using other factors.

Based on mathematical solutions and experimental results, it is widely accepted that besides
impact velocity, impact mass is a key parameter that distinguish the various response
categories [25]. The ratio of the impact mass to panel mass is of great importance to separate
the responses. It has been pointed out by Olsson [26] that the impact response for low-velocity
impact is governed by the ratio of the impact mass to the plates mass. When the impact mass
is larger than twice the plates mass, it is termed as large mass impact by Olsson and the
impact response can be solved using the quasi-static approach of analysis. When the impact
mass is less than 1/5 of the plates mass, it is termed as small mass impact and the solution is a
dynamic impact. On the other hand, impact on plates with large aspect ratios (length/width>3)
is deemed as boundary-controlled impact because the deflections primarily affect only a
fraction of the plate. Since the response of large mass impact is boundary-controlled, the
impacts on plates with large aspect ratios are classified into large mass impact.

Several studies [11, 18, 26, 28] have deduced that the ratio of the impact mass M to effective
mass of the plate
eq
m should be the key criterion in determining the type of impact response. A
dimensionless quantity defined as mass ratio
eq
M
m
m
= is commonly used in the literature. It is
widely accepted that the impact response is quasi-static when the mass ratio 8 m > and this is
the large mass impact defined by Olsson[26]. For impact on circular panels, Stronge[11] used
the ratio
* eq
p
m
m
M
= which is the ratio of the effective mass
eq
m

to the mass of sandwich panel
15
p
M . The ratio
*
m is highly dependent on the energy balance of the stiffnesses of the panel
and is derived as :
1
2
1
*
2
1
(1 ) ln
1 ln
( 1) (3 ) n
1
l 1
s
s s
m s s ds
s


(
| | +
= + +
( |
+ +
\ .

}
(18)
where 1
2
1
<<
|
|
.
|

\
|
=
p
c
R
R
s . When the span length
p
R is much larger than thickness of the panel
and , the ratio
*
m

approaches a maximum value of :
2
*
2
(7 56 121)
( )
54( 3)
max
m

+ +
=
+
(19)
It can be seen that
*
max
(m ) in eqn (19) is not significantly affected by . Varying the value of
from 0.2 to 0.5 resulted in
*
max
(m ) 0.24 s . This is generally the usual condition for a
Kirchhoff plate. When the boundary distance is short enough and 0 , the ratio
*
m

approached a minimum value of :
*
2
1
2
( )
[1 2ln ]
min
m
s
=
+
(20)
Typically the value of
1
0.01 s < , eqn (20) yields
*
min
(m ) 0.02 < . The variation of
*
m with
increasing for
1
0.001, 0.005, 0.01 s = is shown in Figure 6. As can be seen from the figure,
the value of
*
m

increases dramatically for the range of the shear to bending stiffness ratio
4 < . But varies little for 6 > and the curve plateau around
*
m =0.22. For most practical
sandwich plates subjected to low velocity impact, the effective mass
*
m is less than
1
4
. Note
that more kinetic energy will be transferred to the target panel for a larger effective mass. This
means that the Kirchhoff plate performs better in energy absorption during impact.

Swanson[52] pointed out it that for composite plates subjected to impact load, the quasi-static
solution is only valid when the mass ratio 8 m > using the following rule-of-thumb :
16
1
3
i n
e e < (21)
where
i
e is the impact frequency in impact while
n
e is the fundamental frequency of the
panel. This means that the criteria of Olsson[26], Stronge[11] and Swanson [52] are unified.

It is known that the criterion of effective mass ratio 8 m > is reasonable one but somewhat
inconvenient to apply especially for large mass impact. The guideline suggested in this
contribution for large mass impact if any one of the following conditions are satisfied :
I. The impact duration is much larger than the time for flexural wave travelling to the
boundary.
II. The effective mass ratio is larger than 8, i.e. 8 m > .
III. The ratio of the impact mass to the mass of the panel is larger than 2, i.e. 2 m> .
IV. The aspect ratio of the plate is larger than 3: length/width > 3.

This large mass impact response is dominated by the quasi-static deformation of the plate after
the flexural wave has reached a boundary. Since the impact duration is long enough, the
deformation will be in-phase with the contact load. This response is usually modelled using
either the spring-mass model, the energy-balance model or boundary-controlled solution [6, 9,
11, 18, 25-27, 41, 49, 56-60].
6.3 Small mass impact
For low-velocity impact, the impact response will be dynamic one for small mass impact. This
is because the impact event finishes before the first flexural wave reaches the boundary of the
structure. This means the boundary conditions will not affect the impact response and the
model can be analyzed as one with infinite boundary. Olsson[26] mentioned that for small
mass impact, the ratio of the impact mass to plates mass should be less than 0.29 for circular
plate and 0.23 for square plate. He further added that small mass impact is typical of a wave-
17
controlled impact response when the impact mass is less than 1/5 of the mass of sandwich
panel. The response of this impact can be modelled using the modal superposition method,
and Olsson published a closed form solution with more details in Ref. [26, 28, 50]. Some
experimental work on the measurement of deformation due to small mass impact using pulsed
laser holography and other advanced optical instrument have now surfaced in the literature[26,
28].

The shape of the affected area due to impact is circular for isotropic plate while it is elliptical
for composites as shown in Figure 7. The aspect ratio of the ellipse is found to be related to
the flexural stiffnesses along the principal directions for composite plates [47]

1/ 4 11
22
( )
x
y
r D
r D
= (22)
It is widely assumed that the shear effect can be ignored for small mass impact [10, 28, 43, 47,
61]. From eqns (2) and (13), the governing equation for indentation is :
( ) ( )
1
0
0
8
0 0, 0
q q
h h
qk k
M D
v
o o
o o

o o

'' ' + + =

' = =

(22)
For composites,
3
2
q = , the governing equation can be rewritten as :

( ) ( )
1 3
2 2
0
3
0
16
0 0, 0
h h
k k
M D
v
o o
o o

o o

'' ' + + =

' = =

(23)
Introducing the non-dimensional form of indentation
T v
0
o
o = and time
t
t
T
= ; where T is a
constant with the dimension of time and can be obtained from the Hertz contact of eqn. (2) :
1 2
5
2
( )
o h
M
T
v k
= (24)
Eqn. (23) can be thus be written in a dimensionless form :
18

( ) ( )
2 0
0 0, 0 1
o oo o o
o o

'' '
+ + =

'

= =

(25)
where the coefficient
1
2 3
5
0
3( )
16
h
v k M
D

= .
For sandwich panel, 1 q = , the governing equation can be re-written as :
( ) ( )
0
8
0 0, 0 1
h h
k k
M D
o
o o

o o

'' '
+ + =

'
= =

(26)
In non-dimensional format, Eqn. (26) can be written :

( ) ( )
2 0
0 0, 0 1
o o o
o o

'' '
+ + =

'

= =

(27)
where
1
,
16
h
h
Mk M
T
D k

= = .
It is observed that for impact response on composites, the time T is dependent on the impact
velocity but it is not so for sandwich panel. Eqn. (25) is non-linear and can only be solved
numerically. Eqn. (27) however is a homogeneous linear differential equation which can be
solved as :

2
2
2
2
( 1 )
( 1)
1
( 1)
( 1 )
( 1)
1
t
t
t
sin t
e
te
sinh t
e


<

= =

>

(28)
Expanding the non-dimensional parameters in eqn (28), the indentation can be obtained as :
19
2
0
/ 2
/
0
2
0
/ 2
( 1 / )
( 1)
1
( 1)
( 1 / )
( 1)
1
t T
t T
t T
sin t T
v T
e
v te
sinh t T
v T
e


<

= =

>

(29)
The contact force is thus given as
2
0
/ 2
/
0
2
0
/ 2
( 1 / )
( 1)
1
( ) ( 1)
( 1 / )
( 1)
1
h
t T
t T
h
h
t T
sin t T
k v T
e
P t k v te
sinh t T
k v T
e


<

= =

>

(30)
It can be observed that putting 1 > in eqn (28), the impact duration will be infinity which is
unreal. The solution is however finite for 1 < and thus subsequent discussions in this
contribution are applicable only for 1 < . The contact force is [26, 28] :
( )
2
2
0
/ 2
( 1 / )
1
h
t T
sin t T
P t k Mv
e

(31)
Introducing non-dimensional format of the contact force as :
2
2 2
0
( ) ( 1 )
1
( )
t
h
P t sin
k M e
t
v
t
P

= =

(32)
Differentiating right-hand side of eqn. (32) with respect to t yields :

1 2
2
sin 1
1
max
h
M
t
k

or
1
2
cos
1
max
h
M
t
k

(33)
Substituting this result into eqn (29) gives the maximum indentation
2
0
max
2
2 1
1
h
M
v
k
e
t

.
The variation of the non-dimensional force with increasing time and for different value of
is shown in Figure 9. It can be seen that for 0 = , the curve is symmetric and the response is
elastic. And for 0 1 < < , the response is inelastic. It can be anticipated here that larger will
leads to larger indentation or damage area/depth in sandwich panel.
20

Substituting eqn. (31) into (10), the overall displacement of the panel can be obtained as :
( )
2
0
2
1
[1 sin ]
8
1
t
T
Mv e
w t t arccos
T D

| |

= + |
|

\ .
(34)
and in non-dimensional form, the deflection and contact force are :
( )
2
2
sin 1
( ) 2 [1 ]
1
t
t arccos
w t
e

+
=

(35)
( ) ( )
2 2
2 2
sin 1 sin 1
( ) 2 [1 ]
1 1
t t
t arccos t
f t
e e


+
= +


(36)
The curves of the non-dimensional deflection vs contact force are shown in Figure 10 with
0.6 = and 0.3 = . And it can be observed from the figure that the response is highly
dependent on the coefficient . The coefficient is a key parameter in confirming the
response of small mass impact, and it is independent of the boundary conditions as can be seen
in eqns. (25) and (27).

