Sie sind auf Seite 1von 11

Int. J. Impact Engng Vol. 21, No. 6, pp.

461471, 1998
( 1998 Elsevier Science Ltd. All rights reserved
Printed in Great Britain
PII: S0734743X(98)000062 0734743X/98 $19.00#0.00

A NEW ANALYTICAL MODEL TO SIMULATE IMPACT


ONTO CERAMIC/COMPOSITE ARMORS
I. S. CHOCRON BENLOULO* and V. SANCHEZ-GALVEZ
Departamento de Ciencia de Materiales, E.T.S.I. Caminos, Canales y Puertos, Universidad
Polite& cnica de Madrid, Madrid 28040, Spain

(Received 11 June 1997; in revised form 5 September 1997: in final form 8 January 1998)

SummaryA very simple one-dimensional and fully analytical model of ballistic impact against
ceramic/composite armors is presented in this paper. The analytical model has been checked both
with ballistic tests and numerical simulations giving predictions in good agreement with them. The
model allows the calculation of residual velocity, residual mass, and the projectile velocity and the
deflection or the strain histories of the backup material. These variables are important in describing
the phenomenological process of penetration. Described are modifications to previous work of
impact into ceramics combined with a new composites model. The development of this composite
model is based on studies of the impact in yarns, fabrics and finally composites. ( 1998 Elsevier
Science Ltd. All rights reserved

Keywords: ballistic impact, ceramic/composite armor, analytical model.

NOTATION
A percentage of volume which has been delaminated
A cross-sectional area of the projectile
1
APDS armor piercing discarding sabot
c ceramic longitudinal sound speed
c longitudinal speed of sound in the fabric yarns
:
d areal density of the composite
e strain on the fabric under the projectile
E Young modulus of the yarns
F force exerted by each yarn on the ceramic cone
FSP fragment simulating projectile
G energy necessary to delaminate 1 m2 of composite
h ceramic plate thickness
characteristic dimension of the composite panel
M mass of the ceramic cone
#
M mass of the projectile
1
M mass of the projectile at the end of the first phase
11
n number of layers of fabric
-
n number of yarns under the projectile
:
h angle between the yarn and the direction of impact
R failure constant
o ceramic density
#
o projectile density
1
S section of the yarn
t time in seconds, zero is the instant of contact between armor and projectile
t time when the first phase has finished
1
t time of failure of the fabric
&
u5 (t) velocity of the ceramic cone
0
velocity of the projectile at any time
ballistic limit
50
residual velocity
3
striking velocity
4
mass of the projectile in grams.
x5 (t) velocity of the projectileceramic interface

*Corresponding author. Current address: Southwest Research Institute, Engineering Dynamics Department,
PO Box 28510, San Antonio, TX 78228-0510, U.S.A.
Tel.: (210)-522-2657; fax: (210) 522 3042; email: schocron@swri.edu.

