You are on page 1of 14

International Journal of Impact Engineering 50 (2012) 99e112

Contents lists available at SciVerse ScienceDirect

International Journal of Impact Engineering


journal homepage: www.elsevier.com/locate/ijimpeng

Performance of an advanced combat helmet with different interior cushioning


systems in ballistic impact: Experiments and nite element simulations
Long Bin Tan, Kwong Ming Tse*, Heow Pueh Lee*, Vincent Beng Chye Tan, Siak Piang Lim
Department of Mechanical Engineering, National University of Singapore, Singapore

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 23 November 2011
Received in revised form
21 June 2012
Accepted 25 June 2012
Available online 3 July 2012

This paper presents results from both experiments and numerical simulations of frontal and lateral
ballistic impacts on a Hybrid III headform equipped with Advanced Combat Helmets (ACH) of two
different interior cushioning designs, namely the strap-netting system and the Oregon Aero (OA) foam
padding. It aims to study the differences between two different impact orientations for these two
different interior cushioning systems. The experiments involve frontal and lateral ballistic impacts of
a 11.9 g spherical steel projectile traveling at speeds of approximately 200 m/s, striking on the helmetcushion-headform assemblies. The dynamic interaction between various components of the helmetcushion-headform assemblies are investigated with the use of high-speed photography while posttest evaluation of the damaged helmets are performed using visual observation, optical microscopy
and computed tomography (CT) scan. A series of ballistic impact simulations with nite element (FE)
models of the two assemblies reconstructed from CT images, are performed to correlate with experimental results. The commercial software, Abaqus, is used for the FE analyses. In general, there is
reasonable correlation between numerical and experimental observations and on quantitative parameters, such as head accelerations, helmet damage and deections. It is also found that softer foams with
low stiffness are more effective as shock absorbing cushion against ballistic impacts under certain
condition. Additionally, results of the two interior cushioning systems are compared with various injury
criteria to assess their acceleration levels. It is found that, for frontal impact, the helmet with strapnetting system fails both the Wayne State Tolerance Curve (WSTC) and Federal Motor Vehicles Safety
Standards (FMVSS) 218 criteria while the one with OA foam-padding passes both.
2012 Elsevier Ltd. All rights reserved.

Keywords:
Ballistic impact
KEVLAR helmet
Oregon Aero (OA) interior foam padding
Strap-netting interior cushion
Hybrid III headform

1. Introduction
Given the recent rise in terrorism, civil and international
conicts, the number of people aficted with war-related traumatic
head injuries is set to increase. In fact, most of these war-related
head injuries are due to blunt and penetrating ballistic impact.
Based on studies by Mathers et al. (2006) [1] compiled from data
gathered in 2001, ballistic head injuries account for approximately
a quarter of the violent deaths occurring annually worldwide.
Although the ofcial statistical data of the war-related head injury
deaths are fewer as compared to that caused by road trafc accidents, these numbers are often regarded as being underestimated
due to national security reasons [1]. Since most of these war-related
head injuries are due to blunt and penetrating ballistic impacts, in
order to minimize the morbidity and mortality resulting from
* Corresponding authors.
E-mail addresses: tsekm.research@yahoo.com, g0900558@nus.edu.sg (K.M. Tse),
mpeleehp@nus.edu.sg (H.P. Lee).
0734-743X/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijimpeng.2012.06.003

ballistic head injuries, it is particularly important to ensure that


personal protective equipments, especially the combat helmets, are
capable and effective in providing protection from ballistic impact.
The function of combat helmets is to attenuate shock by
absorbing the energy generated during impact through deformation and dissipating the energy rapidly via delamination and failure
of its shell without injuring the wearer. Despite the fact that
modern combat helmets can prevent bullets of handguns and even
some ries from penetrating them, traumatic injuries to both the
skull and brain can still occur due to the excessive mechanical
responses of the helmet and the head. A ballistic impact on
a helmet can generate peak head acceleration which may exceed
the tolerance level of what the tissues can bear, thus causing irreversible damage to these tissues. Injuries to head can also occur
when the bullet has sufcient energy to cause the interior helmet
shell to come in contact with the underlying tissue and this is
known as rear effect [2].
Earlier research in all domains of head protection could be
traced back to the 60 s when experiments, as reported in [3e6] had