6.4 Distinction between large and small mass impact response
Table 3 gives a brief summary of the three main categories of impact responses, their
associated time durations and damage descriptions. The damage by ballistic impact is
perforation or large deformation which is visible while the damages by others are barely
visible and difficult to detect. The characteristics of small mass impact (wave-controlled
impact) and of large mass impact (boundarycontrolled impact) has been extensively
investigated [3, 22, 26, 49, 62]. The following describes the difference between large and
small mass impact :
1) Large mass impact response is highly dependent on boundary conditions while small
mass impact is not too significantly affected by the boundary conditions.
21
2) The deflection and strains are out of phase for small mass impact while they are more
or less in phase during large mass impact. In general, small mass impact causes
localised deformation near contact area while large mass impact results in global
deformation.
3) The impact duration for small mass impact is much shorter than that for large mass
impact. This implies small mass impact gives a higher impact load and causes earlier
damage initiation for the same impact kinetic energy. For composites, small mass
impact tends to cause much larger delaminations than that of large mass impact with
the same kinetic energy.
4) The response of large mass impact is quasi-static while it is dynamic for small mass
impact.
5) There are usually two or more peaks in the contact force history for small mass impact
but only one peak for large mass impact.
6) The degree of indentation is not conspicuous in both small mass and large mass impact.

The work on impact response and damage resistance of composite sandwich panel has been
studied extensively using experiments, theories and numerical analyses [29, 59, 63-65].
Experiments play an important part in the study of impact response especially for composites
and sandwich structures. The fundamental mechanical properties of the constituents of the
sandwich structure such as the core and the facesheets must be accurately determined. Flat-
wise compression tests for bare honeycombs and basic mechanical tests for facesheets are the
common experiments. Quasi-static tests for composites and sandwich panels are usually
conducted to observe the initiation of damage, maximum load, maximum energy absorbed and
extent of damage. The impact tests are usually conducted using either the standard drop
weight or the dart test for low velocity impact, and the gas-gun equipment for high velocity
impact.
22
7 Methods of Solution
This section gives a description of the various methods of solution for the various types of
impact response. The common methods of solution used extensively in the literature are the
spring-mass model, modal superposition method and energy-balanced method.
7.1 Spring-mass model
Spring-mass model is easy to understand and provides an accurate solution for boundary-
controlled impact [2-3]. The model consists of two mass, i.e. the mass of the impactor, M, and
the effective mass of the plate
ef
m

with displacement
1
x

and
2
x , respectively, as shown in
Figure 9. The impactor and the plate is connected by a nonlinear spring,
c
K , and the plate is
connected to the ground by three springs;
m
K representing effective membrane stiffness,
s
K
representing effective shear stiffness and,
b
K representing effective bending stiffness. The
shear and bending stiffness are connected in series thus resulted in a combined stiffness
bs
K .
From the free body diagram of Figure 11, the equations of motion for the system can be
expressed as :
1
3
2 2 2
0
0
eq bs m
Mx P
m x K x K x P

''
+ =

''
+ + =

(37)
with initial conditions:

( ) ( ) ( ) ( )
1 1 2 0 2
0 0 0, 0 , 0 0 x x x x v
'
=
'
= = = .
It is well-known that the contact force P is highly nonlinear and the indentation is defined as
1 2
x x o = . The well known eqn (37) can be solved numerically.

The kinetic energy of a circular plate is [11] :
1
0
1
2
2
p
R
KE w rdr t ' =
}
(38)
23
where is the mass per unit area of the sandwich panel, w' is the velocity of each point on
the plate and,
ef
m

is the effective mass of the panel. The kinetic energy can also be written as
2
2
( )
1
2
ef
KE m w
'
= (39)
where w' is the velocity of mass
ef
m after impact. Equating the two expressions of the kinetic
energy shows that the effective mass of the plate is less than 1/4 of the plate mass[11].
Assuming that the impact energy is large enough, then the deflection will be much larger than
the local indentation which gives
1 2
0 x x = . Therefore the 2-DOF (degree of freedom)
system can be simplified into a single-DOF system as :
( )
3
1 1 1
0
ef bs m
M m x K x K x
''
+ + + = (40)
From static analysis, the respective global bending and shear deflection of a circular plate
subjected to a concentrated force in the centre is given by [66] :
( )
2 2
2(1 )
[1 ( ) ( ) ]
(3 )
b
b p p p
P r r r
w r ln
K R R R

+
= +
+
(41)
( )
( ) ( )
2
2
for 1
1 2l /
,
n
c s
s
c p c
P r
r R w r
K
R R R
(
(
s =
(
+

(42)
( )
( )
( )
2ln /
for ,
1 2ln /
p
c
c p
s
s
R r
P
r R w r
K R R
> =
+

(43)
7.2 Modal superposition method
Rather than representing the system with two lumped mass, the modal superposition method
treats the impacted structure as a continuous infinite DOF system. This method is usually used
when the velocity is not high and yet quasi-static method cannot be used. An impact response
from small impactor mass is a good example of this type of impact response. The governing
equation of a circular sandwich panel is given by [3, 11] :
) ( ) ( ) 1 (
2
2
2
t P t
S
D
t
w
w
t S
D
w D o
V
=
c
c
+ V
c
c
V (44)
24
where
2
1
r r r
c c
V = +
c c
is the Laplace operator, w denotes ( , ) w r s and
2
( ) /
c f c c
S G h h h = + .

This impact response is different from quasi-static response especially when high natural
modes of frequencies of the panel are excited, and also when the mass ratio is small. The
impact frequency is compared to the natural frequency in later section. When the impact
frequency is close to the natural frequency of the plate, the modal superposition method has
been shown to give a reasonably accurate solution. In practice, only the first one or two modes
are significant in a dynamic impact response analysis. The usual relationship between the
contact force and the indentation of eqn (11) is still valid here.
7.3 Energy-balanced method
Based on the assumption of quasi-static behaviour of a low mass structure subjected to a
heavy mass impactor, the energy-balanced method utilizes the principle of conservation of
total energy of the impactor-structure system. This energy model yields only the result of
maximum impact force, and not the entire force-time history. Because of the assumption, this
method is only valid for large mass impact as the response is boundary-controlled.

Mathematically, the kinetic energy of the impacting mass is equal to the sum of the various
energies; contact energy
c
E , membrane energy
m
E , and the combined bending and shear
energy
bs
E . Applying the conservation of energy gives :
2
0
1
2
c m bs
Mv E E E = + + (45)
The energy for indentation as a result of contact between impact and structure can be
expressed as :
0 0
3
0 0
( ) ( )
c
E Pd t t d
o o
o _ o |o o = = +
} }
(46)
25
Assuming quasi-static response, the relationship between contact force P and the overall
deformation of the plate w

can be shown to be :
3
bs m
P K w K w = + (47)
The energy for membrane and for bending-shear is given as
4
0
1
4
m m
E K w = ;
2
0
1
2
bs bs
E K w = (48)
Substituting eqns. (47) and (48) into (45) yields :
0
2 2 4 3
0 0 0
0
2 2 4 ( ) ( )
bs m
Mv K w K w t t d
o
_ o |o o = + + +
}
(49)
where
0
w and
0
o are the maximum deflection and indentation of the panel, respectively.

