Вы находитесь на странице: 1из 19

International Journal of Impact Engineering 96 (2016) 146164

Contents lists available at ScienceDirect

International Journal of Impact Engineering


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / i j i m p e n g

An investigation of the constitutive behavior of Armox 500T steel and


armor piercing incendiary projectile material
M.A. Iqbal a,*, K. Senthil a, P. Sharma b, N.K. Gupta c
a
b
c

Department of Civil Engineering, Indian Institute of Technology Roorkee, Roorkee 247667, India
Terminal Ballistics Research Laboratory, Chandigarh 160003, India
Department of Applied Mechanics, Indian Institute of Technology Delhi, New Delhi 110016, India

A R T I C L E

I N F O

Article history:
Received 30 January 2016
Received in revised form 17 May 2016
Accepted 22 May 2016
Available online 26 May 2016
Keywords:
Armox 500T steel
Armor piercing incendiary projectile
Material characterization
Calibration of JohnsonCook model
Ballistic limit

A B S T R A C T

A detailed investigation has been carried out for studying the constitutive behavior of Armox 500T steel
and armor piercing incendiary projectile (API) material under varying stress state, strain rate and temperature. The characterization of Armox 500T steel showed increase in its strength with increase in stress
triaxiality as well as strain rate. Increment in temperature, on the other hand, induced signicant increase in the material ductility while reducing its strength. The API projectile material remained insensitive
to stress-triaxiality and strain rate; however, it was highly sensitive to thermal effects. Results thus obtained from experiments on the specimens of both the materials were subsequently employed for calibrating
the material parameters of JohnsonCook (JC) ow and fracture model. The calibrated JC model for Armox
500T steel has been validated by numerically simulating the high strain rate tension tests performed on
split Hopkinson pressure bar apparatus.
The ballistic experiments were carried out wherein 8 and 10 mm thick Armox 500T steel target plates
were impacted by 7.62 and 12.7 API projectiles respectively at a normal incidence with a velocity of nearly
830 m/s. The results of the ballistic tests were reproduced through nite element simulations performed on ABAQUS/Explicit nite element code employing calibrated JC model for the target as well as
the projectile material. Experimental and numerical ndings with respect to failure mechanism and ballistic resistance of the target are presented and discussed. It is seen that the computed failure modes
and residual velocities accurately matched the experiments.
Further, the ballistic limit of the target material was obtained numerically and the values obtained
were validated through the RechtIpson empirical model.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
The mechanics of perforation of metal targets due to a projectile impact is associated with viscous and thermal effects, which
signicantly inuence the material behavior of the interacting
bodies. The simulation of perforation process therefore requires a
detailed description of the projectile and the target materials through
a constitutive model capable of incorporating the inuence of
large strain, high strain rate and high temperature besides being
able to model crack propagation, degradation and fracture of the
material [1].
The multiaxial state of stress in a constitutive model is generally expressed in terms of equivalent von-Mises yield stress as a
function of accumulated plastic strain, plastic strain rate and ab-

* Corresponding author. Department of Civil Engineering, Indian Institute of


Technology Roorkee, Roorkee 247667, India. Tel.: +91 9897379426, +91 1332 285866;
Fax: +91 1332 275568.
E-mail addresses: iqbal_ashraf@rediffmail.com; iqbalfce@iitr.ac.in (M.A. Iqbal).
http://dx.doi.org/10.1016/j.ijimpeng.2016.05.017
0734-743X/ 2016 Elsevier Ltd. All rights reserved.

solute temperature [27]. Failure in a ductile material is initiated


by nucleation and coalescence of micro voids originating from the
point of impact and disseminating through the material due to localized plastic ow. Rapid nucleation, coalescence and growth of
voids reduce ductility of the material and hence its fracture strain.
Available studies in literature [2,811] reveal that ductility of a material strongly depends on the stress state, which in turn is related
to stress triaxiality, such that the fracture strain decreases with increase in the stress triaxiality. Material ductility is seen to moderately
decrease with increase in strain rate, while it is signicantly enhanced with increase in temperature beyond a limit [1215]. A
number of available constitutive models describe the failure in metals
in terms of equivalent fracture strain, dened as a function of stress
triaxiality [2,16,17], strain rate and temperature [2,1823]. These
models based on macro mechanics approach vary in their degree
of complexity and robustness in capturing ow and fracture behavior in metals [27,16,19,21].
Wierzbicki et al. [24] carried out a comparative investigation of
the fracture mechanism of 2024-T351 aluminum coupons using
seven different conventional fracture models [XueWierzbicki (XW),

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

Wilkins, JohnsonCook (JC), crash FEM, maximum shear stress,


constant equivalent strain criteria, fracture forming limit diagram].
Specimens of various geometries facilitated the failure of material
through distinct modes of deformation and hence enabled the calibration of different material models and correlation of their
predictions. The maximum shear stress criterion fairly reproduced the failure preceded by shear localization but could not predict
fracture under axi-symmetric loading. The XW model, constructed
under three-dimensional space of invariants, adequately dealt with
all the stress states and predicted the failure mechanisms fairly well.
The Wilkins criterion could not reproduce good results under a combination of large and small range of stress triaxiality due to separable
effect of hydrostatic and deviatoric stress states on the material ductility. The crash FEM model was able to distinguish the ductile and
shear fracture. The JohnsonCook model, FFLD approach and constant equivalent strain reproduced the experimental results well
within a narrow range of stress triaxiality.
Teng and Wierzbicki [25] further employed six distinct fracture models to predict the damage in Weldox 460 E steel and 2024T351 aluminum plates impacted by blunt nosed projectiles.
Assuming the projectile to be rigid, the ow behavior of both the
target materials was simulated employing JohnsonCook plasticity model. The actual and predicted failure mechanism and residual
projectile velocities were compared and the limitations of each fracture model were discussed. The Wilkins fracture model predicted
unrealistic spallation of both the materials due to absence of ductility effects. The maximum shear stress criterion could not predict
the shear plugging under a range of incidence velocities. The resultant critical shear stress caused either premature or incomplete
target failure. The single parameter based modied Cockcroft
Latham model could not reproduce damage as well as residual
velocities in low as well as high ductility materials. The conventional critical strain and the simplied BaoWierzbicki (BW) fracture
models also failed to give satisfactory results for a wide range of
problems. On the other hand, JC failure model reproduced realistic predictions of fracture behavior and residual projectile velocities.
Banerjee [26] studied the limitations of ve different strain rate
and temperature dependent ow stress models by simulating the
behavior of OFHC copper under one dimensional tension and compression tests performed at varying strain rate and temperature.
Stressstrain relations, strain rate and temperature dependence of
yield stress, fracture strain and hardening and softening response
were studied and compared for JC, SteinbergCochranGuinan
Lund (SCGL), ZerilliArmstrong (ZA), mechanical threshold (MTS)
and PrestonTonksWallace (PTW) models. The JC model overestimated the initial yield for the quasi-static test performed at room
temperature and underestimated the rate of hardening for the test
performed at room temperature at 8000 s1 strain rate. The SCGL
model underestimated the softening at high temperature but accurately predicted the yield stress at 8000 s1 strain rate. At room
temperature, the ZA model accurately predicted the quasi-static yield
stress but underestimated the yield stress at 8000 s1 strain rate. The
yield stress at 4000 s1 strain rate was reasonably predicted; however,
the decrement of yield stress with increasing temperature was overestimated. The MTS model underestimated the yield stress at 296 K
temperature and 8000 s1 strain rate while it accurately predicted
the same at 1023 K and 1800 s1. At 1173 K temperature, the model
overestimated the yield stress for the quasi static test and accurately predicted its value at high strain rate. The PTW model explicitly
accounted for the rapid increase in yield stress at strain rate above
1000 s1 and performed better under the tests carried out in compression than in tension. For the Taylor impact simulations, however,
the JC model has reproduced the mushroomed diameter and the
nal length most accurately. All the other models though correctly predicted the nal length but underestimated the
mushroomed diameter.

