Sei sulla pagina 1di 12

This article was downloaded by: [Indian Institute of Technology Roorkee]

On: 30 June 2015, At: 04:23


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

International Journal for Computational Methods in


Engineering Science and Mechanics
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/ucme20

Prediction of Behavior of Ceramic/Metal Composite


Panels Under Two Consecutive Ballistic Impacts
a

A. Prakash , J. Rajasankar , N. R. Iyer , N. Anandavalli , S. K. Biswas & A. K.


Mukhopadhyay
a

CSIR-Structural Engineering Research Centre, Chennai, India

CSIR-Central Glass & Ceramic Research Institute, Kolkata, India


Accepted author version posted online: 06 Feb 2014.Published online: 10 Apr 2014.

Click for updates


To cite this article: A. Prakash, J. Rajasankar, N. R. Iyer, N. Anandavalli, S. K. Biswas & A. K. Mukhopadhyay (2014)
Prediction of Behavior of Ceramic/Metal Composite Panels Under Two Consecutive Ballistic Impacts, International Journal for
Computational Methods in Engineering Science and Mechanics, 15:3, 192-202, DOI: 10.1080/15502287.2014.882431
To link to this article: http://dx.doi.org/10.1080/15502287.2014.882431

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the Content) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions

International Journal for Computational Methods in Engineering Science and Mechanics, 15:192202, 2014
c Taylor & Francis Group, LLC
Copyright 
ISSN: 1550-2287 print / 1550-2295 online
DOI: 10.1080/15502287.2014.882431

Prediction of Behavior of Ceramic/Metal Composite Panels


Under Two Consecutive Ballistic Impacts
A. Prakash,1 J. Rajasankar,1 N. R. Iyer,1 N. Anandavalli,1 S. K. Biswas,2
and A. K. Mukhopadhyay2
1

CSIR-Structural Engineering Research Centre, Chennai, India


CSIR-Central Glass & Ceramic Research Institute, Kolkata, India

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

This article presents a numerical investigation to predict the


behavior of ceramic (Al2 O3 99.5)/metal (Al5083 H116) composite
panels under two consecutive high-velocity impacts of 7.62 mm
sharp-nosed small projectiles. A numerical model is developed using the advanced nonlinear software AUTODYN. The aim of the
study is to predict the impact behavior of ceramic/metal composite panels. The study mainly focuses on the effect of arrangement
of front ceramic tiles having collinear and non-collinear joints on
the impact damage pattern. The novelty of the study presented in
this article is the prediction of high-velocity-impact response under two consecutive and closely spaced hits on composite panels
carried out in a more realistic manner. Numerical responses, such
as depth of penetration, and deformation in back plate and crack
patterns, are found to match well with the experimental results. It
is believed that the outcome of this study is helpful in the design of
a ceramic tile joint arrangement to minimize damage in the target
panel.
Keywords

Ceramic/metal composite, Ballistic impact, Ceramic tile


joints, 7.62 mm projectile

1. INTRODUCTION
In recent years, advanced ceramics have been extensively
used in various innovative applications (e.g., artificial bones,
complete engines, space shuttles, etc.). Advanced ceramics
exhibit excellent characteristic properties such as high melting point, oxidation resistance, high hardness, non-magnetism,

The authors acknowledge the participation of CSIR-Central Glass


and Ceramic Research Institute (CGCRI), Kolkata, in the development
of ceramic-based composite panels and involving authors in the experimental programme at TBRL, Chandigarh (India). This paper is being
published with the permission of the Director, CSIR-SERC, Chennai.
Address correspondence to Amar Prakash, Senior Scientist, CSIRStructural Engineering Research Centre, CSIR Campus, Chennai
600113, India. E-mail: amar@serc.res.in or amarsercm@gmail.com
Color versions of one or more of the figures in the article can be
found online at www.tandfonline.com/ucme.

chemical stability, and less weight. A ceramic-based composite panel is a composite system incorporating ductile metallic
plates bonded with hard ceramic tiles to defeat the ballistic projectile [15]. On impact, the kinetic energy of the projectile
is absorbed in fracturing the ceramic tile, deforming the target
and projectile, and increasing in temperature. The remaining
energy is absorbed by the metallic/composite backing, which
contains the remnants of the projectile and the ceramic debris
[6]. Ceramic/metal composite panels are extensively used in
lightweight armors. One of the limitations with ceramic/metal
composite targets under ballistic impact is that the ceramic tile
gets fragmented and thus the panels become vulnerable for the
next hits. This issue has been overlooked so far in the existing literature. Therefore, it is essential to understand the effect of joint
patterns in ceramic tiles to minimize the impact damage area.
In this way, the multi-hit resistance capability of ceramic/metal
composites can be enhanced significantly. Therefore, the motivation here is to reduce the damage area in targets under multiple
hits.
Many research studies have been reported on analytical, numerical, and experimental investigations on the highvelocity-impact performance of ceramic/metal composite panels. Willkins [1] first implemented numerical analysis of ceramic
armors subjected to normal impact using HEMP code. A few
other pioneering studies have been conducted by Hetherington
and Rajagopalan [2], Navarro et al. [3], Benloulo and Galvez
[4], and Fawaz et al. [5].
Zaera et al. [6] studied both numerically and experimentally the behavior of alumina/aluminum composite panels. It
has been reported that a thicker layer of adhesive causes a large
area to be affected by plastic deformation of the metallic plate in
the composite armor consisting of alumina tiles and aluminum
plate. Roeder and Sun [7] investigated the effects of structural
layering and thermal residual stresses on impact resistance of
alumina/aluminum laminated structures. Layered targets in various thicknesses have been tested under incident velocities in