461
462 I. S. C. Benloulo and V. Sanchez-Galvez

@ dynamic strength of the fragmented ceramic


#
dynamic strength of the intact ceramic
#
dynamic elastic limit of the projectile
1

1. INTRODUCTION
Ceramic backed by composite armors are becoming the subject of many investigations
because their performance against small and medium caliber projectiles is outstanding
when the weight is a design condition, for instance in light weight vehicles, airplane and
helicopter protection or body armors. The main role of the ceramic is the erosion and
rupture of the projectile. The composite absorbs the kinetic energy of the fragments
stopping them. The design of these armors is really complex and need many sophisticated
tools. Empirical methods are the most widely used ones because they offer reliability, but
they are extremely expensive and the results did not give enough information: the history of
the projectile, the trends when changing the configurations or the phenomenological
process cannot be obtained in detail with the experimental approach. Another way to face
the problem is to use hydrocodes to simulate numerically the physical process. Either finite
elements or finite difference schemes need many parameters for material description, which
very often are not even known, for a correct calculation. For instance, the impact of
a tungsten projectile against a ceramic/composite armor may need as many as 50 different
mechanical parameters, including the elastic constants for each material (in all directions),
dynamic yield stress and others without clear physical meaning such as erosion strain, etc.
Another problem is that the calculations are very long and usually need many hours or even
days of powerful computers like workstations. The numerical approach provides a lot of
information but again cannot give trends unless multiple configurations are calculated and
consequently making the design of an armor long and hard. The third approach is the
analytical one. Provided the appropriate assumptions it could be a simple and fast way
allowing to obtain phenomenological information of the penetration mechanics without
losing much accuracy with respect to the numerical models. To the knowledge of the
authors only Florences model [1] appears in the literature as a simple analytical model
being used to analyze impact onto ceramic/composite armor. Interesting tests may be found
in Refs. [24].
The analytical model presented in this paper begins with the problem of the impact of
a projectile into the ceramic already extensively treated in the literature. Then the composite
model is faced with the logic reductionism: yarnfabriccomposite. Combining the two
models a ceramic/composite model is derived and checked against experimental data and
numerical simulations.

2. ANALYTICAL MODEL
The extremely complex problem of a projectile impacting a ceramic backed by a metal
plate has already been studied by Florence [1] in 1969 and more recently by Woodward
[5], den Reijer [6], Hetherington [7] and Zaera [8]. It is not the objective of this paper to
improve their ceramic impact models but simply to use them, together with the model of
impact onto fabrics developed by the authors [9], to build a ceramic/composite one. In the
following the reader can find the most important results obtained by the authors and how
they have been applied to this problem. Two phases have been assumed during the impact.

2.1. First phase


2.1.1. Projectile model. The ceramic/composite analytical model presented in this paper has
been checked with Tungsten projectiles (20, 25 and 30 mm APDS) which are the most
common threats add-on armors have to fragment. Such projectiles are assumed to be rigid
perfectly plastic with a dynamic yield stress of 1. Their geometry is supposed to be
cylindrical.
Analytical model for ceramic/composite armors 463

Nevertheless, the composite model has been checked independently, as a previous step to
the ceramic/composite model, with very different projectiles, usually FSPs, ranging from
0.2 g to tens of grams showing good agreement with experiments, see for instance, Refs.
[910]. In this latter case the steel projectiles were assumed to be rigid.

2.1.2. Ceramic model. When a ceramic is impacted by a projectile a compressive wave


travels from the front to the rear face at the speed of sound, then reflects and becomes
a tensile wave which breaks the ceramic in tension while coming back. Den Reijer [6]
assumes that a ceramic cone is generated at t"6(h/c), h is the thickness of the ceramic tile
and c its longitudinal speed of sound, with an angle of 65, see Ref. [1]. During the
formation of the cone the projectile is being eroded but the ceramic does not move at all.
The part of the projectile that is being eroded is called the plastic part and in this phase its
penetration velocity is x5 "0. The rear of the projectile moves at a velocity (t). This
velocity is governed by Tates equation [11]
d
M "! A . (1)
1 dt 1 1
The additional equation which governs this phase is the geometrical condition for the
projectile:
dM
1"!o A , (2)
dt 1 1
where A1 is the cross-sectional area of the projectile. The initial condition of this ordinary
differential equation is the initial mass of the projectile. Eqns. (12) may solve the first phase
of the problem (Fig. 1).

2.2. Second phase


It starts at t"6 (h/c) and now the whole armor contributes to the slowing down of the
projectile: the rear of the projectile moves at a speed (t), the ceramicprojectile interface
at x5 (t) and the cone at u5 (t). The difference between (t) and x5 (t) gives the erosion rate of
0
the projectile and the difference between x5 (t) and u5 (t) the penetration of the projectile into
0
the ceramic cone (the subscript zero does not mean an initial condition but it is kept to
respect den Reijers notation). Figure 2 explains the mechanisms governing this phase. The
forces F acting on the ceramic cone will be explicitly written in the composite equations
paragraph.