For large mass impact, the relationship between P and o was deducted by the principle of
minimum potential energy [11, 18]
3
P _ o |o = + (50)
where
0
16
3
f
D q
t
_ = and
2
0.488
f
h
| = . Based on quasi-static assumption, the force-
indentation relation in eqn (50) can be simplified by considering that the magnitude of
indentation is roughly of the same order as the thickness of the facesheet, i.e.
2
( ) 1 t |o ~ . Re-
writing eqn (50) as [11] :
1
4
1
( ) ( ) ( ) ( 2 ) ( )
( )
P t t t t
t
o |o _| o
o
= + ~ (51)
where the coefficient
1
0
4
19. 2 8
f
f
D q
h
_| = . This relationship between the contact force and
indentation is in the process of being studied and validated using experiments. According to
eqns (49) and (51), a nonlinear relation between the deflection and indentation can be obtained
as :
3 3
0 0 0 0 bs m
K w K w _ o |o + = + (52)
26
Introducing the non-dimensional quantities of

0
0
f
w
w
h
=

and

0
0
f
h
o
o = into eqn (52) to give :
3 3
0 1 0 2 0 0
0.488 w k w k o o + = +

(53)
where

2
1
f m
bs
h K
k
K
= and

0
2
16 /
3
f
bs
f
D q h
k
K
t
= . The maximum deflection
0
w can be solved
numerically using eqn (53). And subsequently eqns (51) and (52) are used to determine the
contact force and the indentation. Alternatively, the maximum deflection and maximum
indentation can be investigated by solving eqn (53) numerically. The nonlinear relationship of
maximum deflection with maximum indentation is shown in Figure 12 for various stiffnesses
of the plates. Using this method of analysis, there are two limitations; 1) it is impossible to get
an explicit expression for this solution as the equations are highly nonlinear and; 2) the contact
history is not available.
7.4 Finite Element Simulations and Results
It is a common fact that impact experiments are very time consuming and sometimes
expensive, and also they are usually destructive. Thus important time-based information such
as internal stresses and strains are rather difficult to capture. The engineering application of
finite element method has boomed over the years as computing power increased
exponentially. The finite element method has been increasingly used in predicting dynamic
response and modes of failure of structures subjected to impact. The use of finite element
method in providing critical information of stress/strain, wave propagation, energy partition
absorbed in the skin and core have been published [21, 57, 67-71, 97].
7.4.1 Accurate Geometric and Material Finite Element Model
Although a fair amount of success has been achieved in predicting low-velocity impact
damage in monolithic carbon reinforced composite panels [59, 68, 70, 108, 115], less work
were published on composite sandwich panels consisting of composite facesheets supporting
27
Nomex honeycomb cores. One of the major drawback of finite element method is that the
computational expenses for honeycomb sandwich models increase rapidly as the number of
cells in the core increases. Figure 13(a) shows a finite element model of a geometrically
accurate 9-cell honeycomb core subjected to transverse compression, and Figure 13(b) shows
the corresponding experimental failure mode.

In order to increase computational efficiency, the honeycomb core is sometimes meshed with
3D solid elements to represent an equivalent continuum model [104-107, 116], while 2-D
plate or shell elements are used for the skins [104, 107]. However, since the contact load
distribution in the laminate is inherently a 3D problem, the 2-D elements may prove
inadequate when transverse stresses are required for failure analysis. By considering the
honeycomb as a homogeneous material, the response is then described in terms of
macroscopic stresses and strains. Instead of considering the real cellular structure, the
sandwich core is analysed in terms of its effective properties, which have to be determined
using either results of mechanical testing or of analytical approximation. Various analytical
techniques have been proposed to predict the effective continuum properties of the core in
terms of its geometric and material characteristics [109, 110, 117120].

To model damage in the continuum core, several approaches have been proposed. Some
researchers have applied an elliptic yield criterion which accounts for the transverse normal
and shear stresses to predict the onset of core plasticity [105, 113]. Similarly, the core has also
been idealised as an isotropic, elastic-perfectly plastic material in the transverse direction [104,
106, 122]. Atkay et al. [106] proposed an element elimination technique which removes finite
elements upon reaching a threshold stress or strain value. However, as they pointed out [121],
such an approach cannot model a realistic impact response since failed elements also
contribute to the damage resistance even after failure. Horrigan et al. [114] proposed an
isotropic continuum damage model, but this modelling was limited to small deformations
28
since the plastic stress flow did not represent the real damage in the core, as pointed out by
Castanie et al. [123].

Although a continuum core meshed with solid elements may seem a convenient way to
represent the honeycomb core geometrically, errors have been attributed to the continuum
model when it is used to model damage [107, 114]. One reason is that it may be difficult to
simulate the exact damage progression using solid elements due to the discontinuous surfaces
of the honeycomb core in contact with the facesheets. The onset of damage initiation and
propagation in the honeycomb core may be highly sensitive to the local damage distribution
along the cells. Moreover, damage is assumed to progress at an even rate throughout the
continuum model, whereas damage in the test specimens may occur at a distance of
approximately cellular width apart [114]. Thus, the local stress field and damage distribution
may not be accurately represented in the core, especially in the impact damage region.

The limitation of the continuum model can be overcome by adopting a micromechanical
approach, where the actual discrete hexagonal microstructure of the honeycomb is taken into
consideration. Here, each cellular cell is modelled explicitly with shell elements, such that the
final model is an accurate and detailed representation of the real geometry. Previous
investigations have assumed this modelling approach. Nguyen et al. [71] accurately simulated
the impact response for aluminium honeycomb sandwich plates, and found that the structural
response and impact damage resistance of these materials are sensitive to the core geometry.
Atkay et al. [7] also modelled the transverse crushing behaviour of aluminium and Nomex
honeycomb cores and reported good correlation with test data. Similarly, Mohr and Doyoyo
[111, 112] analysed the deformation mechanisms of aluminium honeycombs under large
deformation, and found that the constitutive behaviour of these honeycombs during plastic
collapse is largely controlled by folding systems. All these studies also demonstrate that a
micromechanical model is useful in understanding the effect of core geometry on the
29
structural response. Figure 14(a) show an idealized homogenous core model while Figure 4(b)
shows a geometrically accurate finite element model of the honeycomb sandwich panel [95].
Figure 15 compares the experimental results of maximum force and deflection with increasing
impact energy, with the results obtained using the finite model of Figure 14(b) [95]. It can be
seen that these results are in good agreement.

8 Impact Influence Factors
This section discussed the various factors that have been known to have significant influence
on the impact response, these factors will be now described in some details :

1) Impact Mass
As discussed earlier, the impact mass M dominates the impact response event. Larger impact
mass resulted in larger deflection and longer impact duration [26, 38]. The quasi-static method
is widely used to analyze the low-velocity impact and researchers have concluded that the
mass ratio, m, is a key parameter in classifying this type of impact response [42, 52, 72].
Besides impact velocity in large mass impact, impact mass is the main parameter that
influence impact duration [25, 26].

2) Impact velocity
The magnitude of impact velocity should be considered in conjunction with the impact mass
[25, 26]. In low-velocity impact, it is widely accepted that the impact velocity does not
dominate the impact response but will somewhat influence the contact force and deflection
response [25-26, 38, 47, 49]. The maximum contact force is found to be proportional to the
impact velocity with the condition that damage is barely visible [3, 38]. Increasing impact
velocity will result in larger contact force and deflection, but has little effect on the impact
30
duration [25, 30]. Higher impact velocity will complement the energy-absorption capability of
the target [25, 38].

3) Geometry of Impactor
The geometry of the impactor or indentor is vital parameter that governs the contact response
between the impactor and the top facesheet of the sandwich structure. The effect of impactors
shape on impact response and failure mechanism of sandwich panel has been investigated via
experiments with both hemispherical and flat-ended indenters [27, 73-75]. Extensive
experiments has been conducted by Zhou [27] and it is found that spherical-end impactor will
cause lower ultimate load than the flat-ended impactor due to greater stress concentration
around the edge of impactor. Literature search and review showed that spherical impactor
attracts more research attention [18, 36-37, 42, 72, 76-78] as compared to other shapes of
impactor. It was observed from experimental results that the increase of the radius of the
spherical head of the impactor gives a larger contact force but has little influence on the plate
central displacement [28]. From the results of actual impact tests on [(0/45)
n
/core/(45/0)
n
]
sandwich plates with indentors of different radii, Raju [78-79] reported the effect of indentor
size on the resulting damage initiation, damage propagation and the damage modes. He also
concluded that indentor of smaller radius produced dominantly facesheet damage in the form
of matrix cracking and fibre fractures while indentor of larger radius produced core crushing
that dominated until the fracture initiated in the facesheet in contact.

The geometrical size and property of the impactor are usually related to wavelength of the
panel during impact [13, 80] and they are found to influence impact response [81-82].
Recently, experiments using flat-ended indentors started appearing in the literature [74, 83]. It
was found that flat-ended indentor leads to greater threshold loads and corresponding panel
deflections. The chief reason is that flat-ended indentors provided larger contact area and thus
imposing a greater degree of membrane stretching of skins.
31

4) Boundary conditions
Impact response has been shown to be very sensitive to boundary conditions for large mass
impact while it is not so for small mass impact. In theory, simple support, clamp and free are
common boundary conditions. For the same materials and dimensions, a clamped plate will
give a larger structural stiffness than a simply supported plate. Thus different boundary
condition will affect the magnitude of the bending, shear and membrane stiffness. These
effects were reviewed in Section 6.1. From the published results based on theoretical and
numerical investigation [31], it can be concluded that sandwich structures with larger
structural stiffness will results in larger contact force and smaller deflection when impacted,
and thereby is prone to produce more damage.