192

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

CERAMIC/METAL COMPOSITE PANELS AND BALLISTIC IMPACTS

the range of 100 m/s to 300 m/s. It has been observed that
thick-layered laminates allow less penetration than thin-layer
laminates for the same areal density. Kaufmann et al. [8] have
conducted depth of penetration tests on four different ceramics,
out of which alumina ceramic outperformed silicon carbide and
boron carbide.
Holmquist-Johnson [9] reported a numerical and experimental study of impact performance evaluation of composite panel
made of SiC (front tiles) and aluminum alloy as the backing
plate. A steel projectile of diameter 7.62 mm and mass 8.32 g
impacted on the panels. Lopez et al. [10] studied both numerically and experimentally the effect of adhesive layer thickness
on the efficiency of alumina/aluminum armors using 7.62 AP
projectiles. Two configurations of the ceramic/metal composite panel have been studied for different tile sizes, of which
0.3 mm thickness of adhesive provided a better result. Mangapatnam [11] reported an experimental study of dynamic strength
of epoxy adhesive. Ramakrishna et al. [12] reported experimental work on an uncoated SiC/SiC composite and found that for
velocity above 300 m/s, the projectile penetrated through the
composite target. Hassan et al. [13] investigated the effect of
high-velocity impact of integral armour using the finite element
method. They have showed that the rubber composite interface
fails by delamination, believed to be due to interlaminar shear
stress rather than interlaminar tensile stress.
Sadanandan et al. [14] and Jena et al. [15] have studied
the effect of oblique impact. Sadanandan et al. [14] reported
that the ballistic limit velocity increases with obliquity. Beppu
et al. [16] reported the damage evaluation of concrete plate by
high-velocity impact. In their tests, failure processes of cratering and spalling were captured by a high-speed video camera.
Strabburger et al. [17] studied ballistic behavior of transparent armor ceramics. It was reported that protection efficiency
of ceramic/glass/polycarbonate targets increases as the ceramic
thickness increases. The modelling of high-strain-rate behavior
of materials for ceramic tiles and ductile backplate of metal are
reported in Johnson-Holmquist [18] and reviewed by Lamberts
[19]. Ubeyli [20] experimentally investigated the effect of different types of adhesives on the performance of Al2 O3 /Al2024laminated composite armors against 7.62 AP projectiles. The
results showed that polyurethane exhibited more resistance to
spalling of ceramic tiles than those bonded with epoxy. Karamis
et al. [21] studied the ballistic behavior of composite materials
subjected to high-velocity impact.
A number of recent research studies being conducted in the
field of ceramic/metal armor have shown a vast spectrum of
results. Fernandez et al. [22] have proposed a new constitutive
material model for simulating the behavior of material fragmentation under impact loading. Liu et al. [23] have shown
the method for the preparation of an interface. Ong et al. [24]
have simulated advanced personnel armor using the commercial software AUTODYN [35]. Savio et al. [25] have studied
the ballistic performance of boron carbide ceramic. Feli and
Asgari [26], Daniel et al. [27], and Tasdemirci et al. [28] have

193

focused on the numerical simulation of layered ceramic/metal or


ceramic/fiber composites. Chi et al. [29] have presented a semianalytical approach to studying the ballistic impact response
of the ceramic/metal composite panels. Compton et al. [30],
Kolopp et al. [31], and Gamble et al. [32] recently presented
similar studies of hybrid layered composite targets. Previous
research has shown that numerical simulations often neglect
the effect of the gilding jacket of the projectile; this issue was
studied by Hazell et al. [33], who showed that for correct predictions about ballistic performance, the bullet jacket should be
considered in the model.
From the reported studies, it is observed that emphasis in
prior research has been given to the optimization of the layers
thickness in the design of ceramic-based armours. Most of the
studies have been conducted to determine ballistic efficiency,
effect of ceramic tile thickness, ratio of front ceramic tile and
back plate thickness ratios, and effect of adhesive types and
thickness on impact performance of composite targets. There
is a scarcity of information on the effect of types of ceramic
tile joints on impact resistance and performance under multi-hit
scenarios. The reason for this may be the high demand on computational resources and their non-availability. However, due to
advancements in computational tools and resources, research on
numerical simulations of physical phenomena such as consecutive multi-hits is possible.
The aim of the present study is to find the effect of the
arrangement of front ceramic tiles having collinear and noncollinear joints on the impact damage pattern. The novelty of
the study presented in this article is that the prediction of highvelocity-impact response under two consecutive and closely
spaced hits on composite panels is carried out in a more realistic
manner.

