Anda di halaman 1dari 13

International Journal of Impact Engineering 89 (2016) 1–13

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

Ballistic impact response of Kevlar® reinforced thermoplastic


composite armors
Aswani Kumar Bandaru a,*, Vikrant V. Chavan b, Suhail Ahmad a, R. Alagirusamy b,
Naresh Bhatnagar c
a Department of Applied Mechanics, Indian Institute of Technology Delhi, New Delhi, India
b Department of Textile Engineering, Indian Institute of Technology Delhi, New Delhi, India
c
Department of Mechanical Engineering, Indian Institute of Technology Delhi, New Delhi, India

A R T I C L E I N F O A B S T R A C T

Article history: The ballistic impact response of thermoplastic-based composite armors made from Kevlar® fabric and
Received 14 May 2015 polypropylene (PP) matrix has been investigated against ballistic test standard NIJ-STD 0106.01 Type IIIA.
Received in revised form 27 October 2015 Kevlar® fabrics of different architectures, namely 2D plain woven, 3D orthogonal and 3D angle interlock
Accepted 28 October 2015
fabrics, were produced and used as reinforcements to fabricate composite armor panels, using compres-
Available online 4 November 2015
sion molding technology. Interfacial property between PP and Kevlar® was improved by adding a coupling
agent called maleic anhydride grafted PP. Reduced density was observed in Kevlar® thermoplastic-
Keywords:
based composites as compared to that of the thermoset-based laminates. Ballistic impact tests were
Kevlar®
Thermoplastic matrix imparted with 9 mm full metal jacket (FMJ) on armor panels having different fabric architecture. Ballis-
Ballistic limit tic test results revealed that 2D armor was 2.4–7% more susceptible to damage than 3D armors. Hydrocode
Ballistic impact simulations were carried out using ANSYS AUTODYN v. 14.0 to obtain an estimate for the ballistic limit
Hydrocode simulations velocity and simulate failure modes. Post-impact damage patterns obtained from the simulations were
compared with the experimental results to assess the performance of the simulations. Good correlation
between the hydrocode simulations and experiments was found, both in terms of failure modes and damage
patterns. 3D composite armors were able to confront the 9 mm FMJ projectile; however, the 2D plain
woven armors failed. The increase in the ballistic limit from 2D plain woven armor to 3D orthogonal and
3D angle interlock armors was 16.44% and 20%, respectively, indicating the effect of fabric architecture.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction ical simulations, and analytical models. Silva et al. [11] investigated
the ballistic impact response of Kevlar® 29/Vinylester panels im-
With the requirement of light weight body armors, the need for pacted with a fragment simulating projectile (FSP) (320–360 m/
performance improved fiber reinforced composites is significantly s). Numerical simulations were carried out using AUTODYN
increasing. The architecture of the fabric plays a significant role in commercial software. Post-impact damage patterns obtained from
the protection against ballistic impact and provides a unique bal- the simulations were in good agreement with the experimental
listic penetration resistance for varied orientations. The main results. Further, simulations were extended to obtain ballistic limit
structural parameters of the fabric, which shows the effect on the velocity and residual velocity. Similarly, hydrocode simulations were
ballistic performance, are type of weave (with a twist in the yarns), carried out using AUTODYN to investigate the ballistic response of
yarn crimp, fabric structure, projectile geometry, impact velocity and a Kevlar® helmet by Tham et al. [12]. The response of the helmet
friction [1–5]. from simulations was compared with the ballistic impact test results
The ballistic impact response of thermoset-based composite lami- in terms of post-impact damage. Further, simulations were ex-
nates, such as S2 glass/polyester [6], E-glass/epoxy [7,8], S2 glass/ tended to assess the ballistic resistance of helmet against NIJ-STD-
epoxy [9,10], Kevlar® 29/Vinylester [11] and Kevlar®/epoxy [12], was 0106.01 Type II (9 mm FMJ) and 1.1 g FSP. It was found that the
investigated by several researchers through experiments, numer- helmet was able to stop both the projectiles. Grujicic et al. [13] de-
veloped a material model based on the unit cell method and
integrated with ANSYS/AUTODYN as a user defined subroutine. Dif-
* Corresponding author. Department of Applied Mechanics, Indian Institute of
ferent stages of armor penetration such as shearing of the filament,
Technology Delhi, New Delhi, India. Tel.: +91 11 26596412. delamination, and stretching of the filament on the back face of the
E-mail address: aswani006@gmail.com (A.K. Bandaru). target were observed. Kevlar® fabric was also used as a hybrid layer

http://dx.doi.org/10.1016/j.ijimpeng.2015.10.014
0734-743X/© 2015 Elsevier Ltd. All rights reserved.
2 A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13