5) Thickness of the facesheet
The thickness of the composite facesheet plays an important role in impact responses. With
constant impact energy, the thicker plate results in larger contact force and smaller central
displacement. The impact duration tends to decrease with thicker plates [3, 55, 73].
Experimental results of sandwich panels with cross-ply skins varying from 8 to 16 plies show
that both the threshold load and the initial slope are sensitive to skin thickness[27]. With a
linear fit, it was found that the increase in the threshold load and ultimate load are about 0.19
kN and 0.91kN per ply respectively. Zhou[27] advocated that changing skin thickness not
only influence the flexural rigidity but also change the load transfer between the top skin/
core thereby influence the development of damage mechanisms. In all, the effect of the
thickness of the facesheet in a small range can actually be attributed to increasing bending
stiffness and shear stiffness of the panel and thereby affect the response. Increasing the
thickness of the facesheet may change the load transfer mechanism between the top facesheet
and the core.

32
6) Strain-rate effect
Strain rate is considered one of the fundamental material properties. It is well known that it
contributes significantly to impact response of metallic structures especially in high-velocity
impact. This effect of the strain rate is more significant if combined with small mass impact [3,
11]. Modern materials such as polymer [22] and some type of foam core [67] are very
sensitive to strain-rate. Published literatures on the strain-rate effect on impact response of
composites are few and rare [23]. Mathematical model and theoretical solution considering the
strain rate effect on the impact response of low-velocity impact is at its infancy. The
experimental study on the effect of strain rate on the impact response of sandwich structures
has not been reported yet. This is partly due to two reasons: 1) the impact velocities used in
the laboratory are quite low and the resulting strain rate is therefore not that high and, 2) the
mass of target is small compared to impact mass and thus resulting in a large mass impact. It is
a fact that for large mass impact of metallic structures, the effect of strain rate can be ignored.

7) Compressive and Tensile Preload
It is obvious that compressive preload will increase deflection in the transversely and quasi-
statically loaded composite plates. The energy absorption capability was found to be higher
compared to that of the unloaded plates. Preloading has been to shown to increase the
composites resistance to delaminations which is an important energy absorption mechanism
[69, 84]. The effect of tensile preload on the response of laminated composite plates subjected
to impact loading was investigated with different impactor shapes via experiments [85]. It was
found that increasing pre-tension will decrease the contact duration and deflection while its
effect on the peak force, absorbed energy and damage area diminishes. Other similar works on
the effect of preload on the response of laminated composites are also well-documented [84-
87]. Through experiment on biaxial pre-loaded glass/polyester woven laminates [87], the
ballistic limit was found to be improved approximately 11% higher. Experimental
investigation by Whittingham [86] indicates that the indentation depth, peak force and
33
absorbed energy are independent of the magnitude of the pre-load at low levels of impact
energy while they are becoming significant at higher levels of impact energy. To date,
research work on the effect of preload on the impact response of sandwich structures is not yet
available.

8) Environmental conditions : temperature and moisture
The effect of temperature and moisture on the impact response of sandwich structures has
been investigated experimentally [88-90]. It was found that temperature has a significant
influence on energy absorption [88-89] while having little effect on the effective flexural
stiffness. Lower temperature will reduces composites resistance to delaminations [88]. The
maximum impact force was also found to decrease with increasing temperature from 25
0
C to
75
0
C. Experimental studies on the effect of temperature and moisture to sandwich panel were
conducted extensively by Morganti and his co-workers [91]. However, investigation on the
effect of moisture absorption in sandwich panel on impact response, to the best of the authors
knowledge, is still absent in the literature.

9) Impact angle
The literature reviewed so far deals with the effect of direct impact, that is, the direction of
impact velocity is perpendicular to the surface of target. However, this is not always the case
in practice as the angle of impact angle is a variable. Oblique impact on laminates and other
composites has been investigated experimentally and theoretically [92-94]. Stronge [92] has
derived a mathematical model for the oblique impact response of inflated balls. He found that
the final angular velocity of the ball is dependent on the impact angle for the range of impact
angle between 45
0
from the normal direction. Stronge and his co-workers [95] also
investigated the ballistic limit of oblique impact on sandwich panels. They deduced that for
the both spherical-ended and flat-ended impactor to penetrate thin metallic plates (either
monolithic or layered), the impactor must have a minimum ballistic limit speed at an angle of
34
obliquity between 30
0
and 45
0
from the normal direction. Comparing the results of oblique
impact was with those of direct impact showed that laminated composite plates absorb more
energy during oblique impact [92, 94]. For sandwich panel, ballistic perforations by oblique
impact were investigated extensively using tests and numerical simulation [95]. Results on
sandwich panel subjected to low-velocity impact are however rather limited.

10) Non-Dimensional Coefficient
A non-dimensional coefficient
1
16
h
Mk
D

= that defines maximum indentation and


maximum contact force has been identified as a key parameter in the impact response for
small mass impact. It was found that the parameter is independent of the boundary
conditions and the impact velocity. The details were discussed earlier in Section 6.3.
9 Damage Mechanisms
In this section, a review on damage evolution of facesheet as well as the failure modes and
energy absorption of sandwich structures will be presented in some details.
9.1 Energy Absorption during Impact
The energy absorbed by sandwich structures during impact has been shown to consist of the
following [27, 96-99] :
Strain energy absorbed by the sandwich.
Kinetic energy of the impactor and the sandwich.
Strain energy absorbed by the impactor which is usually ignored as the impactor is
treated as a rigid.
Fracture energy in the facesheet. Experimental results revealed that the energy absorbed
by fracture of the facesheets dominated the energy absorption in both low-velocity and
high-velocity impact. For composite facesheets, this energy has been shown to be
35
contributed from either the tensile fibre stretching/damage or the matrix
cracking/fracture or a combination of both. The impact damaged structure is usually put
through an experiment known compression after impact or CAI to determine its
residual strength [3, 7, 101] and its stiffness degradation [97].
Energy to buckle and crush the core.
Energy to cause debonding between the face-sheet and the core. This situation happens
either when the bond strength is low or when the wave reflection from boundary
produces intense load levels. The quantity of the energy absorbed by debonding is
usually a small fraction of the total energy.
Energy that is lost to friction, sound, heat, etc. This part is usually ignored for it is less
than 3% of the total impact energy.

Various energy-absorption effectiveness factors [100] have been formulated and used to
investigate the energy absorption capability of structural materials and other advanced
materials. The specific energy absorption capabilities of composites and composite sandwich
panels have been proven to be superior to those of metals [98]. In general, published
experimental results revealed that the absorption capability increases with higher impact
velocity and decreases with decreasing transverse stiffness of plates [27, 96-100].
9.2 Damage Evolution during Impact
Although various experimental techniques including ultrasonic inspection, X-rays,
micrography, thermography, stereography, and de-ply technique are available to characterize
the various type of damages in composite sandwich plate after impact [1, 27, 75, 78, 102], the
work on damage evolution is still in its infancy.

Five different competing failure modes has been identified for low-velocity impact damage on
sandwich panels with honeycomb cores [2-3, 29, 53, 60, 74, 79] :
36
- core buckling and debonding,
- core shear and cracking,
- delamination in the top facesheet,
- facesheet matrix cracking and,
- fibre breakage in the facings.

Damage in the core is mainly from the outraging of transverse contact pressure while damage
in top face-sheet is mainly from localised deformation around contact area, similar to that
observed in laminates [43]. It is widely reported that bottom facesheet remained intact during
and after impact in most low-velocity impact tests [27, 53, 102]. Daniel [53] reported that the
stress-strain relationship in the bottom facesheet remains essentially linear throughout the
low-velocity impact event.

Investigation on impact damage evolution of sandwich structures is usually preceded by the
comparison the results of the impact test with those of quasi-static test [3, 27, 53, 74].
Experimental results [53, 75] showed that in general the response of the contact force with
the displacement of impactor may be divided into three stages as shown in Figure 16 :
- Stage I Both the facesheet and the core are in the elastic region without any visible
damage. The response is essentially linear elastic and ends at an initial threshold load
P
1
at displacement u
1
.
- Stage II After the impact force exceeds P
1
, the curve becomes nonlinear with a
sudden drop in structural stiffness indicating the onset of initial damage that may
include core bucking and localised damage in the facesheet. The impact force increases
significantly in this stage mainly due to the membrane effect of the top facesheet. In
quasi-static tests, this drop in the load is obvious and is observed to occur together with
37
a cracking sound. This stage ends with a maximum force P
m
at displacement u
f
. This
maximum force P
m
is usually much larger than the initial threshold force.
- Stage III If the top facesheet is perforated, the stiffness and load carrying capacity of
the impacted sandwich panel will drop dramatically. If there is no penetration, the
impactor will rebound and the panel will have barely visible damage.

Competing damage evolutions of sandwich plates during impact are very sensitive to the skin
thickness, the core density and core thickness, and the indentor nose shape. The core is found
to play an important part in initializing and controlling the sequence in which the various
failure mechanisms occur in low-velocity impact [42, 96, 99]. Mines and his co-workers [60]
claimed that the core crushing dominates overall energy absorption in sandwich panel with
fabric skin in low-velocity impact.