2. DESCRIPTION OF THE PRESENT STUDY


2.1 Criteria for Selection of Material
The main requirements of materials involved in any protective system or armor design are [25]: low density, to reduce
the total weight of the armor; high bulk and shear modulus,
to prevent large deformations; high yielding stress, to preserve
armor resistance to failure; and high dynamic tensile stress, to
avoid material rupture when shock waves appear in the material
after impact. Metals usually fulfill all of the mentioned criteria, except for density. Ceramics, on the other hand, are brittle,
because they get fragmented extensively when tensile waves
appear in the material. However, they fulfill other requirements,
as mentioned above. It is obvious that a single material can not
satisfy all of the requirements. Therefore, ceramic tile backed by
a thin metal plate is the solution to circumvent the limitations of
ceramic and metal. Thus, a ceramic-metal composite combines
the lightweight and high impact resistance of ceramic with the
ductility of metallic materials at the back face.

194

A. PRAKASH ET AL.

25 mm

25 mm
Ceramic tiles

Fixed support

300 mm
Fixed support

300 mm

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

Ceramic tiles

Aluminium alloy backplate

Epoxy resin

300 mm

Aluminium alloy backplate

Epoxy resin

300 mm

(a) Non-collinear joint

(b) Collinear joint


FIG. 1. Typical target geometry.

2.2 Geometry Details


Impact responses of ceramic/metal composite panels under
normal impact of a 7.62 mm sharp-nosed steel projectile with
a caliber radius head of 3.0 are numerically predicted. Impact
velocity of the projectiles having mass of 10.3 g is considered
to be about 820 m/s (more than Mach 2). Two consecutive impacts are numerically generated to execute with a preset time
interval for their activation in succession. The spatial locations
of both impacts are separated to occur on two nearby ceramic
tiles of composite panels. The Lagrangian-approach-based finite element model is generated to simulate the high-velocity
impact on ceramic/metal composite panels using the commercial software AUTODYN [35]. The details of material models
related to high strain rate and large deformation phenomenon as
required for the present simulation are provided. Figure 1 shows
a typical arrangement of ceramic tiles in the composite target
with non-collinear joints and collinear joints.
2.3 Mechanism of Impact Resistance in Ceramic/Metal
Composite Panels
When a projectile impacts on a ceramic/metal composite
panel, the ceramic at the point of contact gets comminuted. This
powder-like material behaves like a fluid in the tip contact area.
That is why researchers, namely Tate-Alekseevski proposed an
equation to model the projectile behavior [3]. More precisely,
at high impact velocity the material behaves hydrodynamically
and is in a plastic state, and this equation is a modification of
Bernoullis equation. This proposed equation has been successfully used since 1966 and shows good agreement with other
methods of analysis. A comprehensive detailed description of
these formulations can be found in the literature [14]. The analytical solution using these equations needs assumptions for
simplification. It is assumed in the existing literature, with regards to analytical solutions, that only the back plate undergoes
deflection; failure modes such as tearing or petaling of the plate

are often neglected; the composite panel is assumed fixed at the


boundaries. Due to shock wave propagation, tensile weak zones
developed within the ceramic tile, which results in a conoid, as
shown in Figure 2a. The cross-section area of this conoid is more
than the projectiles [6]. Thus, the impact force gets distributed
on a larger area of the ductile backplate, as depicted in Figure 2b.
The adhesive layer bonds ceramic tile with metal backplate subjected to interface shear, which results in delamination at the
interface.

FIG. 2. Mechanism of impact resistance in ceramic/metal composite panels.

CERAMIC/METAL COMPOSITE PANELS AND BALLISTIC IMPACTS

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

The backplate undergoes various failure modes, such as


petaling, tearing, plugging, etc., depending upon the projectile energies, strength, and size. Normally, the sharp-nosed projectiles are converted to equivalent cylindrical projectiles in the
analytical approach [4], whereas the numerical simulations consider maximum features of complex geometries, materials, and
physics of the impact process. The details of the numerical
model developed in the present article are provided in the following sub-sections.

3. NUMERICAL SIMULATION
A numerical model using the Lagrangian approach is developed for the simulation of high-velocity impact carried out to
assess the impact resistance of a ceramic-based composite panel.
The details of the target, projectile, and support conditions are
provided in the following sections.