in hybrid composites with other laminates reinforced with carbon the ballistic impact response of Kevlar® thermoplastic laminates ac-
and glass fibers [14–17]. cording to NIJ-STD-0106.01 Type IIIA [27].
Designing of body armors based solely on the experimental data In the present study, Kevlar® 29 yarns were woven to get fabrics
requires huge material and manpower, which is time consuming with three different architectures, namely 2D plain woven (2D-P),
and also uneconomical. Recent advances concerning the ballistic 3D orthogonal (3D-O), and 3D angle interlock (3D-A). Composite
impact response of composite laminates offer the possibility of pre- armor panels were fabricated with PP matrix reinforced with the
venting tests by using numerical simulations such as hydrocodes above three types of fabrics using vacuum assisted compression
[11–14,17] that can reduce the expenses incurred in designing of molding machine. The interfacial property between the Kevlar® and
the body armors. Due to the continuous development of numeri- PP was improved by adding a coupling agent called maleic anhy-
cal algorithms and material models, the accuracy and the dride grafted (MAg)-PP. The objectives of the present work were dual.
applicability of simulation results are increasing. The first objective was to perform a ballistic impact test on Kevlar®/
All the above discussed studies have been concentrated on the MAg-PP (K-MPP) composite armor panels for their perforation
ballistic impact behavior of thermoset-based composite lami- capability against the ballistic test standard NIJ-STD-0106.01 Type
nates. Though thermosets were extensively used as matrix materials, IIIA when impacted by a 9 mm FMJ projectile. The response ob-
the use of these matrices is limited due to a few shortcomings, tained from the ballistic test was used as a benchmark for later
namely, the need for low temperature storage and a long curing comparison with that obtained from hydrocode simulations. Post-
process. Thermoplastic-based composites, on the other hand, are impact damage patterns of the armors were acquired to determine
an alternative to thermoset-based composites due to their long shelf the extent of damage due to different failure modes and fabric struc-
life, short processing time, sufficiently tough, chemical resistant melt- tures. The second objective was to study the influence of fabric
processability, and an ability to be recycled [18,19]. Also, architecture on the ballistic impact response of K-MPP laminates.
thermoplastic composites have relatively low brittleness transi- For the same mass and geometry of the projectile, ballistic limit ve-
tion temperatures, which allow potential improvements in terms locity and energy absorbed by the target were compared for the three
of greater ballistic resistance, higher mechanical toughness, and faster types of thermoplastic-based Kevlar® armor panels, and new find-
manufacturing cycles [20,21]. Studies of Walsh et al. [20,21] re- ings on their ballistic impact response were reported.
ported that the thermoplastic aramid based composites exhibit
improved ballistic performance at a much lighter weight. Song [22] 2. Experimental
studied the influence of microscopic and macroscopic characteris-
tics on the ballistic impact response of thermoplastic composites 2.1. Materials
made of Kevlar® 29/nylon 66, Kevlar® 29/polyetheretherketone
(PEEK), Kevlar® 29 /polycarbonate, Kevlar® 29/Polysulfone, Kevlar® The high performance aramid (Kevlar® 29) fiber tow was con-
KM2/Polysulfone, and Kevlar® KM2 fiber/linear low-density poly- sidered with a linear density of 1000 Denier. Fabrics with three solid
ethylene (LLDPE) laminates. The characteristics, like processing woven structures viz. 2D-plain woven (2D-P), 3D-orthogonal (3D-
temperature, cooling rate, fabric configuration, fiber wetting, polymer O) and 3D-angle interlock (3D-A) were prepared using CCI sample
morphology, and stiffness of the laminate, significantly affected the weaving machine with rapier weft insertion mechanism. Physical
ballistic performance of the composite armor. The major energy ab- parameters of the fabric are given in Table 1a and Table 1b. Micro-
sorbing mechanisms observed were fiber breakage, fiber straining, scopic views of the woven structure are shown in Fig. 1. These fabrics
matrix cracking, and delamination. The review work of Kulkarni et al. were produced with two warp beams, one containing the binding
[19] stated that the thermoplastics have lower tensile strength than yarns and the other containing the ground yarns.
thermosets; as a result, ballistic performance was reduced. There- PP sheets were produced with two different grades, namely,
fore, thermoplastics are used with high ductile fibers like Kevlar® MI3530 and CO15EG, using in-house extrusion facility with nitro-
to enhance the matrix stiffness. Carrillo et al. [23] has studied the gen gas at a pressure of 150 bar.
ballistic response of Kevlar® 129/PP laminates through experimen-
tal studies. The addition of PP matrix to aramid fabrics showed 2.2. Fabrication of laminates
improved impact resistance. However, low adhesion was reported
between Kevlar® fabrics and PP matrix, suggesting for improve- The vacuum assisted compression molding technique was used
ments in the interfacial property. Further, it was reported that for the consolidation of stacked fabrics and resins. The specimens
numerical modeling is required to validate the results obtained were cured at 200 °C. Fiber weight fractions obtained for 2D-P, 3D-O
through experiments. and 3D-A were 60.2%, 64%, and 64%, respectively. Three types of spec-
After a thorough literature review, it is observed that the bal- imen were prepared: first, sixteen layer 2D-P laminate; second, eight
listic impact behavior of thermoset two dimensional (2D) Kevlar®
composite laminates was investigated with woven or unidirection-
al fabrics. The laminates with 2D plain woven fabrics evidence the Table 1
presence of crimp, exhibit poor in-plane stiffness, and suffer more (a) Physical parameters of Kevlar® 29 fabrics. (b) Properties of constituent materials.
damage due to delamination. On the other hand, in 3D fabrics, warp
(a)
and fill tows do not have any crimp, and yarns in z-direction play
Property 2D-P 3D-A 3D-O
a vital role in holding all warp and fill yarns together. The 3D struc-
ture of fabric provides increased areal density, thus increasing the Warp yarns/inch 40 40 40
amount of specific energy absorption [24]. Weft yarns/inch 32 120 120
Areal density (g/m2) 363.75 745.86 780.36
Studies on Kevlar®/PP composite laminates are not conclusive Thickness (mm) 0.64 1.17 1.24
due to low adhesion problem between aramid and PP [23,25,26].
(b)
Further studies are required in terms of enhancing the interfacial
property between Kevlar® fabric and PP matrix. Also, numerical val- Property Tenacity (gpd) Strain (%) Modulus (gpd)
idation is necessary to confirm the experimental results [23]. Ballistic ®
Kevlar yarn 14.91 (6.13)a
2.99 (5.18) 547.30 (10.11)
impact resistance of thermoplastic-based body armors reinforced Polypropylene sheet 17.93 (14.23) 4.08 (14.20) 98.14 (19.18)
with 3D fabrics was also not reported in the open literature. To the a
Value enclosed in parentheses indicates the coefficient of variance of corre-
author’s knowledge, there is no literature available to investigate sponding 15 readings for each sample.
A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13 3

(a) (b) (c)

Fig. 1. Microscopic views of woven samples: (a) 2D-P (b) 3D-O (c) 3D-A.

layer 3D-O laminate; and the third, eight layer 3D-A laminate. The length of the specimen considered was 160 mm, overall gauge length
cross-sectional views of 2D-P, 3D-O and 3D-A laminates are shown was 60 mm, and gauge length of the extensometer was 25 mm. Fig. 4
in Fig. 2. shows the failure in the tested specimen, and it was observed that
Uniform diffusion of PP was observed throughout the lami- the samples were failed within the region measured using the ex-
nate. The density of all the laminates was 4.8–29% (Table 2) less than tensometer. However, the main objective of the present study was
that of the thermoset-based Kevlar® reinforced composites re- the fabrication of a novel thermoplastic-based composite armor panel
ported in the literature. The reduction in the density indicates the and its ballistic response against NIJ-STD 0106.01 Type IIIA standard.
reduced weight of the armor panels.
While fabricating the laminates by using pure PP as a matrix, 2.4. Ballistic test
dispersion of matrix in the fibers was incomplete. This was due to
the reason that PP resins are not cross-linked polymers and possess The ballistic test was performed at the army camp near
high melt viscosity. Therefore, these are characterized by relative- Kapurthala, Punjab, India. The test was carried out at a tempera-
ly low strength and poor creep resistance at slightly elevated ture of 39 °C and humidity of 78%. Each panel was impacted with
temperatures. Due to this nature, they form weak bonds with Kevlar® five rounds as shown in Fig. 5. The firing locations were chosen ac-
fibers [25,26]. Therefore, the adhesion between Kevlar® fabric and cording to NIJ-STD 0106.01, where the location of each shot was not
PP matrix was improved by adding a coupling agent called MAg- affected by previous impact damage. Three types of armor panel were
PP to PP. To select the best suitable combination of MAg-PP and PP, manufactured according to standard body armor dimensions
two layer Kevlar®/PP laminates were fabricated with different com- (301.6 mm × 253.8 mm). For the ballistic test of K-MPP laminate
binations of MAg-PP with MI3530 and CO15EG resins. Low velocity based on the NIJ-STD 0106.01 Type IIIA [27], target was impacted
impact tests were conducted on these laminates using Fractovis Plus by a 9 mm full metal jacket (FMJ) projectile within the velocity range
drop weight impact test machine with an impact velocity of 5.03 m/ of 350–440 m/s. The initial impact velocity is measured by placing
s. Laminates were tested according to ASTM D7136 standard [29] optical sensors in the field which are connected to a data acquisi-
for three specimens with respect to same impact conditions, and tion system.
the corresponding observations are presented in Table 3. From
Table 3, it can be concluded that the combination of 10%MAg- 3. Material modeling
PP + MI3530 exhibited higher energy absorbing capability. Therefore,
final armor panels were manufactured using the matrix combina- The ballistic impact resistance of K-MPP armors with different
tion of 10%MAg-PP + MI3530 PP. fabric architecture was assessed using ANSYS-AUTODYN-3D v. 14.0,
a commercial hydrocode. Hydrocodes are large computer pro-
2.3. Tensile testing of laminates grams that are particularly designed to simulate the response of
structures undergoing large deformation, at high strain rate con-
The basic mechanical properties of the K-MPP composite lami- ditions. An EOS was used to define the relationship between pressure,
nates were obtained by performing quasi-static tensile tests on the density and internal energy by taking account of changes in density
Zwick/Roell universal testing machine (UTM) (Fig. 3). and thermodynamic processes. A strength relation was used to define
The tensile properties of the laminates were measured in ac- suitable equivalent plastic strain, equivalent plastic strain rate and
cordance to ASTM D3039 standard [30]. Three specimens were tested temperature dependence of the yield surface. In addition, a failure
for each material. The dimensions of the samples were selected ac- criteria was implemented i.e. a relation illustrating the (hydrostatic/
cording to ASTM D-3039 standard. According to this standard, the deviatoric) stress and/or strain condition which, when fulfilled, causes

Weft yarns

Weft yarns Binder yarns


Binder yarns

Stuffer yarns
Weft yarns
Warp yarns Stuffer yarns

(a) (b) (c)

Fig. 2. Cross-sectional views: (a) 2D-P (b) 3D-O and (c) 3D-A.
4 A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13

Table 2
Comparison of density.