147

Sjoberg et al. [27] calibrated JC and ZA constitutive models for


alloy 718 (composed of nickel, iron and chromium) and simulated
the instrumented reverse impact experiment to measure force time
histories, which eventually enabled the validation of the model. The
force time history relationship reproduced using the Johnson
Cook model was found to have closer agreement with experiments
in comparison to that from the ZA model.
The selection of a constitutive model for simulating a nite element
problem is primarily governed by its accuracy, robustness, simplistic formulation and physical interpretation of ow and fracture
algorithms. The compatibility of the model with the available nite
element codes, number of material parameters, their availability and
calibration procedure are some practical and important considerations. A phenomenological model capable of representing effective
stress and damage concepts in a simplied manner, employing
minimum number of material parameters, is considered more realistic and practical. The explicit and easy calibration procedure, fewer
number of parameters, availability in most of the nite element codes
and capability of simulating local and global structural behavior are
the reasons for widespread utility and acceptability of the Johnson
Cook model for predicting perforation in metals [12,15,2831].
Although the formulation of the JC model has been somewhat criticized due to its purely empirical approach and uncoupled physical
interpretations [5,13,26,32,33], its uncoupled approach has enabled
exibility of including or omitting the complexity in modeling the
material behavior, which, has proven to be of practical signicance.
In the available studies, the penetration and perforation of metallic
plates have been effectively simulated numerically by employing the
JC model for predicting the ow and the fracture characteristics of
the target and the projectile [14,15,2931,34].
Borvik et al. [12,13] introduced damage in the JC ow stress model
and modied the strain rate sensitivity expression [12] for predicting the plugging phenomenon in 12 mm thick Weldox 460 E steel
plates impacted by blunt nosed projectiles. The model was implemented in explicit nite element code LS-DYNA and also coupled
with an element kill algorithm to allow crack growth during penetration. The simulations of plugging failure of metallic plates
matched the experimental ndings reasonably well. The model was
found capable of describing the main target response and physical material behavior and reproduced ballistic limit within 10%
accuracy. In order to study the inuence of various physical characteristics employed in the model, simulations of plugging failure
were further carried out by including and omitting various model
parameters [12]. The size of element was found to be a vital parameter such that none of the models could predict the ballistic limit
even within 30% accuracy when coarser mesh was used. All the
models showed good agreement with experimental ndings at incidence velocities well above the ballistic limit. However, close to
ballistic limit the difference between the results of various models
became more prominent. The strain rate, temperature and stress
states were all found important parameters, as their omitting introduced signicant errors in the results.
Iqbal et al. [15] performed nite element simulations for studying the ballistic performance of 12 and 16 mm thick mild steel targets
against 7.62 AP projectile at varying obliquity until the occurrence
of projectile ricochet. The target material was characterized and calibrated for obtaining all the material parameters of JC ow stress
and fracture strain model. The calibrated material parameters were
also validated by numerically simulating the high strain rate tension
tests. Experimental and numerical true stressstrain relations were
found to match closely. The target failure mechanism, residual projectile velocity and critical angle of ricochet were also closely
predicted by the model.
The above studies show that the JC model is capable of accurately capturing the ow and fracture material characteristics in
perforation problems involving varying stress-state, strain rate and

148

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

temperature. The model has been calibrated for various materials


including structural and armor steels and aluminum alloys
[1215,30,31]. However, the studies pertaining to the characterization and ballistic evaluation of Armox 500T steel could not be
found in the open literature despite the primary armor application
of the alloy. Further, as far as the small arms projectiles are concerned, the armor piercing projectile is a major threat to body and
vehicle armor and it is signicantly studied in the open literature.
However, the material model of the projectile, which is complete
and accurate, is still not available in the literature [30,31,35,36].
The present work is based on a detailed investigation of the material behavior of Armox 500T steel and armor piercing incendiary
projectiles (7.62 API and 12.7 API) under varying stress state, strain
rate and temperature. The armor as well as projectile materials were
calibrated for obtaining the original JC model parameters through
curve tting method. The calibrated JC model for Armox 500T steel
was also validated by numerically simulating the high strain rate
material tests. The material parameters were subsequently employed for numerically simulating the ballistic trials conducted on
8 and 10 mm thick Armox 500T steel plate targets impacted respectively by 7.62 and 12.7 API projectiles at normal incidence velocity
close to 830 m/s. The experimental and computed ballistic results
obtained with respect to target failure mechanism and residual projectile velocities are presented, compared and discussed. The ballistic
limit of both the targets was also evaluated numerically and the results
were validated through RechtIpson empirical model.
2. Material characterization of Armox 500T steel
The Armox 500T steel is considered suitable for civil and military ballistic applications due to its high strength and high hardness.
This steel was procured in the form of plates of different thicknesses and the coupons for material tests were extracted from the
middle of a 20 mm thick plate. The average hardness of the material was found as 501 BHN on a 10 mm 10 mm 10 mm cubical
specimen with respect to in plane as well as out of plane surface,
with 10% variation in different observations. Chemical composition of the material was studied by energy-dispersive X-ray (EDX)

Table 1
Chemical composition of Armox 500T steel.
Chemical composition

EDX spectroscopy

As per the certicate of


SSAB

C
Si
Mn
P
S
Cr
Ni
Mo
Al
Cu

0.3484
0.2732
0.7672
0.0208
0.0066
0.5042
0.7498

0.0243
0.1429

0.28
0.26
0.91
0.008
0.001
0.5
0.94
0.349
0.053

spectroscopy and the percentages of the ingredients thus determined are given in Table 1.
Smooth cylindrical tension test specimens of 6.25 mm diameter and 32 mm gauge length (see Fig. 1 (a)) were prepared in
accordance with ASTM 370-12. To study any possible anisotropy of
the material, these specimens were extracted in 0, 45 and 90 directions, and quasi-static uniaxial tension tests were performed on
Tinius Olsen H75KS and Milano Controls UTM (universal testing
machine) at a constant strain rate, 1.6 104 s1. The engineering
stressstrain curves plotted in Fig. 1(b) do not show signicant variation, and the material was assumed to be isotropic. In order to
evaluate the true stressstrain relation, the diameter reduction of
the smooth cylindrical specimen was continuously measured up to
fracture with the help of a digital vernier caliper. The true stress
was obtained as the force per unit actual cross-sectional area, F/Ac,
and the true strain, ln (Ai /Ac) where Ai and Ac are the initial and
current cross-sectional areas of the specimen respectively. The
Bridgman [37] correction was applied to the measured true stress
and the corrected true stressstrain relationship thus obtained corresponding to different material orientations is shown in Fig. 2(a).
The corresponding JC hardening curves are presented in Fig. 2(b)
for comparison. The corresponding JC strength parameters A, B and
n obtained for the three material orientations, presented in Table 2,
show insignicant variation.

(a)

1800
Engineering stress (MPa)

(b)

1500
1200
900
600

0
45
90

300
0
0

0.03
0.06
0.09
Engineering strain

0.12

Fig. 1. (a) Schematic of smooth cylindrical specimen and (b) experimental engineering stressstain curves of Armox 500T steel under tension.

2000

2000

1800

1800

1600

True stress (MPa)

True stress (MPa)

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

0 - Experimetns
45 - Experiments
90 - Experiments

1400
1200

149

1600
0 - Johnson-Cook model
1400

45 - Johnson-Cook model
90 - Johnson-Cook model

1200
(a)

(b)

1000

1000
0

0.2

0.4
0.6
True strain

0.8

0.2

0.4
0.6
True strain

0.8

Fig. 2. True stressstrain curves of Armox 500T steel; (a) experimental and (b) tted JC model.

Table 2
Material parameters corresponding to different orientations of Armox 500T steel plates.
Orientation

0
45
90

1372.488
1423.264
1319.869

535.022
499.414
504.696

0.2467
0.2455
0.2892

2.1. Effect of stress triaxiality


The stress triaxiality has been found to have signicant inuence on material ductility [8,38,39] such that there is increase in
material ductility with decrease in stress triaxiality. The initial stress
triaxiality was introduced by producing an articial notch in the cylindrical specimen (Fig. 3 (a)) and measured at the center of the notch
assuming it to be constant. According to Bridgmans analysis [37]
the initial stress triaxiality ratio could be calculated in terms of
maximum stress triaxiality ratio, max
* ;

*max =

1
a

+ ln 1 +

3
2R

(1)

where R is curvature radius of the notch and a is the radius of


the specimen at the middle of the notched region. In the present
study, the initial stress triaxiality ratio has been varied as 1.95, 1.43,
0.927, 0.765 and 0.67 by varying the curvature radius of the notched
specimen as 0.4, 0.8, 2.0, 3.0 and 4.0 mm respectively (see Fig. 3(b)
for the undeformed and Fig. 3(c) for the deformed specimens). The
specimens of different notched curvature radii extracted from 0 orientation were tested under tension on 70-S11U02 Milano Controls
UTM at a constant strain-rate 1.6 104 s1 and the engineering stress
strain relations thus obtained are plotted in Fig. 4. The stress
strain relations were recorded with the help of integrated data
acquisition system and inbuilt load cell based piston of the UTM.
With increase in stress triaxiality, the fracture strain and therefore
the material ductility have been found to decrease. The stress
strain curves of the notched specimens have been found to have
serrations due to the nonlinear specimen geometry. In case of the
smooth cylindrical specimen on the other hand, the serrations have
been found to be insignicant (Fig. 4). Clausen et al. [14] also observed serrations more distinguished in notched specimens than
in smooth cylindrical specimens of 5083 H116 aluminum alloy under
quasi-static tension tests.
2.2. Effect of strain rate
The low and intermediate strain rate uniaxial tests were carried
out on smooth cylindrical specimens of diameter 6.25 mm and gauge