195

3.1 Geometry and Finite Element Discretization


Three-dimensional finite element meshes of the target panel
and projectile used in the finite element model are shown in
Figure 3. The minimum size of the element used is 0.5 mm.
The umber of eight-noded solid Lagrangian finite elements for
the projectile geometry and target bodies is about 14112 and
478085, respectively. After various trials on mesh size, the final
target discretized with uniform refined mesh, as shown in Figure 3, is used. To model the geometry of the non-collinear joint
pattern of ceramic tiles, the fragment/brick feature available in
AUTODYN [35] is adopted. This feature provides ease in creating a geometric model with staggering joints with number of
layers. Although the projectiles actual configuration comprises
thin casing and inner core, for the simulation a monolithic projectile is used to save computational time. It is kept in mind that
the momentum and kinetic energy of the projectile used in the
simulation must be the same as that used during the impact tests.

FIG. 3. Finite element model and mesh for target and projectile.

196

A. PRAKASH ET AL.

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

3.2 Interface Model


The front ceramic tile and ductile backplate of the target are
bonded with the epoxy resin. This bonded interface is modelled using the surface-to-surface contact feature available in
AUTODYN.
3.3 Boundary Conditions
Part of the bottom edge of the ductile backplate in the composite panel is fixed using a bench vice during an impact test.
A clear margin of 75 mm is provided between the front tile
edge and backplate edge of the composite panel (Figure 3). This
boundary condition is modelled by assigning three-dimensional
general velocities as zero for the selected bottom edge. Due
to this condition, any movement in the specified region gets
stopped, which is similar to that of the fixed boundary in quasistatic tests.
3.4 Material Models
It is vital to use improved material models and their input data
when modelling any material under shock loads to obtain better
results. Knowledge of exact model parameters over a wide range
of strain rates is therefore a case of absolute necessity in the
numerical simulation. Three important descriptors are needed
for representing material behavior under high-velocity impact,
as given below:

An equation of state (EOS) which relates the density


(or volume) and internal energy (or temperature) of the
material to pressure;
A constitutive relationship which describes the strength
of the material by relating the stress in the material to
the amount of distortion (strain) required to produce
this stress;
A failure model to predict the failure of material.
In addition to the above requirements, an erosion criterion is
also required as an effective numerical tool to handle severe
mesh distortion in both projectiles and target. However, it needs
careful tuning of the correct values of geometric strains for
the material so that numerical computation is not interrupted.
It is observed from the literature that different values of the
geometric erosion strains may be possible for same materials
because this is not a material property but a numerical technique
to remove severely distorted elements from computation.

3.4.1 Strength and Damage Model


For the inelastic behavior of brittle material (like ceramic)
whose strength is affected due to crushing, a suitable constitutive
model given by Johnson-Holmquist [18] is adopted. Strength
parameters such as yield and shear modulus can be reduced
with the damage as calculation proceeds.
A Johnson-Holmquist model (JH-Model) for both strength
and damage modelling in the ceramic tiles is used. This model
considers both compression and shear-induced strength in ceramic materials. Accumulated damage is computed as the ratio
of incremental plastic strain over the current estimated fracture
strain. The effective fracture strain is pressure-dependent. The
work done in deforming the material elastically in shear can be
converted into a pressure increase; hence volumetric dilatation
occurs. The amount of work which is converted into a dilatation
pressure is controlled through the bulking constant. This can
have values ranging from 0.0 (representing no shear-induced
distance) to 1 (maximum dilatancy effects). If the bulking constant is greater than zero, then the JH model should be used in
conjunction with a polynomial equation of state. The improved
tensile behavior of the model has been used to allow for principal tensile failure initiation in addition to the hydrodynamic
tensile limit. The crack-softening algorithm can also be used in
conjunction with principal stress failure criteria.
3.4.2 Materials Used
A Johnson Cook model [20] for strength and failure is
adopted for the steel used in the projectile. Ceramic (Al2 O3 99.5), epoxy resin, and aluminum alloy (Al5083 H116) are used
in the target composite panel. The material models used for these
materials are defined in Table 1. Instantaneous geometric-strainbased erosion criteria are considered to get rid of severely distorted elements. The input values provided for various material
parameters are given in Table 2.