Composite laminate Density


(g/cm3)

Kevlar® 29/Vynilester [11] 1.40


Kevlar® 29/Epoxy [12] 1.65
Kevlar® 29 & 129/polyvinylbutyral [28] 1.23
Kevlar® 29 (2D-P)/MAg-PP (Present Study) 1.17
Kevlar® 29 (3D-O)/MAg-PP (Present Study) 1.19
Kevlar® 29 (3D-A)/MAg-PP (Present Study) 1.19

Fig. 3. Zwick/Roell UTM cross head holding K-MPP specimen.


the material to break and drop its ability to support normal and shear
stresses.
An advanced material model [31,32], particularly designed to sim-
ulate the shock response of anisotropic materials, was implemented
⎡⎣C k ⎤⎦ = [T k ] [C k ][T k ]
−1 −T
in ANSYS-AUTODYN v 14.0. This model assumes that the compos- (2)
ite material can be idealized as an orthotropic material and considers
anisotropic strength degradation, material anisotropy, shock re- where C k is the elastic stiffness matrix of kth lamina within the prin-
sponse and coupling of volumetric and deviatoric response. cipal coordinate system of laminate and Tk is the transformation
Therefore, in the present study, the target was modeled using ortho matrix and is expressed as:
EOS along with the elastic strength model and material stress/
strain failure criteria. Proper erosion strain and Lagrange processor ⎡ C2 S2 0 0 0 S2 ⎤
were also utilized in the present simulations. ⎢ S2 C2 0 0 0 − S 2⎥⎥

⎢ 0 0 1 0 0 0 ⎥
3.1. Equation of state (EOS) [T k ] = ⎢⎢ 0 0 0 C −S 0 ⎥
⎥ (3)
⎢ 0 0 0 S C 0 ⎥
In general, the behavior of composite laminates can be repre- ⎢ ⎥
sented through a set of orthotropic constitutive relations. The ⎢ −S2 S2 ⎥
0 0 0 C2 ⎥
orthotropic EOS in ANSYS-AUTODYN allows a non-linear EOS to be ⎣⎢ 2 2 ⎦
used in conjunction with an orthotropic stiffness matrix [31]. The
where C2 = cos2 θk, S2 = sin2 θk, S2 = sin 2θk and C2 = cos 2θk.
stress tensor obtained from these constitutive relations is divided
To include non-linear shock effects in the above linear rela-
into hydrostatic (pressure) and deviatoric components. This ap-
tions, it is first desirable to separate the volumetric (thermodynamic)
proach incorporates the possible occurrence of non-linear effects
response of the material from its ability to carry shear loads
(shocks) that can be attributed to the volumetric straining in the
(strength). The strain components are split into their average ε ave
material. The stress–strain relation for an orthotropic material follows
and ε ijd deviatoric components. The constitutive relation becomes
as:
[33]:
⎡σ 11 ⎤ ⎡C11 C12 C13 0 0 0 ⎤ ⎡ ε11 ⎤
⎢σ ⎥ ⎢C ⎡ d 1 ⎤
C 22 C 23 0 0 0 ⎥ ⎢ ε 22 ⎥ ⎢ ε11 + 3 ε vol ⎥
⎢ 22 ⎥ ⎢ 21 ⎥⎢ ⎥
⎢σ 33 ⎥ ⎢C 31 C 32 C 33 0 0 0 ⎥ ⎢ ε 33 ⎥ ⎡σ 11 ⎤ ⎡C11 C12 C13 0 0 0 ⎤⎢ ⎥
⎢σ ⎥ ⎢C 1
⎢ ⎥=⎢ ⎥⎢ ⎥ (1) C 22 C 23 0 0 0 ⎥ ⎢ε 22d
+ ε ⎥
⎢ τ 23 ⎥ ⎢ 0 0 0 C 44 0 0 ⎥ ⎢γ 23 ⎥ ⎢ 22 ⎥ ⎢ 21 ⎥⎢ 3
vol

⎢ τ 31 ⎥ ⎢ 0 ⎢σ 33 ⎥ ⎢C 31 C 32 C 33 0 0 0 ⎥⎢ ⎥
0 0 0 C 55 0 ⎥ ⎢γ 31 ⎥ ⎢ ⎥=⎢ ⎥⎢ d 1 ⎥ (4)
⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ τ 23 ⎥ ⎢ 0 0 0 C 44 0 0 ⎥ ⎢ε 33 + ε vol ⎥
⎣ τ 12 ⎦ ⎣ 0 0 0 0 0 C 66 ⎦ ⎣γ 12 ⎦ 3
⎢ τ 31 ⎥ ⎢ 0 0 0 0 C 55 0 ⎥⎢ ⎥
⎢ ⎥ ⎢ ⎥⎢ γ 23 ⎥
where Cij is the constitutive coefficient matrix; σij, εij are the prin- ⎣ τ 12 ⎦ ⎣ 0 0 0 0 0 C 66 ⎦ ⎢ γ 31 ⎥
cipal stresses and strains; and τij, γij are the shear stress and shear ⎢ ⎥
strains, respectively. It is also useful, since within the AUTODYN en- ⎣ γ 12 ⎦
vironment, each layer of the laminate is not explicitly modeled;
where ε vol = ε11 + ε 22 + ε 33 and ε ave = 31 (ε11 + ε 22 + ε 33 ) .
rather, continuum elements representing equivalent homoge-
Pressure is defined as one-third of the trace of the stresses
neous anisotropic solids are used to represent thick laminates
consisting of a number of repeating lamina. The principal materi- 1
al directions of all the lamina in the laminate do not necessarily P=− (σ 11 + σ 22 + σ 33 ) (5)
3
coincide with the global axis of the laminate. Stiffness matrix for
each lamina has to be expressed within the laminate global mate- By expanding Eq. (4), separating deviatoric and volumetric terms
rial axis using the following transformation: grouped separately, the expressions for the direct stresses are:

Table 3
Impact response parameters of Kevlar® laminate with different PP combinations.

Kevlar® fabric + combination Vel. (m/s) Damage force (N) Peak def. (mm) Peak energy (J) Peak force (N) Total energy (J)

KF+5% MAg-PP + MI3530 5.03 29.708 20.78 62.556 4885.202 74.774


KF+5% MAg-PP + CO15EG 5.03 49.91 19.72 71.397 6344.464 82.486
KF+10% MAg-PP + MI3530 5.03 61.793 24.394 125.344 8861.335 133.287
KF+10% MAg-PP + CO15EG 5.03 39.215 18.439 53.498 5765.75 92.672
A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13 5

include non-linear terms. The first term is replaced with the Mie–
Grüneisen EOS, and the remaining terms act as a correction due to
deviatoric strains. The Eq. (7) takes the following form:

1
[C11 + C 21 + C 31 ] Δε11d
ΔP = ΔPEOS (ε vol , e ) −
3 (9)
1 1
− [C12 + C 22 + C 32 ] Δε 22
d
− [C13 + C 23 + C 33 ] Δε 33
d
3 3
Γ( v )
where the pressure contribution PEOS = Pr (ε vol ) + v [e − er (ε vol )]from
volumetric strains can include the non-linear shock (thermody-
Fig. 4. Uniaxial tensile test specimen. namic) effects and energy dependence as in a conventional EOS. The
parameters Pr (ε vol ) and er (ε vol ) define the material pressure volume
and energy–volume relation along the Hugoniot reference curve.
1 The Grüneisen gamma is defined as:
σ 11 = (C11 + C 22 + C13 ) ε vol + C11ε11
d
+ C12ε 22
d
+ C13ε 33
d
(6a)
3
⎛ ∂P ⎞
Γ (v ) = v ⎜ ⎟ (10)
1 ⎝ ∂e ⎠ v
σ 22 = (C 21 + C 22 + C 23 ) ε vol + C 21ε11
d
+ C 22ε 22
d
+ C 23ε 33
d
(6b)
3
The Grüneisen gamma allows the determination of thermody-
1 namic states away from reference Hugoniot states. Three forms of
σ 33 = (C 31 + C 32 + C 33 ) ε vol + C 31ε11
d
+ C 32ε 22
d
+ C 33ε 33
d
(6c) Grüneisen gamma EOS are available to link with an orthotropic re-
3
sponse in the model: linear EOS, shock EOS and polynomial EOS.
Substituting Eq (6a–c) into Eq (5), the following expression can The volumetric response of the material is defined through the solid
be obtained for pressure: EOS. The shock EOS (Eq. 11) allows the coupling of non-linear EOS
with orthotropic material stiffness.
1 1
P=− [C11 + C 22 + C 33 + 2(C12 + C 23 + C 31 )]ε vol − [C11 + C12 + C13 ]ε11d K
9 3 C0 = (11)
1 1 ρ
− [C 21 + C 22 + C 23 ]ε 22
d
− [C 31 + C 32 + C 33 ]ε 33
d
(7)
3 3
where C0 is the bulk acoustic sound speed, K is the effective bulk
The first term on the right side of Eq. (7) represents the stan- modulus, and ρ is the current material density. The above materi-
dard relationship between the pressure and volumetric strain al model has been used by other authors for simulating the non-
(Hooke’s law) at low compressions. The later terms comprise of cou- linear stress–strain relationships for Kevlar composites [11,12,14].
pling between the pressure and deviatoric strain. The contribution
to the pressure from volumetric and deviatoric components of strain 3.2. Failure model
can clearly be identified in Eq. (7). The first term of Eq. (7) can, there-
fore, be used to define the volumetric (thermodynamic) response Laminated composites exhibit multi-failure modes under bal-
of an orthotropic material in which the effective bulk modulus of listic impact. The failure initiation can be based on any combination
the material K is defined as of material stress and/or strain. After failure initiation, stiffness and
strength properties of the failed material are updated based on the
1
K= (C11 + C 22 + C 33 + 2(C12 + C 23 + C 31 )) (8) mode of the failure initiation. In a laminate, with 11-direction
9 through its thickness, the stress in 11-direction is set to zero, if de-
Under high strain rate ballistic impact conditions, the relation lamination occurs due to through-the-thickness tensile stresses (or
between pressure and volumetric strain is typically non-linear. For strains) or from excessive shear stress (or strain) results in delami-
the inclusion of non-linear shock effects, the contribution to pres- nation. If the failure is initiated in either of these two modes, the
sure from volumetric strain (first term in Eq. (7)) is modified to stress in the 11-direction is instantaneously set to zero, and the strain
in the 11-direction of failure is stored. Subsequently, if the mate-
rial strain in the 11-direction exceeds the failure strain, the material
stiffness matrix is modified as [33]:
11-direction:

⎡ d 1 ⎤
⎢ ε11 + 3 ε vol ⎥
⎡σ 11 ⎤ ⎡0 0 0 0 0 0 ⎤⎢ ⎥
⎢σ ⎥ ⎢0 C 1
C 23 0 0 0 ⎥ ⎢ε 22 d
+ ε ⎥
⎢ ⎥ ⎢
22 22
⎥⎢ 3
vol

⎢σ 33 ⎥ ⎢0 C 32 C 33 0 0 0 ⎥⎢ ⎥
⎢ ⎥=⎢ ⎥⎢ d 1 ⎥ (12)
τ
⎢ 23 ⎥ ⎢ 0 0 0 α C 44 0 0 ⎥ ⎢ε 33 + ε vol ⎥
3
⎢ τ 31 ⎥ ⎢0 0 0 0 αC 55 0 ⎥⎢ ⎥
⎢ ⎥ ⎢ ⎥⎢ γ 23 ⎥
⎣ τ 12 ⎦ ⎣0 0 0 0 0 αC 66 ⎦ ⎢ γ 31 ⎥
⎢ ⎥
⎣ γ 12 ⎦

where α is the residual shear stiffness.


Sometimes, reduction in the shear stiffness results in delami-
nation. A nominal value of 20% is typically used for the parameter
Fig. 5. Impact locations and firing order for testing. α [34]. The 22- and 33-directions are assumed to be in the plane
6 A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13

of the composite, i.e. in fiber directions. Due to excessive stresses 4. Numerical simulations
and/or strains in 22- and 33-directions, in-plane failure mode
occurs. If failure is initiated in these two modes, the stress in the The main objective of the present study is to determine if K-MPP
failed direction is instantaneously set to zero, and the strain at failure armor with different fabric architectures can conform to NIJ Standard-
in the direction of failure is stored. The stiffness matrix is modi- 0106.01 Type IIIA, 9 mm FMJ (full metal jacket). In the numerical
fied as: simulations, the projectile impacted the target with an initial impact
22-direction: velocity within the range of 350–440 m/s. The actual dimensions
of the panel used in the ballistic tests were equivalent to the stan-
⎡ d 1 ⎤ dard body armor dimensions, i.e. 301.6 mm × 253.8 mm. As the
⎢ ε11 + 3 ε vol ⎥
⎡σ 11 ⎤ ⎡C11 0 C13 0 0 0 ⎤⎢ ⎥ ballistic impact phenomenon is localized in nature, only a region
⎢ 0 ⎥ ⎢ 0 1
0 0 0 0 0 ⎥ ⎢ε 22 d
+ ε ⎥ of the real panel was considered. For the selection of panel dimen-
⎢ ⎥ ⎢ ⎥⎢ 3
vol
⎥ sions for simulation, 2D axisymmetric simulations were carried out
⎢σ 33 ⎥ ⎢C13 0 C 33 0 0 0 ⎥⎢ ⎥
⎢ ⎥=⎢ ⎥⎢ d 1 ⎥ (13) on the 2D-P armor panel with three different dimensions in the y
⎢ τ 23 ⎥ ⎢ 0 0 0 αC 44 0 0 ⎥ ⎢ε 33 + ε vol ⎥ direction by keeping the thickness constant. The dimensions con-
3
⎢ τ 31 ⎥ ⎢ 0 0 0 0 αC 55 0 ⎥⎢ ⎥ sidered were Panel A (2.5 × 75 mm2), Panel B (2.5 × 100 mm2) and
⎢ ⎥ ⎢ ⎥⎢ γ 23 ⎥
⎣ τ 12 ⎦ ⎣ 0 0 0 0 0 αC 66 ⎦ ⎢ Panel C (2.5 × 150 mm2). Both the edges of the panels were fixed.
γ 31 ⎥
⎢ ⎥ The panel was impacted with an impact velocity of 426 m/s, and
⎣ γ 12 ⎦ the damage propagation was observed. Fig. 6 (a–c) shows the damage
propagation in different panels at different time cycles till the com-
33-direction:
plete penetration of the projectile.
⎡ d 1 ⎤ From Fig. 6 (a–c) it was found that for an impact velocity of
⎢ ε11 + 3 ε vol ⎥ 426 m/s, the stress waves were interacted with the boundary of the
⎡σ 11 ⎤ ⎡C11 C12 0 0 0 0 ⎤⎢ ⎥ Panel A. However, in the case of Panels B and C, damage propaga-
⎢σ ⎥ ⎢C 1
C 22 0 0 0 0 ⎥ ⎢ε 22 d
+ ε ⎥ tion was localized, and stress waves were not interacted with the
⎢ 22 ⎥ ⎢ 12 ⎥⎢ 3
vol