Raju and Smith [79] developed a damage evolution based on extensive experimental results
with emphasis on the effect of core thickness. Figure 17 illustrates the damage progression in
five stages :
(i) Damage initiation in core and facesheet;
(ii) Damage progression in facesheet and core;
(iii) Facesheet fracture for thick core while core crushing and consolidation for thin core;
(iv) Top facesheet penetration and core crushing for thick core while core consolidation
for thin core;
(v) Damage initiation in the bottom facesheet.

For sandwich structures with larger core density, another damage evolution that arises from
experimental study is the core shear failure mode [67, 103] as shown in Figure 18. For this
mode to occur, Fatt [103] estimated the shear strain in honeycomb core by defining the shear
angle as :
38
core
c
a R
o
=

(65)
Since the laminated facesheets offer the first line of resistance to impact and allows the
transmission of the load to the core, the strength of the laminated facesheet plays an important
role in the energy absorption capability of sandwich panels. Abrate [2-3] remarked that
delaminations are of the first concern since they significantly reduce the strength and stiffness
of the laminate and they also play a vital role at interfaces with different plies orientations. To
investigate the failure mode of facesheet, various failure criteria such as Puck, Tsai-Hill, Tsai-
Wu, Hashin, Chang-Chang, etc were widely used to model the failure response of composite.
Each of these theories has its advantages and disadvantages. However, failure analyses of
composites still have plenty of rooms for improvement and new developments. In 1992, Prof.
M.J. Hinton, Mr. P.D. Soden and Dr. A.S. Kaddour launched and organized a World Wide
Failure Exercise (WWFE), which is referred to as the Failure Olympics of leading
theoreticians in the research field of composite to evaluate the various failure criteria. The
organizer of the exercise stated that much work still remains to be done in order to ensure that
reliable and accurate predictive tools are available for general use.
10 Conclusion
In this contribution, models using numerical method, mathematical method and experimental
investigation of sandwich panels subjected to impact load were reviewed. The impact
responses are classified according the various key parameters, and different methodologies are
listed for different classes of impact. The characteristics of each class of impact responses
were also classified and discussed.

The impact duration is regarded as criterion to distinguish categories of response of sandwich
panels subjected to impact load. Through extensive analysis and investigation, the mass
ratio, m, is selected as the key selection parameter rather than impact duration as it is easier to
39
quantify. The proportion of effective mass to the mass of the panel is investigated to be one
fourth for Kirchhoff plate while the proportion drops drastically when the bending stiffness is
much larger than the shear stiffness. In sandwich structures, the shear stiffness plays a
significant role in the transference of the stresses from top facesheet to the core and then down
to the bottom facesheet.

For low-velocity impact, the impact response is boundary-controlled for large mass impact
and can be modelled using quasi-static method while it is wave-controlled for small mass
impact. Different kinds of impact response could be solved using different methods, which
were discussed. Popular mathematical solutions for impact response were also presented in
some details. The investigation of the various modes of failure in composite sandwich panels,
the energy-absorption effectiveness and the damage mechanism for sandwich panels are
summed up towards the end of the review.
References