4. DETAILS OF IMPACT TESTS


Impact tests were carried out on the ceramic-metal composite panel. The objective of the test was to investigate the
performance of the composite panel under the normal impact
of 7.62 AP projectiles. Penetration depths were measured using
conventional methods with adequate accuracy. The first hit was
considered at the center of the central ceramic tile in the target with collinear joints as shown in Figure 4a, followed by a

TABLE 1
Material models used in simulation
Descriptor
Equation of state
Strength
Failure
Geometric strain

Al2 O3 -99.5

Epoxy Resin

Steel 4340

Al5083 H116

Polynomial
Johnson-Holmquist
Johnson-Holmquist
2.0

Shock
Cowper-Symonds
Hydro (Pmin )
1.5

Linear
Johnson-Cook
Johnson-Cook
2.1

Linear
Johnson-Cook
Hydro (Pmin )
2.0

197

Polynomial
Reference Density (g/cc)
Bulk Modulus A1 (kPa)

Shock
Reference Density (g/cc)
Gruneisen Coefficient
Parameter C1
Parameter S1

Linear
Reference Density (g/cc)
Bulk Modulus (kPa)

Linear
Reference Density (g/cc)
Bulk Modulus (kPa)
Reference Temperature (K)
Specific Heat (J/kgK)
Thermal Conductivity (J/mKs)

EPOXY RESIN

STEEL 4340

ALUMINUM
ALLOY
AL5083H116

Equation of State

ALUMINA
AL2 O3 -99.5

Material
Descriptor
Failure Parameter

2.70E+00
5.83E+07
2.93E+02
9.10E+02
0.00E+00

Johnson-Cook
Shear Modulus (kPa)
Yield Stress (kPa)
Hardening Constant (kPa)
Hardening Exponent
Strain Rate Constant
Thermal Softening Exponent
Melting Temperature, K
Ref. Strain Rate (/s)

2.69E+07
1.67E+05
5.96E+05
5.51E-01
1.00E-03
8.59E-01
8.93E+02
1

7.70E+07
7.92E+05
5.10E+05
2.60E-01
1.40E-02
1.03E+00
1.79E+03

Cowper-Symonds
Shear Modulus (kPa)
1.60E+06
Yield Stress (kPa)
4.50E+04

Johnson-Cook
7.86E+00 Shear Modulus (kPa)
1.59E+08 Yield Stress (kPa)
Hardening Constant (kPa)
Hardening Exponent
Strain Rate Constant
Thermal Softening Exponent
Melting Temperature, K

1.186
1.13
2730
1.493

5.00E-02
3.44E+00
2.12E+00
2.00E-03
6.10E-01
1.79E+03
1
Hydro (Pmin )
Hydro Tensile Limit (kPa)
1.50E+06

Johnson-Cook
Damage Constant, D1
Damage Constant, D2
Damage Constant, D3
Damage Constant, D4
Damage Constant, D5
Melting Temperature (K)
Ref. Strain Rate (/s)

Hydro (Pmin )
Hydro Tensile Limit (kPa)
1.50E+05

Johnson-Holmquist
Johnson-Holmquist
3.8
Shear Modulus (kPa)
1.35E+08
Hydro Tensile Limit (kPa)
2.90E+04
2.00E+08 Model Type
Continuous (JH2) Model Type
Continuous (JH2)
Hugoniot Elastic Limit (kPa)
5.90E+06
Damage Constant, D1
0.001
Damage Constant, D2
1
Bulking Constant
1
Damage Type
Gradual (JH2)
Tensile Failure
Hydro (Pmin)

Strength Parameter

TABLE 2
Material input parameter used for various models

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

198

A. PRAKASH ET AL.

TABLE 3
Comparison of depth of penetrations
Description
Numerical Prediction
Numerical Simulation (with offset in
hit locations bothways)
Experimental

Impact velocity (m/s)

Location of impact

DOP into back


plate (mm)

Joint type in
ceramic tiles

820.9
831.1
820.9
831.1
820.9
831.1

Center of central tile


Center of corner tile
Center of central tile
Center of corner tile
Center of central tile
Center of corner tile

4.1
7.2
6.5
5.1
2.4#
6.3

Collinear joints
Collinear joints
Non-collinear joints
Non-collinear joints
Non-collinear joints
Non-collinear joints

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

This value need to be checked because the comunited ceramic material was found embedded in the crator hole.

second hit at the lower right-side tile. For non-collinear jointed


tiles, the location of hit 1 and hit 2 is shown in Figure 4b.
An impact experiment was conducted on the target having
non-collinear jointed tiles using the test set-up as shown in
Figure 4c.

4.1 Procedure for Impact Test


Well-wrapped ceramic-metal composite panel was
mounted on the bench vice (Figure 4c). One of the
edges of the panel was gripped between the jaws of
the bench vice and the other edges were kept unconstrained.
Arrangements were made to determine the velocity of
the projectile before impact.
Adequate attention was paid before triggering the
7.62AP projectile to ensure that no person stayed near
the composite panel.
First, the composite panel had an impact having projectile velocity 820.9 m/sec at the center of the panel
where two tile edges meet, as shown in Figure 4b.