2.2.1. Projectile equations. Depending on the velocity of the ceramicprojectile interface


x5 (t) the tip of the projectile will be flowing or not. Two possibilities have then to be taken
into account in the equations:

2.2.1.1. First case: xR (t)((t), rigid plastic projectile. There is a plastic zone where the
projectile is being eroded. Tate and Aleksevskii equation is applicable
#1 o (!x5 )2"@ #1 o (x5 !u5 )2 , (3)
1 2 1 # 2 # 0

Fig. 1. Configuration at the end of the first phase. Fig. 2. Phenomenological description of the second
phase.
464 I. S. C. Benloulo and V. Sanchez-Galvez

@ is the dynamic strength of the broken ceramic. For a careful study of the terms like
#
and @ the reader should consult Ref. [12].
1 #
Deceleration of the projectile is found from Tates equation:
d
M (t) "! A , (4)
1 dt 1 1
where at t"t is , the initial velocity in this phase which coincides with the final
1 1
velocity of the first phase. The last equation to be written for the projectile in the first case is
the following geometrical condition:
dM
1"!o A (!x5 ), (5)
dt 1 1
where M at t"t is defined as M , the projectile mass at the end of Phase 1.
1 1 11
2.2.1.2. Second case: xR (t)"(t), rigid projectile. When solving Eqns (3)(5) it could
happen that the velocity of the projectileceramic interface equals or even becomes larger
than the projectiles one. This has obviously no physical sense: the reality is that the
projectile is no longer flowing, so Eqns (35) are no longer applicable. Now, the interface
and the projectile move at the same speed
x5 (t)"(t) (6)
and the force slowing the projectile is exerted by the ceramic. Two subcases should be
considered.
Subcase 1: The velocity of the cone is smaller than that of the projectile (u5 (t)((t)),
0
hence the projectile is penetrating into the ceramic cone until the ceramic is completely
eroded. The formula governing this subcase is Newtons equation
d
M "!@ A , (7)
12 dt # 1
where now the force on the projectile is exerted by the pulverized ceramic and the mass of
the projectile is constant and equal to the mass at the end of the first case.
Subcase 2: It could happen that the ceramic cone, not yet being totally eroded (cone mass
more than 0.1 g), reaches projectiles velocity. Then the two bodies, as a single projectile, will
penetrate the composite backup. To simplify the mathematical aspect, the kinetic energies
of the ceramic cone and projectile are added and a new homogenous projectile with the
total mass and the same total kinetic energy is assumed to impact the composite. The
discontinuity obtained in the velocity is very small because projectile mass is much higher
than the ceramic cone mass.

2.2.2. Ceramic equations. The ceramic cone begins to move at the end of the first phase. It is
pushed at one end by the projectile while, at the back, it is being retained by the composite
backup. Figure 3 shows the forces acting on the cone during the impact process. The
composite material is a textile fabric, for instance, aramid or polyethylene yarns, embedded

Fig. 3. Forces acting on the ceramic cone during the impact.


Analytical model for ceramic/composite armors 465

in an organic matrix, like polypropylene or polyethylene. The matrix is supposed to break


into pieces which stick to the yarns becoming only dead weight for them, implying that the
principal mechanism for slowing down the projectile is the force F exerted by the yarns.
Only warp yarns have been drawn in the figure but weft ones are also taken into account in
the model.
The equation governing the motion of the cone is Newtons equation, but now the mass is
time dependent:
d(M u5 )
# 0 u5 "!2Fu5 cos h!Gn2AQ #@ A u5 ,
dt 0 0 # 1 0
u5 (t"t )"0. (8)
0 1
It is pointed out that Newtons equation has been multiplied by cone velocity in order to
obtain an energy equation. The first and second term in the right-hand side of the equation
will be explicitly explained in the paragraph of the composite equations, they are respon-
sible for the force of the composite on the ceramic and the energy lost by delamination,
respectively. The third term is the force of the projectile on the ceramic. M is the mass of the
#
ceramic cone. Its value can vanish converting the problem in an impact of a projectile onto
a composite.