[1] K. Diamanti, C. Soutis, and J. M. Hodgkinson, Non-destructive inspection of
sandwich and repaired composite laminated structures, Composites Science and
Technology, vol. 65, pp. 2059-2067, 2005.
[2] S. Abrate, Localized impact on sandwich structures with laminated facings,
Internation Journal of Impact Engineering, vol. 50, no. 2, pp. 69-82, 1997.
[3] S. Abrate, Impact on composite structures: Cambridge University Press, 1998.
[4] H. G. Allen, Analysis and design of structural sandwich panels, 1969.
[5] T. A. Anderson, An investigation of SDOF models for large mass impact on
sandwich composites, Composites: Part B, vol. 36, pp. 135-142, 2005.
[6] P. Feraboli, Damage resistance characteristics of thick-Core honeycomb composite
panels, in 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and
Materials Conference, Newport, 2006.
[7] P. Feraboli, and K. T. Kedward, A new composite structure impact performance
assessment program, Composites Science and Technology, vol. 66, pp. 1336-1347,
2006.
[8] Marguerre, K., The optimum buckling load of flexibly supported plate composed of two sheets
joined by a light weight filler, when under longitudinal compression. Deutsche
Vierteljahrsschrist fur Literalurwissenschaft und Giests Geschiechte, 1944. D.V.L: p. 5.
[9] C. C. Foo, G. B. Chai, and L. K. Seah, A model to predict low-velocity impact
response and damage in sandwich composites, Composites Science and Technology,
vol. 68, pp. 1348-1356, 2008.
[10] R. Olsson, Engineering method for prediction of impact response and damage in
sandwich panels, Journal of Sandwich Structures and Materials, vol. 4, pp. 3-19,
2002.
40
[11] D. W. Zhou, and W. J. Stronge, Low velocity impact denting of HSSA lightweight
sandwich panel, International Journal of Mechanical Sciences, vol. 48, pp. 1031-
1045, 2006.
[12] G. Luo, X. Lv, and L. Ma, Dynamics of an impact-progressive system, Nonlinear
Analysis: Real World Applications, vol. 10, pp. 665-679, 2009.
[13] S. Michelle, H. Fatt, and D. Sirivolu, A wave propagation model for the high velocity
impact response of a composite sandwich panel, International Journal of Impact
Engineering, vol. 37, pp. 117-130, 2010.
[14] A. Mishra, and N. K. Naik, Failure initiation in composite structures under low-
velocity impact: Analytical studies, Composite Structures, vol. 92, pp. 436444,
2010.
[15] M. Aktas, C. Atas, B. l. Murat et al., An experimental investigation of the impact
response of composite laminates, Composite Structures, vol. 87, pp. 307313, 2009.
[16] M. Nguyen, and D. Elder, A review of explicit finite element software for composite
impact analysis, Journal of Composite Materials, vol. 39, pp. 375-386, 2005.
[17] D. J. Elder, R. S. Thomson, M. Q. Nguyen et al., Review of delamination predictive
methods for low speed impact of composite laminates, Composite Structures, vol. 66,
no. 4, pp. 677-683 2004.
[18] D. W. Zhou, and W. J. Stronge, Dynamic indentation of lightweight sandwich
panels, in 11th European Conference on Composite Materials, Rhodes, Greece, 2004.
[19] W. J. Cantwell, and J. Morton, Comparison of the low and high velocity impact
response of CFRP, Composites, vol. 20, pp. 545-551, 1989.
[20] M. Shaker, F. Ko, and J. Song, Comparison of the low and high velocity impact
response of Kevlar fiber-reinforced epoxy composites, Journal of Composites
Technology and Research, vol. 21, no. 4, pp. 1-6, 1999.
[21] K. Malekzadeh, M. R. Khalili, R. Olsson et al., Higher-order dynamic response of
composite sandwich panels with flexible core under simultaneous low-velocity impacts
of multiple small masses, International Journal of Solids and Structures7, vol. 43, pp.
66676687, 2006.
[22] N. K. Naik, and Y. Perla, Mechanical behaviour of acrylic under high strain rate
tensile loading, Polymer Testing, vol. 27, no. 4, pp. 504-512, 2008.
[23] B. C. Santiuste, S. Snchez-Sez, E. Barbero et al., Modelling of composite sandwich
structures with honeycomb core subjected to high velocity impact, Composite
Structures, pp. Article in press, 2009.
[24] M. S. H. Fatt, and D. Sirivolu, A wave propagation model for the high velocity
impact response of a composite sandwich panel, International Journal of Impact
Engineering, vol. 37 no. 2, pp. 117-130, 2010.
[25] P. W. Bland, and S. Michelle, A discussion on the role played by velocity in impact
mechanics, in The 19th Conference of Mechanical Engineering Network of Thailand,
Phuket, Thailand, 2005.
[26] R. Olsson, Mass criterion for wave controlled impact response of composite plates,
Composites: Part A 31 (2000) 879887, vol. 31, pp. 879-887, 2000.
[27] G. Zhou, and M. Hill, Investigation of parameters governing the damage and energy
absorption characteristics of honeycomb sandwich panels, Journal of Sandwich
Structures and Materials, vol. 9, no. 4, pp. 309-342, 2007.
[28] D. Zheng, and W. K. Binienda, Effect of permanent indentation on the delamination
threshold for small mass impact on plates, International Journal of Solids and
Structures, vol. 44, pp. 8143-8158, 2007.
[29] D. W. Zhou, and W. J. Stronge, Impact damage on lightweight sandwich panels,
International Journal of Impact Engineering, vol. 35, pp. 1339-1354, 2008.
[30] I. M. Daniel, "Impact response and damage tolerance of composite sandwich
structures," Dynamic Failure of Materials and Structures: Springer, 2009.
41
[31] P. Qiao, Impact and damage prediction of sandwich beams with flexible core
considering arbitrary boundary effects, Journal of Sandwich Structures and
Materials, vol. 9, pp. 411-444 2007.
[32] B. C. Santiuste, S. Snchez-Sez, E. Barbero et al., Modelling of composite sandwich
structures with honeycomb core subjected to high velocity impact, Composite
Structures, vol. 92, no. 9, pp. 2090-2096, 2010.
[33] D. W. Zhou, and W. J. Stronge, Mechanical properties of fibrous core sandwich
panels, International Journal of Mechanical Sciences, vol. 47, pp. 775-798, 2005.
[34] K. L. Johnson, Contact mechanics, Cambridge: Cambridge University Press, 1985.
[35] C. T. Sun, Strength analysis of unidirectional composites and laminates, Internatianl
Comprehensive Composite Materials, vol. 20, pp. 641-666, 2000.
[36] R. Olsson, and H. L. McManus, Improved theory for contact indentation of sandwich
panels, vol. 34, no. 6, pp. 1238-1244, 1996.
[37] S. H. Yang, and C. T. Sun, Indentation Law for Composite Laminates, in Composite
Materials: Testing and Design(Six Conference), 1982, pp. 425-449.
[38] R. Olsson, Analytical prediction of large mass impact damage in composite
laminates, Source: Composites Part A: Applied Science and Manufacturing, vol. 32,
pp. 1207-1215, 2001.
[39] G. G. Corbett, S. R. Reid, and W. Johnson, Impact loading of plates and shells by
free-flying projectiles: a review, International Journal of Impact Engineering, vol.
18, no. 2, pp. 141-230, 1996.
[40] J. P. Dear, H. Lee, and S. A. Brown, Impact damage processes in composite sheet and
sandwich honeycomb materials, International Journal of Impact Engineering, vol. 32,
pp. 130-145, 2005.
[41] P. Hampson, and M. Moatamedi, A review of composite structures subjected to
dynamic loading, International Journal of Crashworthiness, vol. 12, no. 4, pp. 411-
428, 2007.
[42] J. K. Kim, and T. X. Yu, Indentation, penetration and perforation of composite
laminate and sandwich panels under quasi-static and projectile loading, Key
Engineering Materials, vol. 141, pp. 501-552, 1998.
[43] M. O. W. Richardson, and M. J. Wisheart, Review of low-velocity impact properties
of composite materials Composites. Part A, vol. 27, pp. 1123-1131, 1996.
[44] J. F. Larry, On the duration of contact for the Hertzian impact of a spherical indenter
on a maxwell solid, Journal of Solids Structures, vol. 10, pp. 621-624, 1974.
[45] A. A. Aleksandrov, V. I. Ivanov, and G. S. Rosin, Determining the duration of contact
in an impact, Izmeritelnaya Tekhnika, vol. 5, pp. 96-97, 1968.
[46] K. N. Shivakumar, W. Elber, and W. Illg, Prediction of Impact Force and Duration
Due to Low-Velocity Impact on Circular Composite Laminates, Journal of Applied
Mechanics, vol. 52, pp. 674-681, 1985.
[47] R. Olsson, Closed form prediction of peak load and delamination onset under small
mass impact, Composite Structures, vol. 59, pp. 341-349, 2003.
[48] W. J. Stronge, Impact Mechanics: Cambridge University Press, 2000.
[49] P. Robinson, and G. Davies, Impactor mass and specimen geometry effects in low
velocity impact of laminated composites., International Journal of Impact
Engineering, vol. 12, no. 2, pp. 189-207, 1992.
[50] L. Asp, and R. Juntikka, High velocity impact on NCF reinforced composites,
Composites Science and Technology, vol. 69, pp. 1478-1482, 2009.
[51] S. Pashah, M. Massenzio, and E. Jacquelin, Prediction of structural response for low
velocity impact, International Journal of Impact Engineering, vol. 35, pp. 119-132,
2008.
[52] R. Swanson, Limits of quasi-static solutions in impact of composite structures,
Composites Engineering, vol. 2, no. 4, pp. 261-267, 1992.
42
[53] I. M. Daniel, "Impact Response and Damage Tolerance of Composite Sandwich
Structures " Dynamic Failure of Materials and Structures, pp. 191-233, Landon:
Springer, 2010.
[54] S. P. Timoshenko, Theory of Plates and Shells, Woinowsky-Krieger: McGraw-Hill,
1959.
[55] A. C. Ugural, Stresses in Plates and Shells, 2nd edition ed., McGraw-Hill: McGraw-
Hill, 1998.
[56] W. J. Cantwel, and J. Morton, The impact resistance of composite materials---a
review, Composites, vol. 22, no. 5, pp. 347-362, 1991.
[57] C. C. Foo, L. K. Seah, and G. B. Chai, Low-velocity impact failure of aluminium
honeycomb sandwich panels, Composites Science and Technology, vol. 85, pp. 20-28,
2008.
[58] V. J. Hawyes, P. T. Curtis, and C. Soutis, Effect of impact damage on the
compressive response of composite laminates, Composites Part A: Applied Science
and Manufacturing, vol. 32, no. 9, pp. 1263-1270, 2001.
[59] R. K. Luo, E. R. Green, and C. J. Morrison, Impact damage analysis of composite
plates, International Journal of Impact Engineering, vol. 22, pp. 435-447, 1999.
[60] R. A. W. Mines, C. M. Worrall and A. G. Gibson, Low velocity perforation behaviour
of polymer composite sandwich panels, Internation Journal of Impact Engineering,
vol. 21, no. 10, pp. 855-879, 1998.
[61] A. S. Yigit, and A. P. Christoforou, On the impact between a rigid sphere and a thin
composite laminate supported by a rigid substrate, Composite Srrucrures, vol. 30, pp.
169-177, 1995.
[62] R. Ferri, and B. V. Sankar, A comparative study on the impact resistance of
composite laminates and sandwich panels, Journal of Thermoplastic Composite
Materials, vol. 10, pp. 304-315, 1997.
[63] F.-K. Chang, and K.-Y. Chang, A progressive damage model for laminated