length 32 mm (Fig. 1(a)). These tests were performed on Tinius Olsen


H75KS universal testing machine, at strain rate 1.6 1043.3 102 s1.
The inbuilt load cell of the UTM enabled the application of load. The
tension tests at high strain rate (5001000 s1) were performed on
smooth cylindrical specimens of diameter 3 mm and gauge length
13.4 mm (Fig. 5(a)) on split Hopkinson pressure bar apparatus. The
fractured specimens at low, intermediate and high strain rate are
shown in Fig. 5(b) to (d) respectively. The engineering stress
strain curves obtained at low, intermediate and high strain rates
through respective tests are shown in Fig. 6(a). Typical curves for
repetitions performed at strain rate 0.01 s1 and 850 s1 are shown
in Fig. 6(b) and (c) respectively. The reason behind the serrations
obtained in the medium strain rate test (Fig. 6(b)) might be the very
high operation speed of UTM, 90 mm/min. The tests carried out in
the range 1.6 104 s13.3 102 s1 showed insignicant inuence of strain rate on the material behavior with respect to its
hardening, stressstrain relation or fracture behavior. At high strain
rate, however, the strength of the material was seen to have increased signicantly (Fig. 6(a)). Generally, the material tests at strain
rate 10+210+5 s1 are performed on split Hopkinson pressure bar apparatus based on the principle of one dimensional elastic wave
propagation theory [40]. The accuracy of the tension tests performed on split Hopkinson bar is based on the force equilibrium
at the ends of the specimen, which requires the specimen to be
loaded uniformly throughout its length. Proper xity of the specimen is crucial for maintaining the force equilibrium during tension
test. In the present study, therefore, the compatibility between the
threads of the specimen and that of the incident and transmission
bars was carefully maintained to ensure proper xity at both ends.
Moreover, the specimen, being considerably smaller (gauge length
13.4 mm) in comparison to that of the incident (2.0 m) and transmission (2.0 m) bars, has been assumed to have reached the force
equilibrium after several reections of elasto-plastic waves inside
it. The pulse was generated by hitting the 0.5 m long and 20 mm
diameter maraging steel striker to the incident bar. The strain gauges
were xed to the incident as well as transmission bars to record
the incident and transmitted wave respectively with the help of an
oscilloscope. The duration of the pulse was 5 105 s. The nominal
difference in the repetitions of the tension tests may be due to
mechanical/data acquisition error.

2.3. Effect of temperature


The tension tests at varying temperature were carried out on
smooth cylindrical specimens of gauge diameter 6.25 and gauge
length 32 mm (Fig. 1(a)). These tests were performed at 100, 200,
300, 400, 500, 600, 750 and 900 C on servo-hydraulic UTM at the

150

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

(a)

R0.4

R0.8
R2.0

R3.0

(b)

R4.0

R0.4
R0.8

R2.0

(c)
R3.0
R4.0
8

Fig. 3. Notched specimens of Armox 500T steel with different initial notch radii; (a) schematic, (b) un-deformed and (c) deformed.

Engineering stress (MPa)

3000

R0.4
R0.8
R2.0
R3.0
R4.0

2500
2000
1500
1000
500
0
0

0.02

0.04 0.06 0.08


Engineering strain

0.1

Fig. 4. The stressstrain curves of Armox 500T steel obtained under tension for different initial notch radii.

strain rate 1.6 104 s1. A portable furnace of internal diameter


60 mm and height 220 mm was used to vary the temperature of the
specimen. The furnace was rmly held between the platens of the
UTM with the help of steel xture and the specimen was placed
inside the furnace. The temperature of the furnace was controlled
by the high temperature furnace microcontroller. Two thermocouples were inserted in the furnace and held close to the specimen
surface to measure its temperature. As the inside diameter of the
furnace was very small (60 mm), the specimen temperature and the
furnace temperature have been assumed to remain identical. The
rate of temperature increase was kept as 10 C per minute. The required temperature was kept constant for about 30 minutes before
the test, in order to ensure proper heating of the specimen. The longitudinal strain of the specimen was measured with the help of two
LVDTs (linear variable displacement transducer). The ow stress of
the material was found to be signicantly affected due to the variation in the temperature. Initially, the ow stress increased slightly
when the temperature increased up to 200 C (see Fig. 7(a)). This
phenomenon is known as the blue brittle effect, whereby the
strength of steel increases and the ductility decreases on heating.
At 300 C temperature, however, the ultimate stress of the material was seen to reduce while the fracture strain increased. At 600 C

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

(a)

Schematic of the specimen for high strain rate tests

(b) Tested specimens at low strain rate

(c) Tested specimens at intermediate strain

(d) Tested specimens at high strain rate

Fig. 5. Specimens of Armox 500T steel; (a) schematic for high strain rate tension
test and (b), (c) and (d) tested in tension at low intermediate and high strain rates
respectively.

temperature, the ultimate stress reduced abruptly. At 900 C temperature, the material lost all of its strength and the specimen
underwent signicant elongation before breakage. The typical repetitions performed at 400 C temperature are shown in Fig. 7(b). The
test results showed that the yield stress of the material decreased
with increase in temperature while the ow stress exhibited a local
maximum at a temperature of about 200 C. The fracture mode of
the specimens at varying temperature is also shown in Fig. 8. Above
400 C temperature, the diameter of the fractured specimen reduced
and thermal softening became more visible. At 900 C, the softening of material caused purely ductile failure.
3. Material characterization of API projectiles
The armor piercing incendiary projectiles used in the present study
are thermally active projectiles, possessing incendiary material between
the steel core and the jacket. The incendiary gun powder may ignite

151

after coming in contact with the target and induce re in case the
target possesses inammable characteristics. The 7.62 API projectile
(Fig. 9(a)), a common threat to human beings, is generally red through
a standard rie or bipod mounted light and heavy machine guns. The
12.7 API projectile (Fig. 9(b)) is considered a major threat to armored
vehicle, sentry bunker and aircraft, and is usually red by carriage
mounted air defense gun which enables fairly high rate of ring.
The average Vickers hardness (HV) of 7.62 and 12.7 API projectiles measured at their various locations (Fig. 10(a) and (b)
respectively) was found to be 831 and 812 respectively. The ultimate tensile strength of both of these projectiles could be roughly
calculated to be three times the Vickers hardness number [30,31,41],
and is estimated to be approximately 2300 MPa. The chemical composition of 7.62 and 12.7 API cores was studied at three different
locations through the length of the core with the help of EDX spectroscopy, and the constituents are reported in Table 3.
As the size of the projectiles was very small (shank length 20.7
and 24.4 mm respectively for 7.62 and 12.7 API projectiles), it was
dicult to machine them to standard specimen. The machining of
the projectiles was also restricted by the ogival curvature of the nose.
These limitations led to the preparation of miniature specimens of
gauge diameter 2.5 mm and gauge length 10 mm from the shank
of the hardened steel core (see Fig. 11(a) and (b) respectively for
the specimens prepared out of the core of 7.62 and 12.7 API projectiles). These specimens were held through a specially designed
dog bone arrangement (Fig. 11(c) and (d)) and tested under tension
on Tinius Olsen H75KS machine at strain rate 1 104 s1 and corresponding cross head speed 0.1 mm/m (see Fig. 11(e) and (f)
respectively for deformed specimens). The engineering stress
strain curves obtained on the core material of 7.62 and 12.7 API
projectiles are shown in Fig. 12. The specimens underwent insignicant plastic deformation and experienced sudden failure. The
static yield strength of the material was obtained as 0.2% proof stress.
There was a close agreement between the material behavior of
7.62 and 12.7 API projectiles with respect to hardness, chemical composition and stressstrain relation, and therefore further material
tests were carried out only on the 12.7 API projectiles.
The initial stress triaxiality was introduced by machining an articial notch in the smooth cylindrical specimens of the API projectile.
The value of the initial stress triaxiality was varied by varying the
curvature radius of the notch as 0.4, 0.8 and 2.0 mm, keeping a constant gauge diameter of 2.5 mm. These notched specimens were
tested on Tinius Olsen H75KS UTM machine at a constant strainrate 1 104 s1 and corresponding cross head speed 0.1 mm/min.
The un-deformed and deformed specimens are shown in Fig. 13 (a)
and (b) respectively. The engineering stressstrain curves obtained under uniaxial tension at different initial stress triaxialities
(notch radii) are shown in Fig. 14. The fracture strain of the material corresponding to different stress triaxialities has been found to
be nearly the same (0.0110.0125), and its inuence on material
ductility is quite insignicant.
The high strain rate tests in the range 27995333 s1 were performed under compression on cylindrical tablet specimens of 5 mm
diameter and 5 mm height using split Hopkinson pressure bar apparatus. The deformed tablet specimens are shown in Fig. 15. The
strain in the specimen material was measured with the help of strain
gauges xed to the incident and transmission bars and the data were
recorded with the help of an oscilloscope. The engineering stress
was obtained as force per unit initial cross-sectional area.
The true stressstrain curve obtained under uniaxial tension test
performed on cylindrical specimens at quasi-static strain rate,
6 104 s1, is shown in Fig. 16(a). The cylindrical specimens of projectiles did not depict strain localization and neck formation (see
Fig. 11(f)). The actual cross-sectional area of these specimens could
not be measured during the tests as the specimens underwent
sudden failure. Therefore, the true stress, t, and true strain, t, have