2.2.3. Composite equations. In this section a new model, based on the previous works
published by Roylance [13], Cunniff [14], Beaumont [15] and Navarro [16] will be
presented. A simple analytical model of impact onto a fabric has already been published by
Chocron [9] including the basis of the model which is now expanded to ceramic backed by
composite materials. The model has been checked with alumina/aramid and alumina/
polyethylene giving good results when compared with analytical and numerical ones. It
basically states that the kinetic energy of the projectile is transmitted to the backup plate
through two mechanisms: straining and breaking of the yarns, and delamination. The next
two paragraphs focus on these events.
When a point projectile impacts a linear elastic yarn the velocity u5 of the projectile and
0
the strain e are related by

u5 "c J 2eJe(1#e)!e2 . (9)


0 :
This equation is the analytical solution found by Smith [17] for constant velocity impact
and is also valid for nonconstant velocity provided it is only used at the point of impact as
proved by Chocron in Ref. [9]. c is the longitudinal speed of sound in the yarn. Smith also
:
found the angle between the line of impact and the yarn

J2eJe(1#e)!e2
sin h" . (10)
Je(1#e)
The force the yarns exert on the projectile can easily be calculated with
F"EeSn n , (11)
- :
where E is the Youngs Modulus, S the section of the yarn, n the number of layers and h is
-
shown in Fig. 2. n is the number of yarns directly in contact with the impacting body. All
:
the yarns are assumed to have the same strain through the textile thickness to keep the
analytical model simple enough, weave and crimp are not considered. Now that the first
term in the right-hand side of Eqn (8) has become clear the second shall be explained.
Beaumont in his Ph.D. Thesis [15] studied low-velocity impact, around 100 m/s, into
brittle matrix composites and found a law to predict the evolution of delamination during
the impact process: it is assumed that the delamination travels together with the transversal
wave in the fabric, then

A BG C A A B D H
dA 2@3 ~1 1@3
AQ , "4 c 2 1!5 JA (12)
dt : 4c
:
466 I. S. C. Benloulo and V. Sanchez-Galvez

where A is the percentage of volume delaminated and is a characteristic length of the


backup (if it is circular then is the diameter). The term n 2GAQ in Eqn (8) accounts for the
energy lost per unit of time due to delamination. G is the energy required to delaminate 1 m2
of composite in J/m2. This term is only important at low-impact velocities, for high-impact
velocities, at 1000 m/s or more, the transversal wave has no time to progress before the
failure of the panel and then the energy absorbed by delamination is very small (less than
1% of the initial kinetic energy of the projectile).

2.2.4. Failure of the composite backup. A failure criteria of the composite should be used in
order to stop the model calculations when penetration occurs. As it is assumed that the
failure happens just under the projectile a maximum strain or strength criteria would not
work: according to Smiths equation (9) the higher the impact velocity, the greater the strain
implying that both maxima coincide at the beginning of the impact process (when the
velocity of the projectile is the highest). If a maximum strain criterion is applied it will give
failure at the very first microsecond, which has no physical meaning.
A failure model based on the energy absorbed until failure by the backup is another way
to calculate the instant of failure. In this work one part of the energy is supposed to be
absorbed by delamination and the other one by the fabric, as elastic energy. The latter one is
supposed to be a constant amount of energy when failure occurs, independently of the
impact velocity. A similar hypothesis has already been used by Prosser [18]. Then if the
mass of the projectile is constant, as for instance in the case of the impact of a projectile into
a textile fabric
2!2"2 (13)
4 3 50
where is the striking velocity of the projectile, the residual velocity and the ballistic
4 3 50
limit.
The equation of energy for the projectile may be written as