⎢ 0 ⎥ ⎢ 0 0 0 0 0 0 ⎥⎢ ⎥ boundary. Therefore, the lateral dimension of 150 mm2 was envis-
⎢ ⎥=⎢ ⎥⎢ d 1 ⎥ (14)
τ
⎢ 23 ⎥ ⎢ 0 0 0 αC 44 0 0 ⎥ ⎢ε 33 + ε vol ⎥ aged for the simulation which is sufficiently large to obtain the
3 solution without significant disquiets due to reflected waves. Fig. 7
⎢ τ 31 ⎥ ⎢ 0 0 0 0 αC 55 0 ⎥⎢ ⎥
⎢ ⎥ ⎢ ⎥⎢ γ 23 ⎥ shows the finite element (FE) model of the projectile and the target.
⎣ τ 12 ⎦ ⎣ 0 0 0 0 0 αC 66 ⎦ ⎢ γ 31 ⎥ The projectile and the target were analyzed using the Lagrange
⎢ ⎥ processor with a mesh of six-sided brick type (hexahedral) ele-
⎣ γ 12 ⎦
ments. To improve the accuracy of the analysis, smaller cells were
The combined effect of failure in all the three material direc- used near the impact zone, and the mesh size gradually increased
tions results in a material that can only withstand hydrostatic to the outer edges.
pressure, by a change in the material stiffness and strength to iso- A mesh sensitivity analysis was carried in order to ensure that
tropic with no stress deviators and no tensile material stresses [12]. the results obtained are not sensitive to the size of the elements
used. For this purpose, an FE model was modeled to predict the bal-
listic limit of Kevlar® laminate. The mesh sizes considered were
3.3. Erosion criteria
0.2 mm, 0.4 mm, 0.6 mm, 0.8 mm, and 1 mm. The ballistic limit was
estimated for all the mesh sizes of the laminate and compared. The
An erosion algorithm that removes highly distorted elements
differences of the ballistic limit between the mesh size of
when instantaneous geometric strain for erosion exceeds a defi-
0.2 × 0.2 mm, 0.4 × 0.4 mm, 0.6 × 0.6 mm, and 0.8 × 0.8 mm were 0.3%
nite value was implemented. The mass of removing cells is
(374 m/s), 0.06% (375.25 m/s), 0.13% (375.5 m/s), and 0.06% respec-
distributed equally to the remaining nodes. By this distribution, the
tively. Therefore, all the simulations were carried out using mesh
inertia and spatial continuity of inertia are conserved in the mesh.
size equal to or less than 0.8 mm. The projectile was discretized into
The specific value of erosion criteria is defined typically using a value
855 elements.
of 0.5–2 [34]. However, the ballistic impact response is not sensi-
The interaction between the projectile and the target was defined
tive to the variation of erosion strain [17]. Principal strain
using gap interaction logic [33]. With this logic, each segment is sur-
components (εi, εij, i = 1 to 3, j = 1 to 3) are used to calculate the ge-
rounded by a contact detection zone, and the radius of this zone
ometric strain using the following equation:
is defined as gap size. Any nodes entering into this detection zone
2 2 are repelled by a force about the depth of penetration of the nodes
[ (ε1 + ε 22 + ε 32 ) − (ε1ε 2 + ε 2ε 3 + ε 3ε1 ) + 3(ε122 + ε 232 + ε 312 ) ]
12
ε eff = (15) into the detection zone. The gap size used in the present simula-
3
tions was 0.0033 mm.
In the present study, the geometric strain for erosion was con- Numerical simulations were performed to predict the ballistic
sidered as 1.5 for Kevlar® composite targets, and it is similar to the limit velocity. Attempting to obtain this value, the panels were sub-
erosion strain considered in [11,12]. jected to an increasing impact velocity. Projectile velocity history
The above discussed material model was implemented through was recorded as shown in Figs. 15–17. The ballistic limit was mea-
hydrocode simulations to investigate the ballistic impact response sured based on the two standard methods available in the literature.
of armor panels to simulate failure modes and damage. The re- The first method was based on the velocity time history in which
sponse obtained from the simulations was compared with the ballistic limit (Vbl) is defined as the maximum impact velocity at
experimental data to evaluate the performance of the simula- which the projectile can be stopped [11,35]. The second method was
tions. Good correlation between the hydrocode simulations and according to US MIL-STD-662E [36] (V50). According to this method,
ballistic tests was succeeded in terms of damage patterns and failure the V50 can be calculated by taking the average of an equal number
modes. Further, simulations were extended to obtain an estimate of lowest velocities giving full perforations and highest velocities
for ballistic limit and to simulate different failure modes. Attempt- giving partial perforations which occur within a small velocity range
ing to obtain this value, projectile velocity time history was recorded of 38 m/s. The elastic constants obtained from the experimental char-
for various impact velocities. acterization are presented in Table 4.
A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13 7

Time (ms) 0 0.02 0.04 0.05 0.06 0.075 0.1

(a) Panel A: 2.5 x 75 mm2

13

Time (ms) 0 0.02 0.04 0.05 0.06 0.075 0.1


(b) Panel B: 2.5 x 100 mm2

Time (ms) 0 0.02 0.04 0.05 0.06 0.075 0.1

(c) Panel C: 2.5 x 150 mm2

Fig. 6. Damage propagation in panel with different dimensions.

The projectile was made of brass jacket with lead core. The shear projectile material model were adopted from the standard ANSYS
response of the brass jacket and the lead core was modeled using AUTODYN 14.0 material library [39].
Johnson–Cook [37] and the Steinberg–Guinan [38] strength models,
respectively. The high strain rate response of these materials was 5. Results and discussion
modeled using Mie-Gruneinsen EOS. The constants for the
The ballistic impact tests on Kevlar® reinforced thermoplastic-
based armor panels revealed a number of failure modes. The
response of the armor panels from the ballistic impact test was used
as a benchmark for later comparison with that obtained from
hydrocode simulation. Ballistic limits were also predicted using ve-
locity time histories and residual velocities obtained from hydrocode
simulations.

5.1. Damage patterns

(a) (b) Fig. 8 shows the damage in the form of a crucifix that evolves
with velocity according to recognizable patterns for the given impact
Fig. 7. FE model: (a) projectile and (b) target. velocity, and similar observations were reported in [11,12]. The extent
8 A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13

Table 4
Description of material model for Kevlar® laminates of different fabrics under study.

Material 2D-P 3D-O 3D-A

Equation of state Ortho Ortho Ortho


Density (g/cm3) 1.176 1.192 1.192
Young’s modulus 11 (kPa) 14.625e+06 17.204e+06 18.330e+06
Young’s modulus 22 (kPa) 14.625e+06 17.204e+06 18.330e+06
Young’s modulus 33 (kPa) 4.293e+05 5.847e+05 5.847e+05
Poissons ratio 12 0.048 0.025 0.09
Poissons ratio 23 0.182 0.590 0.390
Poissons ratio 31 0.182 0.590 0.390
Strength Elastic Elastic Elastic
Shear modulus (kPa) 6.977e+06 8.392e+06 8.408e+06
Failure Material stress/strain Material stress/strain Material stress/strain
Tensile failure strain 11 0.025 0.087 0.074
Tensile failure strain 22 0.025 0.087 0.074
Tensile failure strain 33 0.010 0.054 0.033
Post failure option Orthotropic Orthotropic Orthotropic
Failed in 11, failure mode 11 only 11 only 11 only
Failed in 22, failure mode 22 only 22 only 22 only
Failed in 33, failure mode 33 only 33 only 33 only
Failed in 12, failure mode 12 & 11 only 12 & 11 only 12 & 11 only
Failed in 23, failure mode 23 & 11 only 23 & 11 only 23 & 11 only
Failed in 31, failure mode 31 & 11 only 31 & 11 only 31 & 11 only
Residual shear stiffness fraction 0.2 0.2 0.2
Erosion Geometric strain Geometric strain Geometric strain
Erosion strain 1.5 1.5 1.5
Type of geometric Strain Instantaneous Instantaneous Instantaneous