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

0.000167/s
0.001667/s
0.010/s
0.033/s
850/s
950/s

2400
2000
1600
1200
800
400

1800
Engineering stress (MPa)

Engineering stress (MPa)

152

(a)

Strain Rate 0.01 1/s

1600
1400
1200
1000
800

Trial 1
Trial 2
Trial 3

600
400
200

(b)

0
0

0.02

0.04 0.06 0.08


Engineering strain

0.1

0.12

0.03
0.06
0.09
Engineering strain

0.12

Engineering stress (MPa)

3000
Strain Rate 850 1/s
2500
2000
1500
Trial 1

1000

Trial 2

500

(c)

Trial 3

0
0

0.02
0.04
Engineering strain

0.06

Fig. 6. Stressstrain curves of Armox 500T steel under tension (a) at varying strain rate, (b) repetitions performed at 0.01 s1 and (c) repetitions performed at 850 s1
strain rate.

2000

20 C
100 C
200 C
300 C
400 C
500 C
600 C
750 C
900 C

Engineering stress (MPa)

(a)
1600
1200
800
400

due to high strength projectile material, the tablet specimens were


prepared of comparatively smaller diameter (5 mm). The reduced
diameter of specimen enabled the development of the required strain
rate at comparatively lesser incidence velocity and thus helped to
avoid the damage to the attached strain gauges.
The specimens (of API projectiles) for the temperature tests had
gauge diameter 3 mm and gauge length 10 mm (see Fig. 17(a)) and
were tested in tension at 27, 400, 520, 600, 700, and 950 C on servohydraulic UTM at the strain rate 6 104 s1. A portable furnace was
employed for heating the specimen to the required temperature controlled by a micro-temperature controller. The specimen was heated
to the desired temperature at the rate of 10 C per minute, and maintained for half an hour at the given temperature in order to get it
heated uniformly before the application of mechanical load. The
inside temperature of the furnace was recorded using the K-type

2000
Engineering stress (MPa)

been calculated using the expressions t = (1 + ) and t = ln (1 + ),


respectively, where and represent the engineering stress and
strain respectively. The true stressstrain curves under uniaxial compression tests performed on tablet specimens at strain rate 2799
5333 s1 (Fig. 16 (b)) have also been calculated using the same
expression. The results of compression tests exhibited serrations possibly due to very high strain rate and high strength of the projectile
material. The results of the quasi-static uniaxial tension tests have
also been added in Fig. 16(b) for comparison. As such, no distinguished inuence of the strain rate has been observed on the
material behavior of the projectile. The Hopkinson pressure bar apparatus employed in the present study for performing high strain
rate tests had the strength of incident and transmission bars approximately 2400 MPa and diameter 20 mm. Therefore, in order to
avoid any permanent damage to the incident and transmission bars

(b)

1750
1500
1250
1000
750

Trial 1

500

Trial 2

250

Trial 3

0
0

0.1

0.2
0.3
0.4
Engineering strain

0.5

0.6

0.05
0.1
0.15
Engineering strain

0.2

Fig. 7. Stressstrain curves of Armox 500T steel under tension (a) at varying temperature and (b) repetitions at 400 C temperature.

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

153

Table 3
Chemical composition of 7.62 and 12.7 API projectiles.

100 C

200 C

300 C

400 C

500 C

Elements

7.62 API projectile (%)

12.7 API projectile (%)

Fe
C
Mn
Si
Cu
Cr
K
S
Ni
P
Al

84.06
12.27
0.63
0.69
0.49
0.42
0.30
0.13
0.72
0.05
0.26

84.49
11.74
0.65
0.67
0.41
0.38
0.23
0.27
0.28
0.35
0.54

1280 MPa; however, it had insignicant inuence on the fracture


strain. With increase in temperature to 600 C, the strength further
decreased to 800 MPa without signicant increase in fracture strain.
When the temperature reached 950 C, the ultimate strength reduced
to 600 MPa and the fracture strain increased to 19%.

600 C

4. Constitutive modeling

900 C

750 C

Fig. 8. Fractured surfaces of the specimens of Armox 500T steel at varying


temperature.

thermocouples held near the center of the specimen. Elongation of


the specimen was measured by two LVDTs (linear variable displacement transducers). The fractured surface of the specimen at
varying temperatures is shown in Fig. 17(b).
The stressstrain curves of the projectile core have been plotted
in Fig. 18, corresponding to varying temperature. The material behavior is seen to be signicantly affected due to temperature
variation. An initial increase in temperature from ambient to 400 C
resulted in the decrease of ultimate stress from 2250 to 1720 MPa
and a signicant increase in fracture strain: almost 12%. A further
increase in temperature to 520 C decreased the ultimate stress to

The ow and fracture behavior of projectile and target material was predicted employing the JohnsonCook elasto-viscoplastic
material model [2,19] available in ABAQUS [42] nite element code.
The material model is based on the von Mises yield criterion and
associated ow rule. It includes the effect of linear thermo-elasticity,
yielding, plastic ow, isotropic strain hardening, strain rate hardening, softening due to adiabatic heating and damage. The equivalent
von Mises stress of the JohnsonCook model is dened as;

 pl
n
pl,  pl, T = A + B ( pl ) 1 + C ln 1 T m
 0

where A, B, n, C and m are material parameters determined from


different mechanical tests. pl is equivalent plastic strain,  pl is equivalent plastic strain rate,  0 is a reference strain rate and T is nondimensional temperature dened as;

Fig. 9. Threats of (a) 7.62 mm and (b) 12.7 mm armor piercing incendiary (API) projectile.

(a)

(2)

(b)

Fig. 10. Smoothened surfaces across the length of (a) 7.62 and (b) 12.7 API projectile.

154

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

7.62 API Projectile

12.7 API Projectile

(a)

(b)

(c)

(d)

(e)

(f )

Fig. 11. (a) Machined specimen of 7.62 API projectile, (b) machined specimen of 12.7 API projectile, (c) and (d) dog-bone holding xtures for the projectile specimens, (e)
fractured specimen of 7.62 API projectile, (f) fractured specimen of 12.7 API projectile.

T = ( T T0 ) ( Tmelt T0 ) T0 T Tmelt

(3)

where T is the current temperature, Tmelt is the melting point temperature and T0 is the room temperature.

Engineering stress (MPa)

2400

Johnson and Cook [2] extended the failure criterion proposed by


Hancock and Mackenzie [8] by incorporating the effect of strain path,
strain rate and temperature in the fracture strain expression, in addition to stress triaxiality. The fracture criterion is based on the
damage evolution wherein the damage of the material is assumed
to occur when the damage parameter, , exceeds unity:

pl
= pl ,
f

2000
1600
1200
800
12.7 API Projectile

400

7.62 API Projectile


0
0

0.004

0.008
0.012
Engineering strain

0.016

Fig. 12. Stressstrain curves of 7.62 and 12.7 API projectiles under tension.

(a) Undeformed specimen

(4)

where pl is an increment of the equivalent plastic strain, fpl is


the strain at failure, and the summation is performed over all the
increments throughout the analysis. The strain at failure fpl is
assumed to be dependent on a non-dimensional plastic strain rate,
 pl
m
; a dimensionless pressure-deviatoric stress ratio,
(where
 0

m is the mean stress and is the equivalent von Mises stress)


and the non-dimensional temperature, T , dened earlier in the
JohnsonCook hardening model. The dependencies are assumed to
be separable and are of the form:

 pl



fpl m ,  pl, T = D1 + D2 exp D3 m 1 + D4 ln 1 + D5T (5)

 0

(b) Deformed specimen

R0.4

R0.8

R2.0

Smooth specimen

Fig. 13. (a) Un-deformed and (b) deformed tension test specimens of API projectiles of different notch radii.

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

2500

2400
Engineering stress (MPa)

Engineering stress (MPa)

2800

155

2000
1600
R0.4 mm
R0.8 mm
R2.0 mm

1200
800
400
0

27 C
400 C
520 C
600 C
700 C
950 C

2000

1500

1000

500

0.004
0.008
0.012
Engineering strain

0.016

Fig. 14. Stressstrain curves of API projectile material under tension as a function
of stress triaxiality.

Fig. 15. Deformed specimens of API projectile material tested under compression
at high strain rate.