P
1 1 1 5
M 2! M 2(t)" n n SEc e2 dt, (14)
2 1 4 2 1 2 : - :
0
which means that the kinetic energy lost by the projectile at any time t has been dissipated
through the yarns under it (e is the strain just under the projectile). Friction between yarns
or between the projectile and the fabric has been neglected in order to gain simplicity. If t is
selected as the failure instant and, as the energy until failure is considered constant
independently of the striking velocity, then

P
1 5&
n n SEc e2 dt"const., (15)
2 : - :
0
where t is the time of failure. Another way to write Eqn (15) is
&

P
5& M 2
e2 dt" 1 50 ,R, (16)
n n SEc
0 : - :
where R is called the failure constant. Thus, knowing (or, in fact, any residual velocity
50
from one single firing test), it is easy to calculate the time of failure for any impact velocity
because R will not change if the configuration of the armor does not change. It is suggested
not deriving R directly from the material properties because of the lack of accuracy in
measuring the section, the modulus (that could depend on the strain rate, see Ref. [19]), or
. With one single test the designer can calculate R, by forcing the analytical and
50
experimental residual velocity to be equal, and then use it to analyze all impact velocities.
Eqns (13)(16) are a summary of the failure model for impact into textile fabrics already
published by Chocron [9]. A more general failure model is easily derived: as R depends on
the configuration, let R be the failure constant for the target configuration i and R be the
i 0
one for a reference target configuration, from Eqn (16) it follows:
(M ) n (n ) (2 )
R "R 1i :0 - 0 50 i . (17)
i 0 (M ) n (n ) (2 )
10 :* - 0 50 0
Analytical model for ceramic/composite armors 467

This formula can be very useful if some empirical value for is assumed or found
50
experimentally. For instance, Van Gorp [20] found that Dyneema (or Spectra) armors
verify for fragment simulating projectiles:
"232 d0.5~1@6 , (18)
50
where d is the areal density of the panel and the mass of the projectile in g.
Then Eqn (17) may be rewritten as
(M )2@3 n
R "R 1* :0 , (19)
* 0 (M )2@3 n
10 :*
which only depends on the mass of the projectile and the number of yarns under it (say the
number of yarns per meter in the fabric). Equation (19) allows the calculation of the failure
constant for any configuration (the projectile, the thickness of the backup panel, the striking
velocity can be changed) if one firing test is provided in order to calculate firstly R .
0
The fabric failure model presented above has been, in the same way, applied for the
analysis of impact onto ceramic/composite armors taking into account that it is only a first
approximation to the actual process. Now the reference case for calculating R should be
0
a ceramic/composite one and the mass considered is not the projectiles one but the mass of
the impacting body.

3. COMPARISON OF EXPERIMENTAL AND ANALYTICAL RESULTS


Ballistic tests have been performed by Empresa Nacional Santa Barbara. The tests
included firing 20 mm APDS projectiles (72 g tungsten projectile at 1250 m/s) against an
add-on armor of ceramic (alumina 99.5%) backed with Dyneema composite. Four config-
urations were tested; due to proprietary nature, these configurations are simply numbered
sequentially. The residual velocity of the projectile was measured by a Doppler system,
while the impact process was recorded in five different X-ray exposures, allowing the
determination of the residual length of the projectile. The reader can find the input data for
one of the configurations calculated in Table 1.
Figure 4 shows the comparison between analytical (crosses) and experimental (circles)
residual velocities. The uncertainty in the Doppler radar measurements was estimated in
$30 m/s. The analytical results show a good agreement compared to experimental ones.