Then, on this damaged composite panel, a second hit


with an impact velocity of 831.1 m/s was made at
location Hit 2, as shown in Figure 4b.
Measurements were taken for the depth of penetration
with the help of a special gauge. The criteria related
to the DOP measurement as reported (NIJ-[34]) was
followed.
Thin wrapping was removed and the composite panel
along with the debris of the fractured tile were photographed as shown(Figure 6c).
5. RESULTS AND DISCUSSION
5.1 Prediction of Impact Response for Target with
Collinear Jointed Tiles (Symmetric Hits)
On first impact at the location marked hit 1 (Figure 4), the
planer stress wave moves in-plane and starts reflecting from the
midpoints of the edges at the same time due to equal distance
from the point of impact. However, it reaches the corner later
in time due to the longer distance as compared to the midpoints
on the edges. During this time lag, stress wave transmission at

FIG. 4. Ceramic tile patterns: (a) collinear joint; (b) non-collinear joint; (c) test set-up.

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

CERAMIC/METAL COMPOSITE PANELS AND BALLISTIC IMPACTS

199

FIG. 5. Damage contours for collinear joints for consecutive hits.

the midpoint of all four edges of the impacted tile takes place.
This transmitted stress wave, originating at the interface, moves
as a circular wave front and is met with edges normal to the
direction of propagation in the adjacent tile. Because of the
early reflection from the sides, a longitudinal crack develops
first and a transverse crack appears later. Transverse cracks in
the adjacent tiles result due to the late reflection of the stress
wave from the distal edge. Similar central tensile cracks develop
in all four adjacent tiles, as shown in Figure 5.
The depth of penetrations obtained from numerical and experimental studies are compared in Table 3. A predicted response for depth of penetration in the case of collinear joints
is obtained as 4.1 mm and 7.2 mm for the first and second
hits, respectively. The corresponding damage contours for front
ceramic tiles are shown in Figure 5.

5.2 Simulation of Impact Test on Target with


Non-Collinear Jointed Tiles (Asymmetric Hits)
To simulate the experimental behavior, the eccentric impact
location is deliberately considered; for this reason, the transmitted stress wave reaches with a delay in the downward ceramic
tile and very few cracks appear, even at the end of 84 s (which
is considered as the end of the first hit response). The main crack
in the bottom tile is mainly due to the reflection from sides only
because the stress wave could not reflect by the time kinetic
energy in the first hit was dissipated.
The damage contours obtained numerically and cracked tiles
experimentally after two consecutive hits are shown in Figure 6.
The developed cracks in ceramic tiles with collinear and noncollinear joints show different patterns.

FIG. 6. Damage contours of non-collinear joints for consecutive hits.

A. PRAKASH ET AL.

FIG. 7. Depth of penetration into the back plate in the composite target (side views).

5.3 Depth of Penetration (DOP)


As per numerical prediction, the projectile is found to penetrate into the backplate to a depth of 6.5 mm and 5.1 mm for the
first hit and second hit, respectively. The experimental values
of penetration depth for first and second hits of non-collinear
jointed tiles were measured as 2.4 mm and 6.3 mm, respectively. It is noted that the comminuted ceramic material was
embedded into the aluminum back plate. Due to this, the reported measured depth of penetration for a central hit of 2.4 mm
needs the proper removal of foreign material for actual DOP in
the aluminum back plate. One of the possible reasons for the
difference between experimental and numerical values of DOP

may be due to the covering of the ceramic tiles by thin linen


in the experiment. Although it was assumed that the thin linen
wrap would not affect impact behavior, it seems to have little
influence on response as per the experimental result. Negligible
bulging (less than 1 mm) at the back face is observed for both
the non-collinear and collinear joint cases, as shown in Figure 7.

5.4 Kinetic Energy Variation


The kinetic energy (KE) variation during both consecutive
hits is shown in Figure 8. Based on the preliminary trial, the time
of cut-off is determined to make the second hit. It can be seen

4.0

4.0
First hit

3.0

Second hit

Kinetic energy, kJ

Kinetic Energy, kJ

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

200

2.0
1.0
0.0

0.00

0.04

0.08
0.12
Time, ms

Collinear joint

0.16

First hit

3.0

Second hit

2.0
1.0
0.0
0.00

0.04

0.08

Time (ms)
Non-collinear joint

FIG. 8. Kinetic energy of the projectile for non-collinear ceramic joints.

0.12

CERAMIC/METAL COMPOSITE PANELS AND BALLISTIC IMPACTS

that the projectiles entire kinetic energy has been dissipated


into various mechanisms, namely deformation of target, heat,
etc. Therefore, the next hit is generated after this time interval.
The pattern of KE dissipation history is found to be the same
for both cases, as shown in Figure 8.
6. SUMMARY AND CONCLUSIONS
Simulations of two consecutive normal impacts of 7.62 mm
sharp-nosed projectiles on ceramic/metal composite panels are
presented. The two impacts are numerically generated with a
preset delay in their activation time.