0.05

0.1
Engineering strain

0.15

0.2

Fig. 18. Stressstrain curves of API projectile material under tension at varying
temperature.

where D1 D5 are material parameters determined from different


mechanical test,  pl is equivalent plastic strain rate and  0 is a reference strain rate.
When material damage occurs, the stressstrain relationship no
longer accurately represents the material behavior, ABAQUS [42].
The use of stressstrain relationship beyond ultimate stress introduces a strong mesh dependency based on strain localization, i.e.,
the energy dissipated decreases with a decrease in element size.
Hillerborgs [43] fracture energy criterion has been employed in the
present study to reduce mesh dependency by considering stressdisplacement response after the initiation of damage.

2400

3000
0.0006 1/s
2500
True stress (MPa)

True stress (MPa)

2000
1600
1200
800

2000
1500
2799 1/s
3733 1/s
5333 1/s
0.0006 1/s

1000

400

500
(a)

0
0

0.003

0.006

0.009

0.012

(b)

0.015

True strain

0.05

0.1
0.15
True strain

0.2

0.25

Fig. 16. Stressstrain curves of API projectile material; (a) quasi static tests performed under tension and (b) comparison of the high strain rate compression tests and quasistatic tension tests.

(a)

(b)

Fig. 17. API projectile specimens for high temperature tension tests; (a) un-deformed and (b) deformed.

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

2100

1.2

1800

1500

Fracture strain

True stress (MPa)

156

1200
900
600
Experiments

300
0

0.8
0.6
0.4
0.2

(a)

Johnson-Cook

Experiments
Johnson-Cook

(b)
0

0.2

0.4

0.6
0.8
True strain

1.2

0.3

0.6 0.9 1.2 1.5


Stress triaxiality

1.8

2.1

1.2

1600

Fracture strain

Engineering stress (MPa)

1.4
2000

1200
800

Experiments

(d)

0
-5

-4

-3

-2 -1 0 1
Log (strain rate)

-5

1800

2.1

1500

1.8
Fracture strain

Engineering stress (MPa)

0.4
0.2

Experiments

(c)

0.6
Johnson-Cook

Johnson-Cook

400

0.8

1200
900
600
300

Johnson-Cook
Experiments

0
0

0.1
0.2
0.3
Homologous temperature

-4

-3

-2 -1 0 1
Log (strain rate)

Experiments
Johnson-Cook

1.5
1.2
0.9
0.6
0.3

(e)

(f)

0
0.4

0.1
0.2
0.3
Homologous temperature

0.4

Fig. 19. Calibration of JohnsonCook model for Armox 500T steel (a) true stressstrain relation, (b) fracture strain as a function of stress triaxiality, (c) engineering stress as
a function of strain rate, (d) fracture strain as a function of strain rate, (e) engineering stress as a function of temperature and (f) fracture strain as a function of temperature.

5. Calibration of Armox 500T steel


The Armox 500T steel has been calibrated for obtaining the material parameters of JohnsonCook model by curve tting through
least square method. The steel plates have been found to be isotropic, hence the tests conducted at 0 orientation have been
employed for the calibration of the complete model. The static yield
strength, A, in the rst bracket of JC ow stress model (Eqn. 2) has
been obtained from the engineering stressstrain relationship
(Fig. 1(b)). The strength of the material has been found to be 1372,
1423 and 1319 MPa respectively at 0, 45 and 90 orientation. The
hardening parameters B and n of the rst bracket of the JC ow stress
n
model were obtained by tting the expression, B ( pl ) , in the measured true stressstrain curve by employing OriginPro 8.5.1 software
(Fig. 19(a)).
The stress-triaxiality parameters, D1, D2 and D3 of the equivalent failure strain (Eqn. 5), were obtained by tting the expression

of the second bracket of fracture strain model, D1 + D2 exp D3 m


with the observed true failure strainstress triaxiality curve
(Fig. 19(b)). The true failure strain of the specimen was obtained

d0
as 2ln , with d0 being the initial specimen diameter and df the
df
fractured diameter. An exponential curve has been tted with experimentally obtained data points using the software OriginPro. The
fracture strain has been found to decrease exponentially with increase in stress triaxiality which signies that the ductility decreases
with increase in stress triaxiality.
The strain rate sensitivity parameter C employed in the second
bracket of the JC ow stress model has been obtained by tting the

 pl
expression 1 + C ln with the measured yield stress versus strain
 0

rate curve (Fig. 19(c)). The yield stressstrain rate relationship indicates that the rate hardening is almost linear in the considered
regime. A linear curve has therefore been tted with the experimental data points as per the requirement of the model. The strain
rate dependent damage parameter D4 has been obtained by tting
the expression of the second bracket of the equivalent fracture strain,

 pl
1 + D4 ln  , with the observed true fracture strainstrain rate

0
relationship indicating decrement of fracture strain with increase
in strain rate (Fig. 19(d)) which has been assumed to be linear.

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

Table 4
Material parameters of Armox 500T steel.
Notation

Numerical value

Modulus of elasticity
Poissons ratio
Density
Yield stress constant
Strain hardening constant

E (N/m2)

(kg/m3)
A (N/m2)
B (N/m2)
n
C
m
0
melt (K)
transition (K)
D1
D2
D3
D4
D5

Cp (J/kg K)
K (W/m k)

201 109
0.33
7850
1372.488 106
835.022 106
0.2467
0.0617
0.84
1 s1
1800
293
0.04289
2.1521
2.7575
0.0066
0.86
0.9
455
47

Inelastic heat fraction


Specic heat
Thermal conductivity

The thermal sensitivity parameter m in the third bracket of the


JC ow stress model has been obtained by tting the corresponding expression 1 T m with the observed yield strength versus
temperature curve (Fig. 19(e)). A prerequisite stress of 1456 MPa,
corresponding to the 200 C temperature, was considered in order
to obtain the best t. The temperature dependent fracture strain
parameter D5 has been obtained by tting the corresponding expression of the fracture strain model, 1 + D5T , with the observed
failure strain versus homologous temperature curve (see Fig. 19(f)).
The calibrated material parameters of JC model for Armox 500T steel
are presented in Table 4.
6. Validation of Armox 500T steel parameters
The calibrated material parameters of the JC model have been
validated by simulating the high strain rate material tests performed on split Hopkinson pressure bar (SHPB) apparatus. The nite
element simulations for the material tests performed under tension
at strain rate 850 and 950 s1 were carried out on ABAQUS/Explicit
code. The axisymmetric model of the tension specimen was developed in accordance with its actual geometry (diameter 3 mm and
gauge length 10 mm) (see Fig. 20). The calibrated JC model was employed to assign the ow and fracture behavior of the Armox 500T
steel (see Table 4). The specimens were considered xed at one end

Actual

FEM

True stress (MPa)

3000

Description

Viscous effect
Thermal sensitivity
Reference strain rate
Melting temperature
Transition temperature
Fracture strain constant

157

2400
1800
Experimental results

1200

Mesh size - 0.05 mm


Mesh size - 0.07 mm

600

Mesh size - 0.09 mm

0
0

0.15

0.3
0.45
True strain

0.6

0.75

Fig. 21. Comparison of experimental and numerical stressstrain curves of Armox


500T steel obtained under tension at 950 s1 strain rate.

while the other end was under the application of pressure in the
direction opposite to the material. The pressure versus time curve
was assigned with respect to the corresponding strain rate. The discretization of the geometric model was carried out using four node
axisymmetric quadrilateral elements with stiffness hourglass control,
and three degrees of freedom at each node. The mesh convergence was also studied for the simulations performed for the
validation of material test. The element size in the specimen geometry was varied as 0.09 mm 0.09 mm, 0.07 mm 0.07 mm and
0.05 mm 0.05 mm. The true stressstrain relationship obtained by
numerically simulating the tension tests with different element sizes
considered has been compared with the corresponding experimental results at 950 s1 strain rate (see Fig. 21). A typical simulation
with element size 0.05 mm 0.05 mm and total number of 85 032
elements in the whole specimen took approximately 450 CPU
minutes. The actual and simulated stressstrain curves have been
found to have close correlation with respect to ow and fracture
behavior. As such, no signicant inuence of the element size has
been found on the numerical predictions. The contours of the normal
stresses have also been plotted in the specimen at the same strain
rate, 950 s1 (see Fig. 22(a)(c)). The ow of stress throughout the
specimen could be seen at various time steps before and after the
development of necking. The stress concentration in the necked
region is clearly visible at the onset of fracture. The variation in temperature of the specimen is also plotted at various time steps in
Fig. 22 (d)(f). The temperature of the specimen became substantial only after the formation of neck and reached maximum, 1197 K,
at the onset of fracture. The actual and predicted fractured diameter of the specimen was found to be 2.32 and 2.08 mm respectively.
7. Calibration of API projectiles
The API projectile material has been characterized by obtaining all the JohnsonCook model parameters by curve tting through
least square method. The yield strength, A, in the rst bracket of
the JC ow stress, Eqn. 2, was obtained from the engineering stress
strain curve (Fig. 12). The strain hardening parameters B and n of
the rst bracket of the JC ow stress model were obtained by tting
n
the expression B ( pl ) with the true stressstrain curve (Fig. 23(a)).
The stress triaxiality parameters, D1, D2 and D3 of the JC failure strain
model, Eqn. 5, were obtained by tting the expression
D + D exp D m
1
2
with the measured true failure strainstress
3

Fig. 20. Geometry of the axisymmetric tension test specimen of Armox 500T steel.

triaxiality curve (Fig. 23(b)). An exponential curve has been tted


with experimentally obtained data points using the software
OriginPro. The model accurately tted the experimental data points

158

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

(d

(c)

(b)

(a)

(e

(f

Fig. 22. (a)(c) Contour plots of normal stress in Pascal and (d)(f) temperature in Kelvin at different time intervals in the tension test specimens of Armox 500T steel at
950 s1 strain rate.

representing exponential decrement of fracture strain with increase in stress triaxiality.