Table 1. Example of input data used in the analytical model

Data description Symbol Value Units

Initial Mass of the projectile M 72.e!3 kg


0
Initial velocity 1250 m/s
0
Projectile diameter D 12e!3 m
Projectile density q 18,100 kg/m3
1
Yield strength of the projectile 3.2e9 Pa
1
Ceramic density q 3840 kg/m3
#
Ceramic sound speed c 10,000 m/s
Ceramic thickness h 20e!3 m
Intact ceramic yield strength 7.5e9 Pa
#
Conminuted ceramic yield strength @ 3.2e9 Pa
#
Characteristic length of the composite 0.25 m
Elastic modulus of the fabric yarns E 100e9 Pa
Cross section of the yarn S 0.215e!6 m2
Number of yarns per meter of fabric n 16.5e2 1/m
Number of layers n 130
-
Longitudinal speed of sound in the yarns c 10,000 m/s
:
Delamination energy per area unit G 0. J/m2
Integration step 1e!9 s
468 I. S. C. Benloulo and V. Sanchez-Galvez

Fig. 4. Comparison of the residual velocities ob- Fig. 5. Comparison of the residual lengths obtained
tained analytically and experimentally for a 20 mm analytically and experimentally for a 20 mm APDS
APDS projectile at 1250 m/s. Circles represent ex- projectile at 1250 m/s. Circles represent experimental
perimental data and crosses the result of the analyti- data and crosses the result of the analytical model.
cal model. Error bars represent the accuracy of the
Doppler radar.

All the four cases can be calculated in 5 s in a regular workstation or about 2 min in
a personal computer.
In Fig. 5, residual length of the projectile from the experiments and the model are
compared. Initial length of the projectile was 35.2 mm. The model provides always the best
performance attained by the ceramic element. Some of the ceramic targets may have some
defects that make them fail before expected, eroding the projectile less than desired and
giving the data dispersion observed.
As it is understood that nine tests, four different configurations, are not enough to
validate the analytical model, an additional series of eleven numerical tests, now changing
the projectile caliber, impact velocities and armor thickness, have been performed with
Autodyn-2D hydrocode, its results being compared with those of the analytical model.

5. NUMERICAL SIMULATION AND COMPARISON WITH


ANALYTICAL RESULTS
Autodyn 2-D is a commercial hydrocode specially designed to simulate impact problems
at high velocities. Material properties are entered directly through subroutines. The eleven
cases presented in Tables 2 and 3 needed around 100 h of calculation on a 486 personal
computer, once all the research and development work of the subroutines was achieved,
while the analytical calculation was performed in 10 m on the same computer.
Two very different projectiles were selected to validate the analytical model: 41.7 and
114.3 g. Table 2 summarizes the configurations used and the residual velocities and residual
lengths for impact of the smaller projectile onto different armors, alumina 99.5%/Dyneema,
at different impact velocities; the thickness of the armor elements was also varied. Table 3 is

Table 2. Configurations calculated and comparison of the results obtained with numerical and analytical methods
for a Tungsten projectile, 10 mm in diameter and 41.7 g

Case Ceram. Comp.


4 3 3 3 3
(m/s) thick. thick. analytic numeric analytic numeric
(mm) (mm) (m/s) (m/s) (mm) (mm)

1 1100 20 20 280 650 12 14


2 1250 20 20 720 780 13 15
3 1400 20 20 990 980 14 14
4 1250 25 20 300 430 9 12
5 1250 15 15 940 870 17 17
Analytical model for ceramic/composite armors 469

Table 3. Configurations calculated and comparison of the results obtained with numerical and analytical methods
for a Tungsten projectile, 14 mm in diameter and 114.3 g

Case Ceram. Comp.


4 3 3 3 3
(m/s) thick. thick. analytic numeric analytic numeric
(mm) (mm) (m/s) (m/s) (mm) (mm)

6 1000 20 20 520 660 23 26


7 1250 20 20 950 940 25 26
8 1500 20 20 1280 1220 26 25
9 1250 20 25 930 910 25 25
10 1250 25 20 850 890 21 23
11 1250 25 25 820 830 21 21

Fig. 6. Graphic comparison of the residual velocities Fig. 7. Graphic comparison of the residual lengths
obtained analytically and numerically for the config- obtained analytically and numerically for the config-
urations of Tables 2 and 3. urations of Tables 2 and 3.