100

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

been performed extensively to investigate on the required head


protection in an event of traumatic head injury. Due to ethical
issues, these tests could only use simplied experimental representation of the human head, for example, spherical metal shells
with lled liquid. Subsequently, biodelic headforms such as Side
Impact Dummies (SIDs) and Hybrids, were used for head injury
prediction [7]. However, different and non-standardized test
procedures used by helmet manufacturers made comparison of
helmets difcult. Moreover, maintenance of experimental equipments and replacement of damaged components incur costs.
Numerical simulations, which serve as cost-effective alternative to
these experimental tests, were then used to understand the
mechanism of head injury during impact and how the use of
helmets attenuates injuries. It was until the 70 s when the rst few
analytical studies [8,9] used a simplied axisymmetric elastic headhelmet nite element (FE) model to investigate the dynamic
response of head-helmet in a localized short-duration impact. A
year later, Khalil et al. [10] conducted a low-velocity ballistic impact
study using a simplied axisymmetric elastic head-helmet FE
model and validated with corresponding experiments. Thereafter,
tremendous efforts had been spent on research of head protective
helmets using nite element method (FEM) [11e17]. Van Hoof et al.
[11,12] performed both experimental and numerical studies on the
response of the woven composite helmet materials to ballistic
impact, and found that the helmet interior exhibited large deformation which could exceed the gap between inner helmet shell and
head [12]. Baumgartner and Willinger [13] simulated high-speed
ballistic impact on aluminum military helmet with their FE head
model which consisted of basic anatomical components. In their
simulations, it appeared that the tissue tolerance limit for
sustaining traumatic brain injury was not reached, despite of the
prediction of skull fracture. Aare and Kleiven [14] studied the
effects of helmet shell stiffness and impact angles direction on the
load levels in human head during a ballistic impact, using a more
detailed FE model of human head. They concluded that the helmet
shell deections should not exceed the initial gap between the

helmet shell and the head in order to prevent the rear effect. Tham
et al. [15] conducted ballistic tests to determine the response of the
KEVLAR helmet, which was then used as a benchmark for
comparison against FE simulation results. It was found that the
KEVLAR helmet was capable of deecting the full-jacketed bullet
traveling at 358 m/s. However, no head models were incorporated
in Thams simulation and hence the analyses were restricted to
mechanical responses of the helmet without any insights to head or
brain injuries. Another study by Yang and Dai [17] focused on
evaluation of the rear effect by having the helmeted FE head model
impacted by bullet at different impact angles. The severity of head
responses was then assessed using head injury criterion (HIC).
Nonetheless, controversy arises over numerical validations of
these studies against experimental data since the numerical and
experimental head model used was non-identical. Most of these
numerical studies used FE head models with more realistic human
anatomical features. On the other hand, their validations were
against data obtained from ballistic experiments on biodelic
headforms which are completely different from the FE head
models. Researchers who used different head models for their
numerical and experimental work will have to contend with
accuracies of transition of impact forces from the helmet to the
head between their two models, which further complicates the
analyses. Therefore, it would be more appropriate to adopt an
identical head model for both numerical and experimental tests.
In the present study, frontal and lateral ballistic impact tests are
conducted to determine the responses of the ACH helmet and the
Hybrid III headform in terms of deformation and acceleration.
These experimental responses are used as a benchmark for the
validation of numerical solutions coming from the same FE model
assembly of ACH helmet, interior cushioning system and Hybrid III
headform. The validation of the head responses from FEA further
allows the verication of correct contact or interaction properties
applied between the backplane of the helmet shell interior and of
the cushioning system against the head. Of late, focus has also been
paid to the helmet shell materials [18e24] and helmet liners or

Fig. 1. Advanced combat helmet (ACH) with (A) strap-netting (Helmet 1) and (B) Oregon Aero (OA) interior foam cushioning system (Helmet 2). Experimental setup and impact
sites for (C) frontal and (D) lateral ballistic impact test on Hybrid III head and neck.