composites containing stress concentrations, Journal of Composite Materials, vol. 21,
no. 9, pp. 834-855, 1987.
[64] P. A. Fomitchov, A. K. Kromin, S. Krishnaswamy et al., Imaging of damage in
sandwich composite structures using a scanning laser source technique, Composites:
Part B, vol. 35, pp. 557-562, 2004.
[65] F. J. Yang, and W. J. Cantwell, Impact damage initiation in composite materials,
Composites Science and Technology, pp. Article in press, 2009.
[66] D. Zenkert, An introduction to sandwich construction, London: Emas Publishing,
1995.
[67] F. Zhu, G. Lu, D. Ruan et al., Plastic Deformation, Failure and Energy Absorption of
Sandwich Structures with Metallic Cellular Cores, International Journal of Protective
Structures, vol. 1, no. 4, pp. 507-541, 2010.
[68] M. q. Nguyen, D. j. Elder, and J. Bayandor, A Review of Explicit Finite Element
Software for Composite Impact Analysis, Journal of Composite Materials, vol. 39,
no. 4, pp. 375-386, 2005.
[69] K. M. Mikkor, R. S. Thomson, and I. Herszberg, Finite element modelling of impact
on preloaded composite panels, Composite Structures, vol. 75, no. 1-4, pp. 501-513,
2006.
[70] V. Tita, J. d. Carvalho, and D. Vandepitte, Failure analysis of low velocity impact on
thin composite laminates: Experimental and numerical approaches, Composite
Structures, vol. 83, no. 4, pp. 413-428, 2008.
[71] M. Q. Nguyen, S. S. Jacombs, R. S. Thomson et al., Simulation of impact on
sandwich structures, Composite Structures, vol. 67, no. 2, pp. 217-227, 2005.
[72] T. Anderson, and E. Madenci, Graphite/epoxy foam sandwich panels under quasi-
static indentation, Engineering Fracture Mechanics, vol. 67, pp. 329-344, 2000.
43
[73] J. Tomblin, B. S. T. Lacy, S. Hooper et al., Review of Damage Tolerance for
Composite Sandwich Airframe Structures, DOT/FAA/AR-99/49, University of
Maryland, Virginia, 1999.
[74] G. Zhou, Static behaviour and damage of thick composite laminates, Composite
Structures, vol. 36, no. 1-2, pp. 13-22, 1996.
[75] G. Zhou, M. Hill, J. Loughlan et al., Damage Characteristics of Composite
Honeycomb Sandwich Panels in Bending under Quasi-static Loading, Journal of
Sandwich Structures and Materials, vol. 8, pp. 55-89, 2006.
[76] J. L. Abot, I. M. Daniel, and E. E. Gdoutos, Contact law for composite sandwich
beams, Journal of Sandwich Structures and Materials, vol. 4, no. 2, pp. 157-173,
2002.
[77] S. M. Lee, and T. K. Tsotsis, Indentation failure behavior of honeycomb sandwich
panels, Composites Science and Technology, vol. 60, pp. 1147-1159, 2000.
[78] K. S. Raju, and J. S. Tomblin, Damage Characteristics in Sandwich Panels subjected
to Static Indentation using Spherical Indentors, in 42nd
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials
Conference, Seattle, WA., 2001.
[79] K. S. Raju, and B. L. Smith, Impact Damage Resistance and Tolerance of
Honeycomb Core Sandwich Panels, Journal of Composite Materials, vol. 42, no. 4,
pp. 385-413, 2008.
[80] L.-l. Wang, Foundations of stress waves: Elsevier Science, 2007.
[81] J. P. Hou, and C. Ruiz, Soft body impact on laminated composite materials,
Composites: Part A, vol. 38, pp. 505-515, 2007.
[82] A. F. Johnson, and M. Holzapfel, Modelling soft body impact on composite
structures, Composite Structures, vol. 61, pp. 103-113, 2003.
[83] I. Daniel, "The influence of core properties on failure of composite sandwich beams."
pp. 63-74.
[84] S. Heimbs, S. Heller, and P. Middendorf, Low velocity impact on CFRP plates with
compressive preload: Test and modelling, International Journal of Impact
Engineering, vol. 36, no. 10-11, pp. 1182-1193, 2009.
[85] T. Mitrevski, I. H. Marshalla, and R. S. Thomson, Low-velocity impacts on preloaded
GFRP specimens with various impactor shapes Composite Structures, vol. 76, no. 3,
pp. 209-217, 2006.
[86] B. Whittingham, I. H. Marshall, and T. Mitrevski, The response ofcomposite
structures with pre-stress subject to low velocity impact damage, Composite
Structures, vol. 66, pp. 685988, 2004.
[87] S. K. Garca-Castilloa, S. Snchez-Seza, J. Lpez-Puentea et al., Impact behaviour
of preloaded glass/polyester woven plates, Composites Science and Technology, vol.
69, no. 6, pp. 711-717, 2009.
[88] S. Sanchez-Saez, E. Barbero, and C. Navarro, Analysis of the dynamic flexural
behaviour of composite beams at low temperature, Composites Science and
Technology, vol. 67, pp. 26162632, 2007.
[89] M. D. Erickson, A. R. Kallmeyer, and K. G. Kellogg, Effect of Temperature on the
Low-velocity Impact Behavior of Composite Sandwich Panels, Journal of Sandwich
Structures and Materials, vol. 7, pp. 245-264, 2005.
[90] A. Salehi-Khojina, and M. Mahinfalah, Temperature effects on Kevlar/hybrid and
carbon fiber composite sandwiches under impact loading, Composite Structures, vol.
78, no. 2, pp. 197-206, 2005.
[91] F. Morganti, M. Marchetti, and G. Reibaldi, Effects of moisture and thermal ageing
on structural stability of sandwich panels, Acta Astronautica, vol. 11, pp. 489-508,
1984.
44
[92] N. K. Gupta, and V. Madhu, An experimental study of normal and oblique impact of
hard-core projectile on single and layered plates, International Journal of Impact
Engineering, vol. 19, no. 5-6, pp. 395-414, 1997.
[93] W. J. Stronge, and A. D. C. Ashcroft, Oblique impact of inflated balls at large
deflections, International Journal of Impact Engineering, vol. 34, pp. 1003-1019,
2007.
[94] P. J. Hazel, G. Kister, C. Stennett et al., Normal and oblique penetration of woven
CFRP laminates by a high velocity steel sphere, Composites Part A: Applied Science
and Manufacturing, vol. 39, no. 5, pp. 866-874, 2008.
[95] D. W. Zhou, and W. J. Stronge, Ballistic limit for oblique impact of thin sandwich
panels and spaced plates, International Journal of Impact Engineering, vol. 35, pp.
1339-1354, 2008.
[96] J. Dean, C. S. Dunleavy, P. M. Brown et al., Energy absorption during projectile
perforation of thin steel plates and the kinetic energy of ejected fragments,
International Journal of Impact Engineering, vol. 36, no. 10, pp. 1250-1258, 2009.
[97] C. C. Foo, Energy absorption characteristics of sandwich structures subjected to low-
velocity impact, Nanyang Technological University, Singapore, 2008.
[98] V. Skvortsova, J. Keplerb, and E. Bozhevolnaya, Energy partition for ballistic
penetration of sandwich panels, International Journal of Impact Engineering, vol. 28,
pp. 697-716, 2003.
[99] L. Sun, R. F. Gibson, F. Gordaninejad et al., Energy absorption capability of
nanocomposites: A review, Composites Science and Technology, vol. 69, pp. 2392
2409, 2009.
[100] N. Jones, Energy-absorbing effectiveness factor, Impact Loading of Lightweight
Structures vol. 37, no. 6, pp. 754-765, 2009.
[101] V. Kostopoulos, A. B. Karapappas, A. Vavouliotis et al., Impact and after-impact
properties of carbon fibre reinforced composites enhanced with multi-wall carbon
nanotubes, Composites Science and Technology, vol. 70, no. 4, pp. 553-563, 2010.
[102] P. H. W. Tsang, Impact resistance and damage tolerance of composite sandwich
panels, Thesis (Ph. D.), Dept. of Aeronautics and Astronautics, Massachusetts
Institute of Technology, Massachusetts 1994.
[103] M. S. H. Fatt, and K. S. Park, Dynamic models for low-velocity impact damage of
composite sandwich panels Part B: Damage initiation, Composite Structures, vol.
52, no. 3-4, pp. 353-364, 2001.
[104] M. Meo, A.J. Morris, R. Vignjevic, and G. Marengo. Numerical simulation of
lowvelocity impact on an aircraft sandwich panel. Composite Structures, 62:353360,
2003.
[105] T. Besant, G. A. O. Davies, and D. Hitchings. Finite element modelling of low velocity
impact of composite sandwich panels. Composites Part A: Applied Science and
Manufacturing, 32:11891196, 2001.
[106] L. Aktay, A. F. Johnson, and M. Holzapfel. Prediction of impact damage on sandwich
composite panels. Computational Materials Science, 32(3-4):252260, 2005.
[107] L. Aktay, A. F. Johnson, and Bernd-H. Kroplin. Numerical modelling of honeycomb
crush behaviour. Engineering Fracture Mechanics, 75(9):26162630, 2008.
[108] G. A. O. Davies and X. Zhang. Impact damage prediction in carbon composite
structures. International Journal of Impact Engineering, 16(1):149170, 1995.
[109] J. Zhang and M. F. Ashby. The out-of-plane properties of honeycombs. International
Journal of Mechanical Sciences, 34(6):475489, 1992.
[110] L. J. Gibson and M. F. Ashby. Cellular solids: Structures & properties. Pergamon
Press, 1988.
[111] D. Mohr and M. Doyoyo. Large plastic deformation of metallic honeycomb:
orthotropic rate-independent constitutive model. International Journal of Solids and
Structures, 41:44354456, 2004.
45
[112] D. Mohr and M. Doyoyo. Deformation-induced folding systems in thin-walled
monolithic hexagonal metallic honeycombs. International Journal of Solids and
Structures, 41:33533377, 2004.
[113] L. Karger, J. Baaran, and J. Tessmer. Rapid simulation of impacts on composite
sandwich panels inducing barely visible damage. Composite Structures, 79(4):527
534, 2007.
[114] D. P. W. Horrigan, R. R. Aitken, and G. Moltschaniwskyj. Modelling of crushing due
to impact in honeycomb sandwiches. Journal of Sandwich Structures and Materials,
2:131151, 2000.
[115] F. K. Chang, H. Y. Choi, and S. T. Jeng. Study on impact damage in laminated
composites. Mechanics of Materials, 10:8395, 1990.
[116] D. Jiang and D. Shu. Local displacement of core in two-layer sandwich structures
subjected to low-velocity impact. Composite Structures, 71:5360, 2005.
[117] F. Meraghni, F. Desrumaux, and M.L. Benzeggagh. Mechanical behavior of cellular
core for structural sandwich panels. Composites Part A: Applied Science and
Manufacturing, 30:767779, 1999.
[118] J. Hohe and W. Becker. A mechanical model for two-dimensional cellular sandwich
cores with general geometry. Computational Materials Science, 19:108115, 2000.
[119] W. S. Burton and A. K. Noor. Assessment of continuum models for sandwich panel
honeycomb cores. Computer Methods in Applied Mechanics and Engineering,
145:341360, 1997.
[120] L. Gornet, S. Marguet, and G. Marckmann. Numerical modelling of Nomex
honeycomb cores: Failure and effective elastic properties. In III European Conference
on Computational Mechanics, 2006.
[121] M. Q. Nguyen, S. S. Jacombs, R. S. Thomson, D.Hachenberg, and M. L. Scott.
Simulation of impact on sandwich structures. Composite Structures, 39:375386,
2005.
[122] B. Castanie, C. Bouvet, Y. Aminanda, J. J. Barrau, and P. Thevenet. Modelling of low-
energy/low-velocity impact on nomex honeycomb sandwich structures with metallic
skins. International Journal of Impact Engineering, 35:620634, 2007.
[123] A. N. Palazotto, E. J. Herup, and L. N. B. Gummadi. Finite element analysis of low-
velocity impact on composite sandwich plates. Composite Structures, 49:209227,
2000.