Fig. 8. Residual Velocity versus Striking velocity for the 10 mm projectile against ceramic/dyneema
armor, calculated analytically and numerically. The thicknesses were 20 mm/20 mm.

the summary of the results for the larger projectile. The results are also shown graphically in
Figs 6 and 7 to assess better the correlation of the analytical and numerical models.
Although not experimentally determined it is estimated that cases 1, 4 and 6 are close to
the ballistic limit of the armor. The residual velocity versus impact velocity for the analytical
model is shown in Fig. 8 for the 41.7 g projectile and 20 mm thick ceramic, the same
thickness as that of the Dyneema composite. The numerical calculations are also shown in
this figure. It is observed that the results are very similar except where the ballistic limit is
approached. Near the ballistic limit, residual velocity versus striking velocity response is
very sensitive to initial conditions, e.g. a small change in the striking velocity results in
a large change in the residual velocity. The slope of the versus curve at the ballistic
3 4
limit is, in fact, theoretically infinite. This is also clearly observed in the experiments
reported by Anderson et al. [21]. Near the ballistic limit the details of the failure process
470 I. S. C. Benloulo and V. Sanchez-Galvez

Fig. 9. Velocity history of the rear of the projectile, calculated numerically and analytically.

become increasingly important. Hence, because the numerical and analytical models have
different failure models it is expected that near the ballistic limit the results will not match so
well. But apart from the cases close to the ballistic limit, the analytical predictions are quite
good considering the number of simplifications. This could permit an easy design of the
light-weight armor as hundreds of cases may be analyzed in a short period of time.
The numerical simulation also allows the analysis of the history of stresses, displace-
ments, velocity, etc. For instance, Fig. 9 plots the velocitytime history of the projectile for
the numerical and analytical models. It is seen that for the first 11 or 12 ls, while the ceramic
is not moving, the velocities match perfectly. This result confirms that the Tate model is very
good for impact of a rod onto a rigid surface. The second phase is when 11 ls(t(20 ls;
the projectile pushes the ceramic cone against the composite. Numerical and analytical
solutions are quite different here. In fact, this is expected because of the derivation of the
analytical composite model: The composite is assumed to fail abruptly at around 20 ls, and
no erosion process has been defined in order to keep the simplicity of the model. On the
other hand, the numerical code evolves more smoothly and naturally due to the (assumed)
erosive process during penetration. Perhaps the next step to take in the analytical model is
to develop a simple erosive model, like the failure of each ply in the composite.

5. CONCLUSIONS
A simple analytical model of high-velocity impact onto ceramic/composite has been
developed. The model is easily encoded, for instance, in FORTRAN, and a complete
impact/penetration problem takes a few seconds to run on a personal computer. The model
is divided into three phases of penetration: intact ceramic, fractured ceramic and initial
response of the composite substrate, and fabric response and failure. The assumptions of
each phase of the model were checked separately, giving reasonable engineering results.
Predictions from the combined model are in relatively good agreement with a limited set of
experimental data. The model was also compared with results from numerical simulations.
There was relatively good agreement on the residual velocity and residual length of the
projectile for most of the cases compared. However, near the ballistic limit velocitywhere
the details of the failure are importantdiscrepancies existed. This is to be expected since
failure is modeled completely differently in the analytical model and finite element ap-
proaches. The approach presented in this paper is a first step in developing an analytical
model of the impact against ceramic/composite armors.

AcknowledgementsAuthors are indebted to Philip Cunniff, Res. Engineer at U.S. Army Natick RD and E Center,
MA, for his help, dedication and useful conversations.
Authors would also like to thank, Dr Charles E. Anderson, Dr James D. Walker, from Southwest Research
Institute, and Prof. Carlos Navarro, from Universidad Carlos III de Madrid, Spain, for their suggestions and
Comision Interministerial de Ciencia y Tecnolog a, Empresa Nacional Santa Barbara and Ministerio de
Educacion y Ciencia for the economical support.
Analytical model for ceramic/composite armors 471