The strain rate sensitivity parameter C in the second bracket of
the JC ow stress was obtained by tting the corresponding ex
 pl
pression of the second bracket, 1 + C ln , with the measured
 0

yield stress versus strain rate curve (Fig. 23 (c)). Due to limited
number of material specimens and the limitation of the experimental facility, the strain rate sensitivity of projectile could not be
studied corresponding to the entire range of strain rate which may
be developed in the material during the perforation process. It should
be noted, however, that the material has been found to be insensitive to strain rate such that the strength has been found
approximately identical in the considered strain rate regime (104
s110+3 s1) (see Fig. 16). The strain rate sensitivity parameter C has
been calibrated to be 0.0076, indicating insignicant inuence of
strain rate on the equivalent stress. Eventually, the strain rate dependent fracture strain parameter, D4, has been assumed to be zero
for the projectile material.
The thermal sensitivity parameter m given in the third bracket
of the JC ow stress was obtained by tting the corresponding expression 1 T m with the actual yield stresstemperature curve
(Fig. 23(d)). The nonlinear curve tted as per the requirement of the
model has been found to closely represent the observed pattern of
the decrement of ow stress with increasing temperature. The temperature dependent fracture strain parameter D5 was obtained by
tting the expression of the third bracket of fracture strain model,
1 + D5T , with the measured fracture straintemperature curve

(Fig. 23 (e)) in accordance with the constitutive relation. The material parameters of the API projectile thus obtained are enlisted
in Table 5.

Table 5
Material parameters of API projectile.
Description

Notation

Numerical value

Modulus of elasticity
Poissons ratio
Density
Yield stress constant
Strain hardening constant

E (N/m2)

(Kg/m3)
A (N/m2)
B (N/m2)
n
C
m
0
melt (K)
transition (K)
D1
D2
D3
D4
D5

Cp (J/kg K)
K (W/m k)

200 109
0.3
7850
1657.71 106
20855.6 106
0.651
0.0076
0.35
1 s1
1800
293
0.0301
0.0142
2.192
0.0
0.35
0.9
455
47

Viscous effect
Thermal sensitivity
Reference strain rate
Melting temperature
Transition temperature
Fracture strain constant

Inelastic heat fraction


Specic heat
Thermal conductivity

8. Ballistic experimentation
The ballistic experiments were carried out by ring the 7.62 API
projectiles through sniper rie on 8 mm thick Armox 500T steel plate
targets placed at 10 m distance from the rie, and by 12.7 API projectiles through air defense gun on 10 mm thick Armox 500T steel
plate targets placed at a distance of 15 m from the gun. The mass
of 7.62 API projectile was 5.5 grams and that of 12.7 API projectile
was 30.06 grams. Both of these projectiles were red at a constant

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

0.038

2400

Johnson-Cook model
Experiments

2000

0.036

1600
1200

Experiments
Johnson-Cook model

800

Fracture strain

True stress (MPa)

159

0.034

0.032
400

(a)

(b)

0.03
0

0.001 0.002 0.003 0.004 0.005


plastic strain

0.3

0.6
0.9
1.2
Stress triaxiality

1.5

2600
Engineerg stress (MPa)

Experiments

(c)

Johnson-Cook Model

2500
2400
2300
2200
2100
-4

-3

-2

-1 0 1 2
Log (Strain rate)

Experiments
Johnson-Cook Model

2000

Experiments
Johnson-Cook model

1.75
1.5
Fracture strain

Engineering stress (MPa)

2500

1500
1000

1.25
1
0.75
0.5

500

0.25

(d)
0

(e)

0
0

0.1

0.2
0.3
0.4
0.5
Homologus temperature

0.1
0.2
0.3
0.4
Homologus temperature

0.5

Fig. 23. Calibration of JohnsonCook model for API projectile material (a) true stressstrain relation, (b) fracture strain as a function of stress triaxiality, (c) engineering
stress as a function of strain rate, (d) engineering stress as a function of temperature and (e) fracture strain as a function of temperature.

incidence velocity of about 830 m/s at normal incidence. The xtures for holding the target as well as the gun were made of
structural steel. The square target plates of span 500 mm 500 mm
were held onto these xtures with the help of heavy nuts and bolts
and tightened effectively to enable the xed boundary. The target
holding xture could be suitably adjusted in horizontal and vertical plane in order to enable the adjustment of the impact location
on the target surface.
The impact and the residual velocities of the projectile were measured with the help of infrared optical sensors. Two such infrared
emitters were placed at 6 and 8 m from the muzzle of the sniper
rie and at 11 and 13 m from the muzzle of the air defense gun.
Two more emitters, each at 2 and 4 m behind the target, were placed
to measure the residual velocity. The projectiles were recovered after
perforation of the target with the help of the recovery platform of
the cushion pad bundles. The alignment of the gun, optical devices,
target and the projectile catcher was carefully maintained with

respect to the projectile trajectory. The Photron Fastcam-APX RS highspeed video camera system was also employed to measure the
residual projectile velocity and to record the perforation phenomenon. The camera was operated at about 9000 frames per second
frequency at a resolution of 640 480 pixels.
9. Numerical simulation for ballistic evaluation
The Armox 500T steel targets of thicknesses 8 and 10 mm were
modeled as three dimensional deformable continuum on ABAQUS/
CAE. The actual span of the target was 500 mm 500 mm; however,
in order to economize the computational problem, a reduced span
of 200 mm 200 mm was modeled. It was observed during the ballistic tests that the inuence of the impact does not spread beyond
this region. The projectiles were also modeled as three dimensional deformable body with the shank diameter, shank length, total
length and length of ogival nose 6.06, 20.75, 28.4, and 7.65 mm

160

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

850
0.55 mm
0.45 mm
0.35 mm
0.25 mm
0.2 mm

800
700
600

Residual velocity (m/s)

Residual velocity (m/s)

900

500
400
300

0.55 mm
0.45 mm
0.35 mm
0.25 mm
0.2 mm

800
750
700
650
(b)

(a)

200

600
0

20

40
Time ( s)

60

80

20

40
60
Time ( s)

80

100

Fig. 24. Residual velocity of (a) 7.62 and (b) 12.7 API projectile at different mesh congurations.

respectively for 7.62 API projectile, and 10.9, 24.4, 52.6 and 19.1 mm
respectively for 12.7 API projectile. As discussed above, the steel core
of the projectile has been modeled for all the nite element simulations, assuming that the brass jacket was stripped off and had no
inuence on the perforation process (Borvik et al. [30]). The ow
and fracture behavior of the target as well as projectile material were
modeled employing the JC model calibrated in the present study.
The target was meshed with eight node linear hexahedral elements with hourglass control. The mesh sensitivity for the target
was studied by varying the element size in the impact zone as 0.55,
0.45, 0.35, 0.25 and 0.2 mm3, giving 15, 18, 23, 32 and 40 elements at the thickness of 8 mm thick target and 19, 23, 29, 40 and
50 elements at the thickness of 10 mm thick target respectively. The
projectiles were impacted on the target corresponding to each mesh
conguration, at incidence velocity 835 m/s, and the residual velocities were found to be 424, 397, 345, 329 and 328 m/s respectively
for 7.62 API projectile and 682, 668, 650, 644 and 643 m/s respectively for 12.7 API projectile (see Fig. 24(a) and (b) respectively). Thus,
the residual velocity of 7.62 and 12.7 API projectiles converged corresponding to element size 0.25 mm3 and 0.35 mm3 respectively.
The residual velocities thus obtained have been found to have close
agreement with the actual results (see Fig. 25(a) and (b) respectively). The hexahedral elements of 1 mm3 were employed to
discretize the projectile. The contact between the projectile and target
was modeled by employing the kinematic contact algorithm available in ABAQUS. The projectile was considered as master and the
through thickness contact region of the target as node based slave
surface. A coecient of friction of 0.02 was assumed between the
projectile and target (Borvik et al. [30,44]). It should be noted that
the friction between the projectile and target is dependent upon
the normal pressure, sliding velocity and temperature between the
surfaces and as such does not remain constant. However, the ac-