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

101

Fig. 2. Experimental setup and measurement devices of the ballistic impact test. (A) Accelerometers mounted within the Hybrid III headform. (B) Strain gauge mounted on helmet
interior surface of the helmet shell. (C) A 3 m long gun barrel of the ballistic gas gun and steel spherical projectile (in the insert). (D) Close-up of the launcher unit of the gas gun
connected to the gas cylinder and gun barrel. (E) Measuring instruments at the target chamber. (F) High-speed camera.

cushion system [25,26]. This present study also aims to compare


the effectiveness of different interior cushioning systems, namely
the strap-netting type and foam padding type. Moreover, post-test
failure observations through the uses of optical microscopy and
computed tomography (CT) scan allow the additional comparison
with the FEA predictions of the damage to the helmet. Lastly, results
from the helmeted headform with two different interior cushioning
systems are also compared with various injury criteria.
2. Methods and materials
2.1. Experimental procedure
The present work deals with the evaluations of frontal and lateral
ballistic impacts on the military Advanced Combat Helmet (ACH)
with strap-netting system (Helmet 1) and Oregon Aero (OA) interior
foam cushioning system (Helmet 2), using the Hybrid III headform
(Fig. 1). In order to obtain the acceleration readings within the
headform, a spherical steel projectile of mass of 11.9 g is launched
from a ballistic gas gun to strike the front and side of the helmet at

Fig. 3. Modeling methodology for the ACH helmet-interior cushion-Hybrid III headform assembly.

102

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

Fig. 4. (A) The FE model of ACH helmet-interior cushion-Hybrid III headform assembly, generated from CT data by Mimics. (B) Sectional and lateral view of the helmet-interior
strap-netting-headform assembly (Helmet 1). (C) Meshed models of the helmet, headform, strap-netting system and OA foam padding system. (D) Sectional and lateral view of
the helmet-interior OA foam padding-headform assembly (Helmet 2).
Table 1
Number of elements and types of elements in the two FE helmet-cushion-headform models.
Components

Projectile

Helmet

No. of elements
Element type

2560
Linear hexahedral

26,296

Total no. of elements

1,91,703 (For Case 1); 2,89,903 (For Case 2)

respective impact velocities of 205 m/s and 220 m/s, which constitute a kinetic energy of about 250 J for the impacting projectile. The
resulting accelerations are then correlated to numerical results to
aid understanding of the effect of interior cushion properties in
attenuating blunt impact responses. Head accelerations during the
ballistic impacts, are recorded by an accelerometer which is
mounted within the headform (center of gravity) (Fig. 2A) while
strain gauges are attached on the inner surface of the helmet to
monitor the strain magnitudes and exural frequencies experienced
during impact (Fig. 2B). The bi-axial strain gauges are oriented at
approximately 45 from the front and side of the helmet. Such
conguration not only prevents the sensors from being affected
under direct impact, but also measures the hoop and radial strains of

Helmet interior cushioning

Hybrid III headform

OA foam
(for case 1)

Strap-netting
(for case 2)

Outer
skin

Inner
core (skull)

7428
Linear
hexahedral

1,05,628
Linear
hexahedral

62,616
Linear tetrahedral

92,803

Fig. 5. Resulting material orientation dened for the helmet model.

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

103

Table 2
Mechanical properties of advanced combat helmet (ACH).
E11 (MPa)

E22 (MPa)

E33 (MPa)

y12

y13/y32

G12 (MPa)

G23/G13 (MPa)

r (kg/m3)

Additional description

18,000

18,000

4500

0.25

0.33

770

2600

1230

1. With modeling of inter-ply delamination of


Helmet (4 layers of elements) through thickness.
2. Modeled degradation of helmet properties
with failure initiation criteria.

Table 3
Failure properties of advanced combat helmet (ACH).
X1t (MPa) X1c (MPa) X2t (MPa) X2c (MPa) X3t (MPa) X3c (MPa) S12 (MPa) S13 (MPa) S23 (MPa) Sn (MPa)
555