of damage in the present study was lower compared to that of the it is found to be dominated by fiber stretching and shearing. The
reported results, indicating the influence of thermoplastic ogive nose of the projectile also showed an effect on the penetra-
matrix. tion through stretching of fibers adjacent to the nose. The damage
Among the three armor panels, the projectile did not pene- was in the form of a cone with open in the direction of impact pe-
trate the armor panels with 3D-O and 3D-A fabrics. However, full riphery. Cone formation was due to the compression of the target
perforations were observed in armor panels with 2D-P fabrics. The material and lead to the upflow of the material (Fig. 10).
impact tests on the 2D-P armor panels uncover some of the failure
modes such as delamination, shear plugging, matrix cracking, in- 5.2. Comparison of test and numerical results
plane failure, and fiber crushing. In these laminates, experimental
results at velocities between 350 m/s and 376 m/s showed partial Damage patterns observed in the ballistic impact test showed
penetration of the projectile, i.e. remained inside the armor. Beyond different failure modes such as matrix cracking, delamination, fiber
this velocity, all the shots showed full perforation. failure, fiber crushing, and shear plugging. The three K-MPP lami-
From hydrocode simulations, different damage patterns were ob- nates exhibited different failure modes depending on their fabric
served in the armor panels. These damage patterns represent architecture. It should be noted that the ballistic impact param-
different failure modes, which occur in a composite laminate under eters, such as projectile mass and the projectile geometry, were kept
the ballistic impact. Propagation of the damage in the impact test the same for all the armor panels.
was not captured due to the limitations in the impact test. However, Fig. 11 shows the damage pattern observed in the 2D-P armor
through simulations, it was captured at different cycles as shown panel for an impact velocity of 426 m/s. Complete perforation of the
in Fig. 9. target was seen in hydrocode simulation as well as in the ballistic
When a projectile impacts the target, a compressive wave gen- impact test with shear plugging failure. From the figure, matrix,
erates through-the-thickness of the target and gets reflected as a cracking and delamination can be observed. The cracks were ob-
tensile wave from its inner surface causing delamination between served along the warp and weft direction on the top surface, and
the plies. As the projectile penetrates, the deformation in the target shear plugging at the bottom surface.
changes. During the initial stage of penetration, compression and Fig. 12 shows the failure near impact zone and shear plugging
displacement of the target material are dominant. At later stages, at the back of the target (major failure mode). The shear plug area

Fig. 8. Damage patterns of the composite target.


A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13 9

Fig. 9. Damage propagation through simulations (400 m/s).

from the impact test exhibited a square shape area of 103.67 mm2, Fig. 13 shows the damage patterns observed in an armor panel
and from simulation, similar profile was obtained with an area of with 3D-O fabrics when impacted with an impact velocity of 426 m/s.
112.16 mm2. Fiber crush near the impact zone was also one of the In the vicinity of the impact zone, diffused damage was observed.
major failure modes that was observed during the impact tests and The failure modes observed were delamination, fiber crush, matrix
hydrocode simulations. cracking, and fiber failure. At this impact velocity, the projectile was

Fig. 10. Stretching of fibers and upflow of the material.

Fig. 11. Damage patterns observed in both experimental and hydrocode simulations (2D-P laminate).

Front face Back face

(a) Impact zone (b) Shear plugging at the back of the target

Fig. 12. Failure observed in the 2D-P laminate.


10 A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13

Table 5
Damage area at front and back surfaces (v = 426 m/s).

Type of Surface Impact Simulation Difference


laminate test (mm2) (mm2) (%)

2D-P Front 71.25 78.16 9.69


Back 103.67 112.16 8.19
3D-O Front 69.52 76.59 9.23
Back - - -
3D-A Front 66.04 70.88 7.32
Back - - -

Fig. 13. Damage patterns in 3D-O laminate. It can be observed that the extent of damage in the vicinity of
impact in 2D-P panels was 2.4% higher than that of the 3D-O panels
and 7.3% higher than that of the 3D-A panels. Among the three armor
struck in the target showing partial penetration with negligible panels, 3D-A panel showed better impact resistance than that of
backface signature. the remaining two panels, which can be further improved to study
Fig. 14 shows the damage that occurred in 3D-A fabrics rein- the behavior at various impact velocities with different thickness
forced armor panel when impacted with an impact velocity of of the target and effect of different projectiles.
426 m/s, and partial penetration was observed. Damage patterns ob-
tained from hydrocode simulations were in good correlation with 5.3. Ballistic limit (Vbl/V50)
the test results, which shows the capability of the simulations. Due
to the higher interlaminar strength of the 3D-A fabric structure, the Fig. 15 shows the velocity time histories of the projectile im-
delamination observed was very low in 3D-A armor panels com- pacting the 2D-P laminate. From Fig. 15, it can be observed that the
pared to the remaining armor panels. laminate with 2D-P fabrics decelerates the projectile slowly due to
Based on the above comparison of the damage patterns for dif- the presence of crimp. Therefore, the warp and fill yarns have to
ferent armor panels, the following inferences can be explained first decrimp before they can start to elongate in tension, thereby
relating to the fabric architecture: slowing down the projectile.
From Fig. 15a it can be observed that for an impact velocity of
– The warp and weft yarns in the 2D-P laminates are undulated 380 m/s and 400 m/s, the residual velocities observed were 68 m/s
because of the interlacing with each other. Moreover, in 2D-P and 131 m/s, respectively. However, for an impact velocity of 376 m/s,
laminates due to the presence of crimp, the factors that harm the residual velocity approaches to zero and subsequently nega-
the impact performance are: (a) more time for yarns to decrimp tive. At this instant, the target absorbed most of the impact energy
before they develop tension through elongation, (b) lower areal of the projectile and thereby resisted its penetration. Hence, Vbl of
density, and (c) greater weaving damages. 2D-P laminate obtained from numerical simulations was 376 m/s.
– When the projectile impacts the target, the stress waves spread Fig. 15b shows the estimation of the V50 as 377.5 m/s by taking the
in in-plane fiber direction and through-the-thickness direc- average of six initial impact velocities within the range of 38 m/s.
tion. z-yarns in the 3D panels reinforce the system in the thickness Fig. 16 shows the velocity time histories of the projectile im-
direction, thus effectively preventing the occurrence of pacting 3D-O laminate at various impact velocities. By analyzing
delamination. Fig. 16a, one can observe that, for an impact velocity of 400 m/s, the
– In 3D fabric structure, the z-yarns play a vital role in holding the residual velocity was negative, indicating rebounding of the projectile.
weft and warp yarns all together. Due to the presence of z-yarns, For an impact velocity of 450 m/s, the residual velocity ap-
areal density increases. Due to the increased areal density, warp proaches zero, indicating that the target absorbed most of the kinetic
and weft yarns contribute high in-plane stiffness and strength. energy of the projectile. This is the maximum velocity at which the
As the areal density is high, initial momentum transfer between target stopped the projectile, and for a velocity beyond this limit,
the projectile and the target occurs up to a great extent and in- full perforations were observed. Therefore, it can be concluded that
creases the specific energy absorption. the Vbl of a 3D-O laminate was 450 m/s. According to US military
standard, the V50 obtained was 452.5 m/s (Fig. 16b).
From the impact tests and hydrocode simulations, damage area Velocity time histories of the projectile impacting 3D-A lami-
observed on front and back surfaces was consistent. Table 5 pres- nate are presented in Fig. 17. During the ballistic impact test and
ents the experimentally and numerically obtained damage areas for simulations, 3D-A laminate showed partial penetration for an impact
the same mass and geometry of the projectile. velocity of 426 m/s. For an impact velocity of 500 m/s, projectile per-
forated carrying a residual velocity of 91 m/s. It indicates that the
ballistic limit is between 426 m/s and 500 m/s. Partial penetra-
tions were observed up to an impact velocity of 470 m/s above which
complete perforations were observed. Thus, the Vbl of 3D-A lami-
nate was 470 m/s. Fig. 17b shows the V50 of 471.2 m/s measured
according to US military standard.
An interesting observation in the velocity history is the decel-
eration of the projectile. Initially, the projectile decelerates faster
in the case of laminates reinforced with 3D fabrics (Figs. 16a and
17a) than that of the 2D-P laminates (Fig. 15a). This is due to the
absence of crimp in 3D fabrics and locking of yarns in their posi-
tion of the matrix. Hence, the deformations, energy absorptions, and
momentum transfers are spread out further, decreasing the occur-
Fig. 14. Damage patterns in 3D-A laminate. rence of damage. However, when z-yarns near the impact zone begin
A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13 11