400
380
360
340
320

(a)

Numerical results
Experimental results
680
660
640
(b)

620

300
0

The experimental and the numerical ballistic results for 8 mm


thick target against 7.62 API projectiles are presented in Table 6. The
residual velocity of the projectile has been predicted within 5% deviation from the experimental results. The projectile failed the target
by making a circular hole, producing minor spalling of material
around the hole and a bulge at the rear surface (Fig. 27). The diameter of the hole was 8.1 mm and the height of rear surface bulge
4.2 mm. The size of hole, bulging of the target at the rear surface
and the erosion of material from the front surface have been quite
accurately reproduced through numerical simulations (Fig. 27(a) and
(b)). Although spalling of the material at the front surface observed during experiments could not be exactly replicated, the high
stresses developed in that region during perforation process

Residual velocity (m/s)

Residual velocity (m/s)

420

10. Results and discussion

700

Numerical results
Experimental results

440

curate determination of the coecient of friction and its dependence


on different parameters during high speed penetration is highly complicated. Krafft [45] computed the energy loss during ballistic
penetration by measuring the torsional adhesion of spinning projectile during penetration through torsion-type Hopkinson bar by
assuming that the friction resisting rotation is equivalent to that resisting axial penetration. It was concluded that the total energy loss
due to friction during longitudinal motion is not more than 3% of
the total energy loss of the projectile. Bowden and Freitag [46]
studied the friction between the metal surfaces by trapping a spinning ball between symmetrically arranged at plates. It was
concluded that the sliding friction decreases with increase in the
velocity to vanishingly small. In the present study, the velocity of
the projectile is considerably high; therefore, the assumption of a
constant coecient of friction is quite reasonable. A typical nite
element model for projectile and target is shown in Fig. 26.

10

20
30
40
Number of elements

50

10
20
30
40
50
Number of elements (mm)

Fig. 25. Mesh convergence for (a) 8 mm and (b) 10 mm thick Armox 500T steel target.

60

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

161

Section X-X

Plan

Enlarged view of the impact region

Fig. 26. Typical nite element model of Armox 500T steel target and 12.7 API projectile.

Table 6
Ballistic resistance of 8 mm thick Armox 500T steel target against 7.62 API projectile.
Impact velocity (m/s)

Residual velocity (m/s)

823.62
828.02

Experimental results

Numerical results

334.28
343.74

349.6
358.1

Experiments

provided a fairly accurate assessment of the material behavior


(Fig. 27(a)).
The actual and the predicted residual projectile velocities of 12.7
API projectile are presented in Table 7 as a result of impact on 10 mm
thick Armox 500T steel target. The residual velocities are reproduced numerically within 5% deviation from the actual results. The
projectile made a circular hole with diameter of 13.12 mm,

Numerical simulation

(a)

(b)

Fig. 27. Failure modes of (a) front and (b) rear surface of 8 mm thick Armox 500T steel target against 7.62 API projectile.

162

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

Table 7
Ballistic resistance of 10 mm thick Armox 500T steel target against 12.7 API projectile.
Impact velocity (m/s)

Residual velocity (m/s)

841.89
835.21
831.67

Experimental results

Numerical results

686.37
Perforated
663.82

708.99
697.24
691.60

produced a small bulge at the rear surface of target and also caused
nominal spalling of material from the front surface, (Fig. 28). The
diameter of hole has been exactly predicted through the numerical simulations and the spalling of material around the hole has also
been reproduced (Fig. 28(a)). At the rear surface, however, no sign
of scabbing was observed (Fig. 28(b)), either through experiments
or nite element simulations. The failure mode at the rear surface
also indicated the signs of hole enlargement and this behavior was
conrmed through numerical simulations.
The numerical simulations were also carried out at different incidence velocities for obtaining the residual projectile velocities and
eventually the ballistic limit of 8 and 10 mm thick targets (see
Tables 8 and 9 respectively). The ballistic limit velocity, V50, was calculated as the average of the highest projectile velocity not giving
perforation and the lowest projectile velocity giving complete perforation of the target. The ballistic limit has been found to be
706.5 m/s for 8 mm thick target against 7.62 API projectile and
501 m/s for 10 mm thick target against 12.7 API projectile.
After the computation of ballistic limit velocity, the calculation
for residual projectile velocity corresponding to a given incidence
velocity has also been carried out using RechtIpson model (see
Tables 8 and 9). The residual velocities were calculated based on the
following empirical model originally proposed by Recht and Ipson
[47],
1

vr = a ( vip vblp ) p

(6)

Experiments

Table 8
Determination of ballistic limit of 8 mm thick target against 7.62 API projectile.
Impact
velocity (m/s)

823.62
760
730
720
716
710
703

Residual
velocity (m/s)

Model
constants

Numerical
results

RechtIpson
model results

358.10
229.37
161.30
131.40
101.20
39.20
0.00

373.82
237.38
148.93
109.12
89.65
51.38
0.00

1.0

1.8

Ballistic
limit (m/s)

706.5

where vi , vr and vbl are initial, residual and ballistic limit velocity, and a and p are the model constants. The least square method
was used in order to obtain the best t of the model with the numerical results for calibrating parameters a and p. For 8 mm thick
target against 7.62 API projectile, parameters a and p were found
to be 1 and 1.8, respectively, and for 10 mm thick target against 12.7
API projectile, 1 and 1.99, respectively. The calculated and numerically simulated residual projectile velocities have been found to have
close correlation for both the targets employed in the present study
(see Fig. 29).
11. Conclusions
The material behavior of Armox 500T steel and armor piercing
incendiary (API) projectiles has been characterized under varying
stress states, strain rate and temperature. The results of the material characterization enabled the calibration of the JohnsonCook
ow and fracture model. The ballistic trials have been conducted
on Armox 500T steel targets against 7.62 and 12.7 API projectiles
and the results thus obtained reproduced numerically on ABAQUS/

Numerical simulation

(a)

(b)

Fig. 28. Failure modes of (a) front and (b) rear surface of 10 mm thick Armox 500T steel target against 12.7 API projectile.

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

Table 9
Determination of ballistic limit of 10 mm thick target against 12.7 API projectile.
Impact
velocity (m/s)

Residual
velocity (m/s)

831.67
700
600
550
520
510
505
497

Model
constants

Numerical
results

RechtIpson
model results

691.60
491.90
337.25
224.03
146.20
104.34
66.30
0.00

663.83
488.88
330.15
226.93
139.28
95.39
63.44
0.00

1.0

1.99

Ballistic
limit (m/s)

Acknowledgment
The authors gratefully acknowledge the nancial support provided by ARMREB, DRDO through the research grant no. ARMREB/
MAA/2011/127 for carrying out the present study.
References

8 mm thick - Numerical results - 7.62 API


8 mm thick - Recht-Ipson model - 7.62 API
10 mm thick - Numerical results - 12.7 API
10 mm thick - Recht-Ipson model - 12.7 API

Residual velocity (m/s)

800

600

400

200

0
400

475

the residual projectile velocities for given incident velocities were


also calculated by the RechtIpson model. These showed a close
agreement with the values obtained through nite element
computations.

501

Explicit nite element code using the calibrated material models


for the target and the projectile.
The fracture strain of Armox 500T steel material decreased from
1.0 to 0.05 as the stress triaxiality increased from 0.33 to 1.95, indicating a decrease in ductility with increase in stress triaxiality.
The strain rate in the range 1.6 1043.3 102 s1 showed insignicant inuence on material behavior; however, as it increased to
850950 s1, the material showed high strain rate sensitivity. The
yield strength increased from 1365 to 1405 MPa with increase in
strain rate from 1.6 104 to 3.3 102 s1; however, at 950 s1 strain
rate, it reached 2207 MPa.
The strength of Armox 500T steel increased slightly when the
temperature increased up to 200 C due to blue brittle effect.
However, with further increase in temperature, 300900 C, the
strength decreased while the fracture strain increased consistently. The specimens underwent signicant elongation, 54%, at
900 C temperature.
The projectile material underwent insignicant plastic deformation under tension and experienced sudden failure at the ultimate
stress, 2300 MPa. The projectile material has also been found to be
insensitive to stress triaxiality and strain rate; however, it described high thermal sensitivity. The strength of material decreased
and the fracture strain increased consistently with increase in
temperature.
The nite element simulations have been carried out for computing residual velocities of the 7.62 and 12.7 API projectiles by
impacting these on 8 and 10 mm thick targets respectively at given
impact velocities. The computed residual velocities and modes of
deformations of both the plates matched the experiments very well.
The ballistic limit of these targets was computed numerically and

1000

163

550
625
700
Impact velocity (m/s)

775

850

Fig. 29. Comparison of numerical results with those obtained from RechtIpson
model.