REFERENCES
1. A. L. Florence, Interaction of projectiles and composite armor, Internal Report, US Army, August, 1969.
2. J. F. Mackiewicz, Advanced ceramic/composite armour for the defeat of small arms. Proc. ightweight
Armours Systems Symp. 95, 1995, Paper 24.
3. A. Davey, Ceramic faced bristol armour. Proc. ightweight Armor Systems Symp., Shrivenham 95, UK, June
1995, Paper 23.
4. C. Navarro, M. A. Mart nez, R. Cortes and V. Sanchez-Galvez, Some observations on the normal impact on
ceramic faced armours backed by composite plates. Int. J. Impact Engng, 13(1), 145156 (1993).
5. R. L. Woodward, A simple one-dimensional approach to modeling ceramic composite armour defeat. Int. J.
Impact Engng 9(4), 455474 (1990).
6. Paulus Cornelis den Reijer, Impact on Ceramic Faced Armour. Ph.D. Thesis, Delft University, 1991
7. J. G. Hetherington, The optimization of two component composite armours. Int. J. Impact Engng, 12(3), 1992,
409414
8. R. Zaera and V. Sanchez-Galvez, Analytical model of ballistic impact on ceramic/metal lightweight armours.
Proc. Ballistics 96 Conf., vol. T13 San Francisco, 2328 September 1996, 487495 (1996).
9. Isaias Sidney Chocron Benloulo, J. Rodr guez and V. Sanchez Galvez, A simple analytical model to simulate
textile fabric ballistic impact behavior. extile Res. J. Tentative date: July (1997).
10. I. S. Chocron-Benloulo and V. Sanchez-Galvez, Impact resistance of polymeric matrix composites. Proc.
ICCE/3 Symp., July 1996.
11. A. Tate, A theory for the deceleration of long rods after impact. J. Mech. Phys. Solids, 14 387399 (1967).
12. Z. Rosenberg and E. Dekel, A critical examination of the modified Bernoulli equation using two-dimensional
simulations of long rod penetrators. Int. J. Impact Engng, 15 (5), 711720 (1994).
13. D. Roylance, A. Wilde and G. Tocci, Ballistic impact of textile structures. extile Res. J. 3441 (1973).
14. Philip M. Cunniff, An analysis of the system effects in woven fabrics under ballistic impact. extile Res. J.,
62(9), 495509 (1992).
15. N. Beaumont, Contribution a` lEtude de lImpact dune Bille sur une Plaque en Materiau Composite, Ph.D.
Thesis, 1991
16. C. Navarro, J. Rodr guez and R. Cortes, Analytical modeling of composite panels subjected to impact loading.
J. de Physique I, Colloque C8, supplement au J. de Physique III, 4, 515520 (1994).
17. J. C. Smith, F. L. McCrackin and H. F. Schiefer, Stressstrain relationships in yarns subjected to rapid impact
loading. Part V: wave propagation in long textile yarns impacted transversely. extile Res. J., 288302 (1958).
18. R. A. Prosser, Penetration of nylon ballistic panels by fragment-simulating projectiles. Part I: a linear
approximation to the relationship between the square of the V50 or Vc striking velocity and the number of
layers of cloth in the ballistic panel. Part II: mechanism of penetration. extile Res. J., 6185 (1988).
19. I.S. Chocron Benloulo, J. Rodr guez, M.A. Mart nez and V. Sanchez Galvez, Dynamic tensile testing of aramid
and polyethylene fiber composites. Int. J. Impact Engng 19(2), 135146 (1997).
20 E. H. M. van Gorp, L. L. H. van der Loo and J. L. J. van Dingenen, A model for HPPE-Based lightweight
add-on armour. Ballistics 93 (1993).
21. Charles E. Anderson Jr., Volker Hohler, James D. Walker and Alois J. Stilp, Time-resolved penetration of long
rods into steel targets. Int. J. Impact Engng, 16(1), (1995) 118 (1995).

Das könnte Ihnen auch gefallen