500 387
2D plain woven, Impact Velocity (m/s) 2D plain woven
375 , 376, 380, 400 Full perforation Partial perforation
384
400
381

300 Ballistic limit = 377.5 m/s

Velocity (m/s)
378
Velocity (m/s)

375
200
372

100
369

366
0
0.000 0.025 0.050 0.075 0.100 0.125 0.150 0 1 2 3 4 5 6 7
Time (ms) Impact number

(a) (b)

Fig. 15. Ballistic limit velocity for 2D-P laminates: (a) velocity time histories, (b) US standard.

to fail, the fabric structure becomes compromised, with little left higher than that of the 2D-P laminates and 4.2% higher than that
holding the warp and weft layers together. Thus, the velocity history of the 3D-O laminates. However, the percentage variation between
rapidly levels off, indicating complete fabric penetration. During the Vbl and V50 was up to 0.55%. The hierarchy of the ballistic limit ob-
time between which the warp and weft yarns at the impact zone tained is 3D-A > 3D-O > 2D-P.
have failed, the friction between the projectile and the target is the The variation of residual velocities with impact velocities for the
dominant mechanism by which the projectile velocity further three laminates predicted from simulations is shown in Fig. 18. For
reduces. A gradual decrease in the slope of the velocity history before 2D-P laminate, as impact velocity increases beyond the ballistic limit,
it completely levels off indicates a greater level of friction. This was an increase in residual velocity was observed with high rate.
observed to a greater extent in 2D-P laminates and to a lesser extent Whereas, in the case of 3D-O and 3D-A laminates, the increase was
in 3D laminates. not that steep. This is due to the architecture of the fabric that slows
The effect of fabric architecture was investigated by comparing down the projectile for further penetration.
the ballistic limit velocity. From Table 6, it can be observed that the Table 7 presents the ballistic impact response of the K-MPP lami-
ballistic limit velocity Vbl of 3D-A laminates was found to be 20% nates reinforced with fabrics of different architecture. By analyzing

3D orthogonal
525 480
Full perforation Partial perforation
3D orthogonal, Impact Velocity (m/s)
450
400, 450, 455 , 500
470

375
460
Ballistic limit = 452.5 m/s
Velocity (m/s)
Velocity (m/s)

300
450
225

440
150

75 430

0
420
0.000 0.025 0.050 0.075 0.100 0 1 2 3 4 5 6 7
Time (ms) Impact number

(a) (b)

Fig. 16. Ballistic limit velocity for 3D-O laminates: (a) velocity time histories, (b) US standard.
12 A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13

600 485
3D angle interlock, Impact Velocity (m/s) 3D angle interlock
Full perforation Partial perforation
450, 460, 470, 480, 500
480

400 475
Velocity (m/s)

Velocity (m/s)
Ballistic limit = 471.2 m/s
470

200
465

460

0
455
0.000 0.025 0.050 0.075 0.100
0 1 2 3 4 5 6 7
Time (ms)
Impact number

(a) (b)

Fig. 17. Ballistic limit velocity for 3D-A laminates: (a) velocity time histories, (b) US standard.

the results presented in Table 7, the effect of fabric structure on the between 2D-P laminate and 3D-A laminate was 36%. By compar-
ballistic impact response of the K-MPP laminates was studied in ing the ballistic response parameters between 3D-O laminates and
terms of residual velocity, energy absorption, and ballistic limit. For 3D-A laminates, it can be observed that, for the same thickness of
the same mass and geometry of the NIJ-STD-0101.06 Type IIIA pro- the target and same projectile, the energy absorbed at ballistic limit
jectile, fabric structure plays a significant role in the ballistic impact by 3D-A laminate was 8.33% higher than that of the 3D-O laminate.
behavior of K-MPP laminates.
The increase in the ballistic limit from 2D-P laminates to 3D-A 6. Conclusions
laminates leads to an increase in the energy absorption. It was found
that the maximum difference of energy absorbed at the ballistic limit Three types of composite armor panels were manufactured using
between 2D-P laminate and the 3D-O laminate was 30.18%, and PP matrix and Kevlar® 29 fabrics with architectures of 2D plain
woven (2D-P), 3D orthogonal (3D-O), and 3D angle interlock (3D-
A). The adhesion between the Kevlar® fabric and PP was improved
Table 6 by adding a coupling agent called maleic anhydride grafted (MAg)-
Predictions of ballistic limit.
PP. The post-impact damage patterns of NIJ-STD-0106.01 Type IIIA
Type of Ballistic limit composite armor were obtained from ballistic impact tests. The
Laminate velocity (m/s)

Vbl V50 Difference (%)

2D-P 376 377.5 0.39 Table 7


3D-O 450 452.5 0.55 Ballistic impact performance of 2D-P, 3D-O and 3D-A laminates for NIJ-STD-
3D-A 470 471.2 0.25 0101.06 Type IIIA.

Type of Vi (m/s) Vr (m/s) Type of KEi (J) KEf (J) Ea (J)


laminate penetration

240 2D-P 350 −25.28 PP 490.00 - -


376 −2.32 PP 565.50 - -
2D-P, 3D-O, 3D-A 400 131 FP 640.00 68.64 571.35
200 426 157 FP 725.90 98.59 627.30
450 198 FP 810.00 156.81 653.18
160 500 240 FP 1000.00 230.4 769.60
Residual Velocity (m/s)

3D-O 400 −28 PP 640.00 - -


426 −11.24 PP 725.90 - -
120 450 −3.2 PP 810.00 - -
480 55.88 FP 921.60 12.49 909.11
80 490 62.74 FP 960.40 15.74 944.65
500 74.57 FP 1000.00 22.24 977.75
3D-A 400 −31.72 PP 640.00 - -
40 426 −24.74 PP 725.90 - -
450 −14.19 PP 810.00 - -
0 470 −0.126 PP 883.60 - -
480 40.47 FP 921.60 6.55 915.05
360 390 420 450 480 510 490 53.27 FP 960.40 11.35 949.05
-40 Impact Velocity (m/s) 500 61.27 FP 1000.00 15.01 984.98

Vi = impact velocity, Vr = residual velocity, PP = partial perforation, FP = full perfora-