[1] Zukas JA. High velocity impact dynamics. New York: Wiley; 1990.
[2] Johnson GR, Cook WH. Fracture characteristics of three metals subjected to
various strains, strain rates, temperatures and pressures. Engg Fracture Mech
1985;21:3148.
[3] Zerilli FJ, Armstrong RW. Dislocation-mechanics-based constitutive relations
for material dynamics calculations. J Appl Phys 1987;61:181625.
[4] Zerilli FJ, Armstrong RW. Constitutive relations for the plastic deformation of
metals. In: Schmidt SC, Shaner JW, Samara GA, Ross M, editors. High-pressure
science and technology. Colorado: American Institute of Physics, Colorado
Springs; 1993. p. 98992.
[5] Harding J. The development of constitutive relationships for material behaviour
at high rates of strain. In: Harding J, editor. Mechanical properties at high strain
rates of strain, Conf. Ser. No. 102: Session 5. Oxford, UK: Institute of Physics;
1989. p. 189203.
[6] Steinberg DJ, Cochran SG, Guinan MW. A constitutive model for metals
applicable at high-strain rate. J Appl Phys 1980;51:1498504.
[7] Steinberg DJ, Lund CM. A constitutive model for strain rates from 104 to 106
s1. J Appl Phys 1989;65:152833.
[8] Hancock JW, Mackenzie AC. On the mechanisms of ductile failure in highstrength steels subjected to multi-axial stress-states. J Mech Phys Solids
1976;24:14769.
[9] Mackenzie AC, Hancock JW, Brown DK. On the inuence of state of stress on
ductile failure initiation in high strength steels. Engrg Fracture Mech
1977;9:16788.
[10] Mirza MS, Barton DC, Church P. The effect of stress triaxiality and strain-rate
on the fracture characteristics of ductile materials. J Mater Sci 1996;31:45361.
[11] Holland D, Halim A, Dahl W. Inuence of stress triaxiality upon ductile crack
propagation. Steel Res 1990;61:5046.
[12] Borvik T, Hopperstad OS, Berstad T, Langseth M. A computational model of
viscoplasticity and ductile damage for impact and penetration. Eur J Mech A
Solid 2001;20:685712.
[13] Borvik T, Langseth M, Hopperstad OS, Malo KA. Ballistic penetration of steel
plates. Int J Impact Engg 1999;22:85586.
[14] Clausen AH, Borvik T, Hopperstad OS, Benallal A. Flow and fracture characteristic
of aluminum alloy AA 5083-H116 as function of strain rate, temperature and
triaxiality. Mat Sci Engg 2004;A364:26072.
[15] Iqbal MA, Senthil K, Bhargava P, Gupta NK. The characterization and ballistic
evaluation of mild steel. Int J Impact Engg 2015;78:98113.
[16] Wilkins ML, Streit RD, Reaugh JE. Cumulative-strain-damage model of ductile
fracture: simulation and prediction of engineering fracture tests. Tech. rep.
UCRL-53058, Lawrence Livermore Laboratory, University of California, Livermore,
CA 94550, 1983.
[17] Cockcroft MG, Latham DJ. Ductility and the workability of metals. J Inst Metals
1968;96:339.
[18] Samanta SK. Dynamic deformation of aluminium and copper at elevated
temperatures. J Mech Phys Solids 1971;19:11735.
[19] Johnson GR, Cook WH. A constitutive model and data for metals subjected to
large strains, high strain rates and high temperatures. In: American Defense
Preparedness Association, editor. Proceedings of the 7th International
Symposium on Ballistics. The Hague, Netherlands: 1983. p. 5417.
[20] Bammann DJ, Chiesa ML, Horstemeyer MF, Weingarten LI. Failure in ductile
materials using nite element simulations. In: Jones N, Wierzbicki T, editors.
Structural crashworthiness and failure. Liverpool, UK: Elsevier Applied Science;
1993. p. 154.
[21] Camacho GT, Ortiz M. Adaptive Lagrangian modelling of ballistic penetration
of metallic targets. Int J Comp Meth Appl Mech Engng 1997;142:269301.
[22] Preston DL, Tonks DL, Wallace DC. Model of plastic deformation for extreme
loading conditions. J Appl Phys 2003;93:21120.
[23] Abed F, Makarem F. Comparisons of constitutive models for steel over a wide
range of temperatures and strain rates. J Engg Mater Tech 2012;134:110.
[24] Wierzbicki T, Bao Y, Lee YW, Bai Y. Calibration and evaluation of seven fracture
models. Int J Mech Sci 2005;47:71943.
[25] Teng X, Wierzbicki T. Evaluation of six fracture models in high velocity
perforation. Engg Fracture Mech 2006;73:165378.
[26] Banerjee B. An evaluation of plastic ow stress models for the simulation of
high temperature and high strain rate deformation of metals, arXiv:cond-mat/
0512466v1. (2005) 143.
[27] Sjoberg T, Sundin KG, Oldenburg M. Calibration and validation of plastic high
strain rate models for alloy 718. XII International Conference on Computational

164

[28]

[29]

[30]

[31]
[32]
[33]
[34]

[35]

[36]

M.A. Iqbal et al. / International Journal of Impact Engineering 96 (2016) 146164

Plasticity. Fundamentals and Applications, September 35, Barcelona, Spain


(2013).
Iqbal MA, Chakrabarti A, Beniwal S, Gupta NK. 3D numerical simulations of
sharp nosed projectile impact on ductile targets. Int J Impact Engg 2010;37:
18595.
Brvik T, Clausen AH, Eriksson M, Berstad T, Hopperstad OS, Langseth M.
Experimental and numerical study on the perforation of AA6005-T6 panels.
Int J Impact Engg 2005;32:3564.
Borvik T, Dey S, Clausen AH. Perforation resistance of ve different high-strength
steel plates subjected to small-arms projectiles. Int J Impact Engg 2009;36:948
64.
Kilic N, Ekici B. Ballistic resistance of high hardness armor steels against 7.62
mm armor piercing ammunition. Mat Des 2013;44:3548.
Arsecularatne JA, Zhang LC. Assessment of constitutive equations used in
machining. Key Engg Mater 2004;2746, 277282.
Borvik T, Hopperstad OS, Berstad T, Langseth M. Numerical simulation of
plugging failure in ballistic penetration. Int J Solids Struct 2001;38:624164.
Gupta NK, Iqbal MA, Sekhon GS. Experimental and numerical studies on the
behavior of thin aluminum plates subjected to impact by blunt- and
hemispherical-nosed projectiles. Int J Impact Engg 2006;32:192144.
Buchar J, Voldrich J, Rolc S, Lisy J. Ballistic performance of dual hardness armor.
In: Carleone J, Orphal D, DEStech Publications, editors. 20th International
Symposium on Ballistics. Orlando, Florida, September, 2002. p. 237.
Niezgoda T, Morka A. On the numerical methods and physics of perforation
in the high-velocity impact mechanics. World J Engg 2009;41416.

[37] Bridgman PW. Studies in large plastic ow and fracture. New York: McGraw-Hill;
1952.
[38] Hopperstad OS, Borvik T, Langseth M, Labibes K, Albertini C. On the inuence
of stress triaxiality and strain rate on the behavior of structural steel. Part I.
Experiments. Eur J Mech A Solid 2003;22:113.
[39] Thomson RD, Hancock JW. Ductile failure by void nucleation, growth and
coalescence. Int J Fracture 1984;26:99112.
[40] Kolsky H. Stress waves in solids. Oxford: Clarendon Press; 1953.
[41] Kilic N, Bedir S, Erdik A, Ekici B, Tasdemirci A, Guden M. Ballistic behavior of
high hardness perforated armor plates against 7.62 mm armor piercing
projectile. Mat Des 2014;63:42738.
[42] ABAQUS/Explicit Users Manual, Version 6.8, 2008.
[43] Hillerborg A, Modeer M, Petersson PE. Analysis of crack formation and crack
growth in concrete by means of fracture mechanics and nite elements. Cem
Conc Res 1976;6:77382.
[44] Borvik T, Olovsson L, Dey S, Langseth M. Normal and oblique impact of small
arms bullets on AA6082-T4 aluminum protective plates. Int J Impact Engg
2011;38:57789.
[45] Krafft JM. Surface friction in ballistic penetration. J Appl Phys 1955;26:124853.
[46] Bowden FP, Freitag EH. The friction of solids at very high speeds. I. Metal on
metal; II. Metal on diamond. Proc Royal Soc 1958;248:35061.
[47] Recht RF, Ipson TW. Ballistic perforation dynamics. J Appl Mech 1963;30:384
90.

Вам также может понравиться