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

It is observed that the ceramic tile arrangement and


impact locations influence the damage caused due to
high-velocity multiple impacts on adhesively bonded
ceramic/metal composite targets.
Damage contours obtained from the simulations indicate the tensile failure of material at a scale from 0 to 1,
and exhibit the shock wave propagation phenomenon.
One can clearly identify the effect of joints from these
contours.
Maximum depth of penetration found for collinear
joints was 7.2 mm at the corner tile hit, whereas for
non-collinear joints it was 5.1 mm, compared with the
experimental value of 6.3 mm. Therefore, a variation
in the depth of penetration obtained is in the range of
1419%.
An eccentricity in the location of actual impact during
tests has been observed from the crack patterns obtained. Therefore, to simulate this behavior, a suitable
offset regarding actual impact locations in tests has
been considered.
Both experimental and numerical responses exhibited
a close agreement in terms of depth of penetration as
well as crack patterns typical for a non-collinear joint
pattern of tiles. Further investigations are necessary to
arrive at the optimum size of the tile and effective joint
patterns for reducing the damage in the target.
The understanding about the damage under consecutive impacts and the crack patterns can provide significant input for designing protective ceramic/metal
composite panels.

REFERENCES
1. M.L. Wilkins, Mechanics of Penetration and Perforation, Intl. Journal Engineering Science, vol. 16, pp. 793807, 1978.
2. J.G. Hetherington and B.P. Rajagopalan, An Investigation into the Energy
Absorbed During Ballistic Perforation of Composite Armors, International
Journal Impact Engineering, vol. 11(1), pp. 3340, 1991.
3. R. Zaera and G.V. Sanchez, Analytical Modelling of Normal and Oblique
Ballistic Impact on Ceramic/Metal Lightweight Armours, International
Journal of Impact Engineering, vol. 21, pp. 133148, 1998.
4. C. Benloulo and G.V. Sanchez, A New Analytical Model to Simulate Impact
onto Ceramic/Metal Composite Armours, International Journal of Impact
Engineering, vol. 21, pp. 461471, 1998.

201

5. Z. Fawaz, W. Zheng, and K. Behdinan, Numerical Simulation of Normal


and Oblique Ballistic Impact on Ceramic Composite Armours, Composite
Structures, vol. 63, pp. 387395, 2004.
6. R. Zaera, S.S. Sanchez, C. Perez, and C. Navarro, Modelling of Adhesive
Layers in Mixed Ceramic/Metal Armours Subjected to Impact, Composite:
Part A, vol. 31, pp. 823833, 2000.
7. B.A. Roeder and C.T. Sun, Dynamic Penetration of Alumina/Aluminum
Laminates: Experiments and Modeling, International Journal of Impact
Engineering, vol. 25, pp. 169185, 2001.
8. C. Kaufmann, D. Cronin, M. Worswick, G. Pageau, and A. Beth, Influence
of Material Properties on the Ballistic Performance of Ceramics for Personal
Body Armour, Shock and Vibration Journal, vol. 10(1), pp. 5159, 2003.
9. T.J. Holmquist and G.R. Johnson, Characterization and Evaluation of Silicon Carbide for High-Velocity Impact, Journal of Applied Physics, vol. 97,
pp. 493502, 2005.
10. J. Lopez-Puente, A. Arias, R. Zaera, and C. Navarro, The Effect of Thickness of Adhesive Layer on the Ballistic Limit of Ceramic/Metal Armours,
An Experimental and Numerical Study, International Journal of Impact
Engineering, vol. 32, pp. 321336, 2005.
11. A. Mangapatnam and V. Parameswaran, Dynamic Strength of Adhesive
Single Lap Joints at High Temperature, International Journal of Adhesion
and Adhesives, vol. 28, pp. 321327, 2008.
12. T. Ramakrishna Bhatt, R. Sung Choi, M.L. Cosgriff, S.D. Fox, and N. Kang
Lee, Impact Resistance of Uncoated SiC/SiC Composite, Materials Science
and Engineering, vol. 476, pp. 2028, 2008.
13. M. Hassan, Y. Zhu, A. Haque, A. Abutalib, U. Vaidya, S. Jeelani, B. Gama,
and B. Gillespie Fink, Investigation of High-Velocity Impact on Integral
Armor Using Finite Element Method, International Journal of Impact Engineering, vol. 24, pp. 203217, 2000.
14. S. Sadanandan and J.G. Hitherington, Characterization of Ceramic/Steel
and Ceramic/ Aluminium Armours Subjected to Oblique Impact, International Journal of Impact Engineering, vol. 19, pp. 811819, 1997.
15. P.K. Jena, N. Jagtap, K. Sivakumar, and B.T. Balakrishna, Some Experimental Studies on Angle Effect in Penetration, International Journal of
Impact Engineering, vol. 37, pp. 489501, 2010.
16. M. Beppu, K. Miwa, M. Itoh, M. Katayama, and T. Ohno, Damage Evaluation of Concrete Plate by High-Velocity Impact, International Journal of
Impact Engineering, vol. 35, pp. 14191426, 2008.
17. E. Strabburger, Ballistic Testing of Transparent Armour Ceramics, Journal
of the European Ceramic Society, vol. 29, pp. 267273, 2009.
18. T.J. Holmquist and G.R. Johnson, Modeling of Ceramic Dwell and Interface
Defeat, Ceramic Armor Material Design, pp. 309316, 2001.
19. A.P.T.M.J. Lamberts, Numerical Simulation of Ballistic Impacts on Ceramic <aterial, MT07.33, Department of Mechanical Engineering, Materials Technology PDE Automotive B.V. CAE, Eindhoven Univ. of Tech,
http://www.mate.tue.nl/mate/pdfs/8206.pdf.
20. M. Ubeyli, O.R. Yildirim, and O. Bilgehan, Investigation on the Ballistic
Behavior of Al2 O3 /Al-2024 Laminated Composites, Journal of Material
Processing Technology, vol. 196, pp. 356364, 2008.
21. M.B. Karamis, F. Nair, and A.A. Cerit, The Metallurgical and Deformation Behaviors of Laminar Metal Matrix Composite after Ballistic Impact,
Journal of Material Processing Technology, vol. 209, pp. 48804889, 2009.
22. D.F. Fernandez, R. Zaera, and J. Saez, A Constitutive Equation for Ceramic
Materials Used in Lightweight Armors, Computers and Structures, vol. 89,
pp. 23162324, 2011.
23. G. Liu, C. Ni, Q. Xiao, F. Jin, G. Qiao, and T. Lu, Preparation and Interface Structures of Metal-Encased SiC Composite, Armors with Interpenetrating Structure, Rare Metal Materials and Engineering, vol. 40(12),
pp. 20762079, 2011.
24. C.W. Ong, C.W. Boey, R.S. Hixson, and J.O. Sinibaldi, Advanced Layered
Personnel Armor, International Journal of Impact Engineering, vol. 38(5),
pp. 369383, 2011.
25. S.G. Savio, K. Ramanjaneyulu, V. Madhu, and T. Balakrishna Bhat, An
Experimental Study on Ballistic Performance of Boron Carbide Tiles, International Journal of Impact Engineering, vol. 38(7), pp. 535541, 2011.