Fig. 18. Impact velocity-post impact residual velocity. tion, KEi = initial kinetic energy, KEf = final kinetic energy, Ea = energy absorbed.
A.K. Bandaru et al./International Journal of Impact Engineering 89 (2016) 1–13 13

application of hydrocode simulations in assessing the ballistic re- [7] Naik NK, Doshi AV. Ballistic impact behavior of thick composites: analytical
formulation. AIAA J 2005;43(7):1525–36.
sistance of 3D armor panels was presented. The material modeling
[8] Naik NK, Shrirao P. Composite structures under ballistic impact. Compos Struct
was defined through a material model that couples the non-linear 2004;66:579–90.
anisotropic constitutive equations to the equation of state (EOS). As [9] Bandaru AK, Ahmad S. Numerical simulation of progressive damage of laminated
a summary of this work, the following conclusions were drawn: composites under ballistic impact. Safety, Reliability, Risk and Life-Cycle
Performance of Structures and Infrastructures. Taylor & Francis Group; 2013.
[10] Yen CF. A ballistic material model for continuous-fiber reinforced composites.
– The thermoplastic Kevlar® composite laminates yielded a 16– Int J Impact Eng 2012;46:11–22.
29% reduction in the density over conventional thermoset- [11] Silva MAG, Cismasiu C, Chiorean CG. Numerical simulation of ballistic impact
on composite laminates. Int J Impact Eng 2005;31:289–306.
based laminates, while maintaining an equivalent protection level. [12] Tham CY, Tan VBC, Lee HP. Ballistic impact of a KEVLAR® helmet: experiment
– Post-impact damage patterns from the ballistic impact tests re- and simulations. Int J Impact Eng 2008;35(5):304–18.
vealed that the 3D Kevlar® armor panels were able to confront [13] Grujicic M, Glomski PS, He T, Arakere G, Bell WC, Cheeseman B. Material
modelling and ballistic-resistance analysis of armor-grade composites reinforced
the 9 mm FMJ projectile unlike 2D Kevlar® armor panels. with high-performance fibers. J Mater Eng Perform 2009;18:1169–82.
– The existence of z-yarns in 3D fabrics greatly improved the bal- [14] Bandaru AK, Vetiyatil L, Ahmad S. The effect of hybridization on the ballistic
listic performance by engaging failed layers that would otherwise impact behavior of hybrid composite armors. Compos Part B 2015;76:300–
19.
separate from the rest of the layers. Due to the absence of crimp [15] Muhi RJ, Najim F, de Moura MFSF. The effect of hybridization on the GFRP
in 3D fabrics, their strength retentions were much higher than behavior under high velocity impact. Compos Part B 2009;40:798–803.
that of the 2D-P fabric. [16] Grujicic M, Pandurangan B, Koudela KL, Cheeseman B. A computational analysis
of the ballistic performance of light weight hybrid composite armors. Appl Surf
– The global damage and failure modes from hydrocode simula-
Sci 2006;253:730–45.
tions were in good agreement with ballistic impact test results. [17] Tan P. Ballistic protection performance of curved armor systems with or without
For the same mass and geometry of the projectile, the extent of debondings/delaminations. Mater Des 2014;64:25–34.
damage on the front surface of the 2D-P laminate was 2.4% higher [18] Vieille B, Casado VM, Bouvet C. About the impact behavior of woven-ply carbon
fiber-reinforced thermoplastic- and thermosetting-composites: a comparative
than that of the 3D-O laminates and 7.3% higher than that of the study. Compos Struct 2013;101:9–21.
3D-A laminates. [19] Kulkarni SG, Gao XL, Horner SE, Zheng JQ, David NV. Ballistic helmets – their
– The ballistic limit velocity was increased with the change in the design, materials, and performance against traumatic brain injury. Compos
Struct 2013;101:313–31.
fabric architecture from 2D-P to 3D-O and 3D-A. The ballistic limit [20] Walsh SM, Scott BR, Spagnuolo DM. The development of a hybrid thermoplastic
velocity of 3D-A laminates was 20% higher than that of the 2D-P ballistic material with applications to helmets. U.S. Army Research Laboratory,
laminates and 4.2% higher than that of the 3D-O laminates. Aberdeen Proving Ground, MD. ARL-TR-3700; 2005. 19 p.
[21] Walsh SM, Scott BR, Spagnuolo DM, Wolbert JP. Hybridized thermoplastic
– For same mass and geometry of the projectile, the hierarchy of aramids: enabling material technology for future force headgear. Aberdeen
overall ballistic performance of NIJ-STD 0106.01 Type IIIA armor Proving Ground, MD: U.S. Army Research Laboratory; 2006. 8 p.
made of Kevlar®/PP laminates consisting of different fabric ar- [22] Song JW. Thermoplastic composite for ballistic application [Ph.D. dissertation].
University of Massachusetts at Lowell; 1986. 118 p.
chitectures is as follows: [23] Carrillo JG, Gamboa RA, Flores-Johnson EA, Gonzalez-Chi PI. Ballistic
3D-A > 3D-O > 2D-P. performance of thermoplastic composite laminates made from aramid woven
– The present study emphasized the use of thermoplastic-based fabric and polypropylene matrix. Polym Test 2012;31(4):512–19.
[24] Shockey DA. Improved barriers to turbine engine fragments: interim report II.
composite laminates with different fabric architectures and design
Report no: DOT/FAA/AR-99/8, II; 1999.
prospects that could be used to engineer a lightweight body [25] Choi CH, Ok YS, Kim BK, Ha CS, Cho WJ, Shin YJ. Melt rheology and property
armor that meets the requirements prescribed in test stan- of short aramid fiber reinforced polyethylene composites. J Korean Ind Eng Chem
dards from government agencies. 1992;3(1):81–7.
[26] Petrie EM. Hand book of adhesives and sealants. McGraw-Hill; 2000.
[27] National Institute of Justice-USA. Ballistic resistance of body armor. (NIJ)
Acknowledgement Standard 0101.06; 2008.
[28] Gower HL, Cronin DS, Plumtree A. Ballistic impact response of laminated
composite panels. Int J Impact Eng 2008;35(9):1000–8.
The authors acknowledge the dedicated efforts of Col. Anil Yadav, [29] ASTM D7136. Standard test method for measuring the damage resistance of a
Army field camp, Kapurthala, Punjab, India, for conducting a field fiber-reinforced polymer matrix composite to a drop-weight impact event, 2007.
[30] ASTM D 3039. Standard test method for tensile properties of polymer matrix
test on the laminates under study. composite materials, 2000.
[31] Hiermaier S, Riedel W, Hayhurst CJ, Clegg RA, Wentzel CM. Advanced Material
References Models for Hypervelocity Impact Simulations, EMI-Report No. E43/99, ESA CR(P)
4305; 1999.
[32] Hayhurst CJ, Hiermaier SJ, Clegg RA, et al. Development of material models for
[1] Behera B, Hari P. Woven textile structure-theory and applications. Woodhead Nextel and Kevlar-Epoxy for high pressures and strain rates. Int J Impact Eng
publishing series in textiles. CRC Press; 2010. 1999;23:365–76.
[2] Pan N, Lin Y, Wang X, Postle R. An oblique fiber bundle test and analysis. Text [33] AUTODYN® composite modelling, Revision 1.3, ANSYS Inc.; 2011.
Res J 2000;70(8):671–4. [34] AUTODYN Theory Manual, Revision 4.3, Century Dynamics Inc.; 2005.
[3] Yunqiao R, Richard FJ. A modeling and experimental study of the influence of [35] Twaron in Soft body armor. Wuppertal: AKZO; 1993.
twist on the mechanical properties of high-performance fiber yarns. J Applied [36] Military Standard V50 ballistic test for armor. MIL-STD-662E; 1987.
Polym Sci 2000;77(9):1938–49. [37] Johnson GR, Cook WH. A constitutive model and data for metals subjected to
[4] Briscoe BJ, Motamedi M. The ballistic impact characteristics of aramid fabrics: large strains, high strain rates and high temperatures. In: Proceedings of the
the influence of interface friction. Wear 1992;158:229–47. seventh international symposium on ballistics, Hague, Netherlands; 1983.
[5] Cheeseman BA, Bogetti TA. Ballistic impact into fabric and compliant composite [38] Steinberg DJ. Equation of state and strength properties of selected materials.
laminates. Compos Struct 2003;61:161–73. UCRL-MA-106439. Livermore, CA: Lawrence Livermore National Laboratory;
[6] Cheng WL, Scott L, Shigeru I. High velocity impact of thick composites. Int J 1991.
Impact Eng 2003;29(1–10):167–84. [39] ANSYS AUTODYN. ANSYS Workbench Release 14.0; 2011.

Anda mungkin juga menyukai