46
M
v
o
R
i
R
p
R
c
a
h
c
h
f
h
f
M
v
o
R
i
R
p
R
c
a
h
c
h
f
h
f


Figure 1 Basic model for sandwich panel subjected to low-velocity impact




c
o
C
c
max
o
o
q
Linear
Crushing
Consolidation
(Densification)
c
o
C
c
max
o
o
q
Linear
Crushing
Consolidation
(Densification)

Figure 2 Typical stress-strain relationship for bare honeycomb




R
c
P
r
R
c
P
r

Figure 3 Pressure distribution of the contact area between two bodies

47
-Finite Element
-Experiment
-Modal
superposition
Medium-mass impact
Mass criteria
-Finite Element
-Experiment
-Modal superposition
-Wave-controlled solution
Small mass impact
(Wave-controlled impact)
Impact Response
-Experiment
-Finite Element Analysis
-Stress-wave theory
Ballistic impact
Non ballistic impact
-Finite Element
-Experiment
-Spring-mass model
-Energy-balance model
-Boundary-controlled solution
Large mass impact
(Boundary-controlled impact)
Impact velocity, v
Impact duration, t
Mass criteria
h
t
E
>>
h
t
E
~
2
p
M
m
>
1
2
5
p
M
m
< <
1
5
p
M
m
<
-Finite Element
-Experiment
-Modal
superposition
Medium-mass impact
Mass criteria
-Finite Element
-Experiment
-Modal superposition
-Wave-controlled solution
Small mass impact
(Wave-controlled impact)
Small mass impact
(Wave-controlled impact)
Impact Response
-Experiment
-Finite Element Analysis
-Stress-wave theory
Ballistic impact
Non ballistic impact
-Finite Element
-Experiment
-Spring-mass model
-Energy-balance model
-Boundary-controlled solution
Large mass impact
(Boundary-controlled impact)
Non ballistic impact
-Finite Element
-Experiment
-Spring-mass model
-Energy-balance model
-Boundary-controlled solution
Large mass impact
(Boundary-controlled impact)
-Finite Element
-Experiment
-Spring-mass model
-Energy-balance model
-Boundary-controlled solution
Large mass impact
(Boundary-controlled impact)
-Finite Element
-Experiment
-Spring-mass model
-Energy-balance model
-Boundary-controlled solution
Large mass impact
(Boundary-controlled impact)
Impact velocity, v
Impact duration, t
Mass criteria
h
t
E
>>
h
t
E
~
2
p
M
m
>
1
2
5
p
M
m
< <
1
5
p
M
m
<


Figure 4 Solution methods for different categories of impact

48

Figure 5 Response differences between large and small mass impact [10]



0
10
20
30
40
50
5 10 15 20 25 30
Span/thickness
K
s
/
K
b
c f f c c
G E h h R 100 , 15 = = =
c f f c c
G E h h R 80 , 10 = = =
c f f c c
G E h h R 20 , 10 = = =
0
10
20
30
40
50
5 10 15 20 25 30
Span/thickness
K
s
/
K
b
c f f c c
G E h h R 100 , 15 = = =
c f f c c
G E h h R 80 , 10 = = =
c f f c c
G E h h R 20 , 10 = = =

Figure 6 Variation of the shear-bending stiffness ratio with span-thickness ratio

49
0
0.05
0.1
0.15
0.2
0.25
0 2 4 6 8 10
s1=0.001
s1=0.005
s1=0.01
*
m
b s
K K
0
0.05
0.1
0.15
0.2
0.25
0 2 4 6 8 10
s1=0.001
s1=0.005
s1=0.01
*
m
*
m
b s
K K

Figure 7 The effective mass ratio vs shear-bending stiffness ratio for circular panels



Figure 8 Illustration of small mass impact in composites


=0.9
=0.5
=0.2
=0
=0.9
=0.5
=0.2
=0

Figure 9 Closed form solution for dimensionless contact force history
50

0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 0.5 1 1.5 2 2.5 3
Non-Dimensional Time
N
o
n
-
D
i
m
e
n
s
i
o
n
a
l

D
i
s
p
l
a
c
e
m
e
n
t

.

=0.3
=0.6
0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 0.5 1 1.5 2 2.5 3
Non-Dimensional Time
N
o
n
-
D
i
m
e
n
s
i
o
n
a
l

D
i
s
p
l
a
c
e
m
e
n
t

.

=0.3
=0.6

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 0.5 1 1.5 2 2.5 3
Non-Dimensional Time
N
o
n
-
D
i
m
e
n
s
i
o
n
a
l

F
o
r
c
e
=0.3
=0.6
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0 0.5 1 1.5 2 2.5 3
Non-Dimensional Time
N
o
n
-
D
i
m
e
n
s
i
o
n
a
l

F
o
r
c
e
=0.3
=0.6

(a) Displacement vs Time (b) Force vs Time
Figure 10 Plots of non-dimensional deflection and contact force with non-dimensional time.




M
m
ef
k
c
k
m
k
b
k
s
x
1
x
2
M
m
ef
k
c
k
m
k
b
k
s
x
1
x
2


Figure 4 Spring-mass model for impact analysis



51
0
2
4
6
8
10
12
14
0 0.5 1 1.5 2
Non-Dimensional Deflection
N
o
n
-
d
i
m
e
n
s
i
o
n
a
l

I
n
d
e
n
t
a
t
i
o
n

k1=1, k2=1
k1=0.1, k2=1

Figure 5 Nonlinearity of maximum deflection with maximum indentation for different
combination of stiffnesses




(a) Finite Element

(b) Experimental
Figure 13 Comparing the predicted with the actual deformed core [97]



(a) Equivalent continuum core model (b) Real cellular core model [57]

Figure 14 Comparison of the finite element models

52
0
1
2
3
4
5
6
7
8
0 5 10 15
Impact Energy (J)
(
m
m
)

o
r

(
k
N
)
FEM (Max Deflection)
Expt(Max Deflection)
FEM (Max Force)
Expt (Max Force)

Figure 15 Impact Results of Aluminium Sandwich Panels (97)



III. Fracture
stage
Load drop
P
1
II. Core buckling
and facesheet
damage
I.Elastic stage
u u
f
u
1
P
P
m
III. Fracture
stage
Load drop
P
1
II. Core buckling
and facesheet
damage
I.Elastic stage
u u
f
u
1
P
P
m

Figure 16 Typical stages of load/displacement curve for composite sandwich panel subjected
to impact load.

53

Figure 17 Damage evolution in sandwich plates with (a) thick core and (b) thin core [79].


Figure 18 Core shear failure for (a) sandwich beam [67] and (b) sandwich panel after impact
[103].
54

Boundary
condition
Edge
condition
Bending stiffness
b
K Membrane stiffness
s
K
Clamped
Immovable
3
2 2
4
3(1 )
r
r p
E h
R
t


2
(353 191 )
648(1 )
r r
r p
E h
R
t


Movable
3
2 2
4
3(1 )
r
r p
E h
R
t


2
191
648
r
p
E h
R
t

Simply
supported
Immovable
3
2
4
3(3 )(1 )
r
r r p
E h
R
t
+

1 2
( )
r
r
p
E h
G
R
t

Movable
3
2
4
3(3 )(1 )
r
r r p
E h
R
t
+

2 2
( )
r
r
p
E h
G
R
t

Where h is the thickness of the plate and Elastic Modulus
r
E ,
Poissons ratio
r
,
1
( )
r
G and
2
( )
r
G are functions of
r
[46] :
4 3 2
1 4
4
3 2
1 191 41 32 40
( ) { (1 ) (1 ) (1 ) (1 )
(3 ) 648 27 9 9
(1 ) 8 1
[ 2(1 ) 8(1 ) 16(1 ) 16]}
3 1 4
r r r r r
r
r
r r r
r
G

= + + + + + + +
+
+
+ + + + + + + + +

4 3 2
2 4
191
(1 ) 41(1 ) 96(1 ) 120(1 ) 72
24
( )
9(3 )
r r r r
r
r
G

+ + + + + + + +
=
+

Table 1 Stiffness of a circular isotropic or quasi-isotropic plate subjected to central load


Boundary
condition
Bending stiffness,
b
K Shear stiffness,
s
K
Clamped
2
16
(3 )(1 )
b
p
D
K
R
t

=
+

2
4 ( )
(1 2ln / )
c c f
s
c p c
G h h
K
h R R
t +
=
+

Table 2 Stiffness of a sandwich plate subjected to impact load

Impact type
Impact
duration
Boundary
dependency
Damage Detectible
Ballistic impact Transient Independent Perforation Yes
Small mass impact Short Independent
Large
deformation

No
Large mass impact Long Dependent Less damage No
Table 3 Comparison of the various impact problem types

Você também pode gostar