202

A. PRAKASH ET AL.

Downloaded by [Indian Institute of Technology Roorkee] at 04:23 30 June 2015

26. S. Feli and M.R. Asgari, Finite Element Simulation of Ceramic/Composite


Armor Under Ballistic Impact, Composites Part B: Engineering, vol. 42(4),
pp. 771780, 2011.
27. B. Daniel, R.F. Alfredo, F.M.A. Sergio, C.L.M. Francisco, and V.D.
Maurcio, Ballistic Impact Simulation of an Armour-Piercing Projectile on Hybrid Ceramic/Fiber Reinforced Composite Armours, International Journal of Impact Engineering, vol. 43, pp. 6377,
2012.
28. A. Tasdemirci, G. Tunusoglu, and M. Guden, The Effect of the Interlayer
on the Ballistic Performance of Ceramic/Composite Armors: Experimental
and Numerical, International Journal of Impact Engineering, vol. 44, pp.
19, 2012.
29. R. Chi, A. Serjouei, I. Sridhar, and G.E.B. Tan, Ballistic Impact on Bi-layer
Alumina/Aluminium Armor: A Semi-analytical Approach, International
Journal of Impact Engineering, vol. 52, pp. 3746, 2013.

30. B.G. Compton, E.A. Gamble, and F.W. Zok, Failure Initiation During Impact of Metal Spheres onto Ceramic Targets, International Journal of Impact
Engineering, vol. 55, pp. 1123, 2013.
31. A. Kolopp, R. Samuel, and B. Christophe, Experimental Study of Sandwich
Structures as Armour Against Medium-velocity Impacts, International Journal of Impact Engineering, in press.
32. E.A. Gamble, B.G. Compton, and F.W. Zok, Impact Response of Layered
SteelAlumina Targets, Mechanics of Materials, vol. 60, pp. 8092, 2013.
33. P.J. Hazell, G.J.A. Thomas, D. Philbey, and W. Tolman, The Effect of Gilding Jacket Material on the Penetration Mechanics of a 7.62 mm Armourpiercing Projectile, International Journal of Impact Engineering, vol. 54,
pp. 1118, 2013.
34. National Institute of Justice (NIJ), Standard-0101.06, Ballistic Resistance
of Body Armor, July 08, http://www.ojp.usdoj.gov/nij.
35. AUTODYN, Release 12.1, ANSYS, Inc., Canonsburg, PA, USA, 2009.

Potrebbero piacerti anche