555

555

555

1050

1050

77

the helmet directly. Our experiment uses a ballistic gas gun, which
primarily comprises of the launcher unit, gun barrel and target
chamber (Fig. 2C and D). The launcher unit is a high-pressure air
chamber connected to a gas cylinder, a pressure transducer/indicator and the gun barrel (Fig. 2D). In order to propel the projectile
into the gun barrel, pressure valves are used to lock in the desired gas
pressure and for instantaneous release of the gas pressure. Each steel
spherical projectile, which weighs 11.9 g and of 14.2 mm diameter
(See the insert of Fig. 2C), is placed at the front end of the gun barrel
and locked before the launcher unit is screwed tight to the barrel.
Light Emitting Diodes (LEDs) are used to measure the projectile
speed as it passes through the LED channel of known length (Fig. 2E).
The time the projectile takes to hit the two LEDs are recorded by an
oscilloscope whereby the projectile speed can be subsequently
calculated from the known distance traveled and time taken
(Fig. 2E). After checking that all the measurement sensors are ready
and that the gas pressure has stabilized, the hand lever is quickly
depressed to release the gas pressure instantaneously to propel the
projectile into the gun barrel. The straight gun barrel keeps the
propelled projectile moving in the direction of the target chamber.
The target chamber allows the test specimen to be rigged up via
clamping on a steel platform.
The entire impact sequence is recorded using high-speed
camera photography for further analysis of the dynamic behavior
of the helmet and foam. The high speed camera is set at 30,000
frames per second, 1/90,000 shutter speed, and yields a screen
resolution of 512 x 512 pixels (Fig. 2F). A very bright halogen lamp is
used to illuminate the specimen since at high frame rate captures,
the amount of light entering the lens will be reduced to result in
dimmer images.

1060

1086

Ss (MPa)

Additional
description

823.2
588
1. Models for intralaminar failure
(4 layers) (4 layers)
using Hashin-Fabric Criteria.
2. Models for interlaminar failure
using surface traction criteria.
3. Model for property degradation.

associated with woven and laminated fabric shells. The damaged


helmets are also sent for CT scans to provide a glimpse of the
internal structures of the helmet so as to investigate its integrity
after impact test. The CT scan images are then used to validate the
failure mechanisms observed in optical microscopy, and helps
quantify the spatial extent of ber breakage and inter-ply delamination. The technique aids in visualizing and determining
measurement of the deformed region for purposes of failure characterization or for validation of FE models.
2.3. Finite element models
The methodology used for model creation of the helmet and the
Hybrid III headform is given in the owchart in Fig. 3. Due to the
complexity in geometry for various parts of the combat helmet and
the Hybrid III headform, the models are not drawn from scratch.
Instead, the geometrical information of the two helmet-headform
assemblies with different interior cushioning systems was scanned by the Siemens Sonotron Sensation CT scanner and is extracted
as axial images (in-plane resolution of 512 x 512 pixels with a pixel
size of 0.68 m and slice thickness of 0.6 mm). These CT images are
then imported into Mimics v13.0 (Materialise, Leuven, Belgium) e
a medical image processing software, for segmentation (separation
of various components such as the helmet, headform, foam or
straps netting) and reconstruction of the two assembled models
(Fig. 4). A semi-automatic meshing technique is employed in
HyperMesh v10.0 (Altair HyperWorks, Troy, MI, USA) to optimize
between computational efciency and element quality, with
average element size of 4 mm and aspect ratio of 1.64. 4-noded

2.2. Post-test observations


After the ballistic tests, the damaged helmets are scrutinized
using the optical microscope to observe the failure mechanisms
Table 4
Mechanical properties of projectile, interior cushioning system as well as Hybrid III
headform [30,31].
E (MPa)
Steel spherical projectile
Polymer skin
Magnesium-alloy skull
Foam padding

Main loop/netting/cross straps


Front and back cushion

180,000
0.3
14,000
0.4
45,000
0.35
Direct compression
data from
experiment
60
0.25
18
0.25

r (kg/m3)
7935
437
1800
164

400
200
Fig. 6. In-house uniaxial compressive data of OA foams at different strain rates.

104

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

tetrahedral elements are assigned to various components of the


headform while 8-noded hexahedral elements are assigned to the
helmet and foam or straps. Although the actual physical ACH
contains more than 20 laminate layers, the FE model of the ACH in
our study is simplied with just 4 layers to reduce computational
cost and increase computational efciency, in the view of the

studys primary aim of comparing effectiveness of different cushioning systems for different impact orientation. It shall also be
noted that these 4 layers of the FE helmet model demonstrate the
delamination of layers. Fig. 4 shows the assembled components of
the two helmet-interior cushion-headform models. The number of
elements and element types are also provided in Table 1.

Fig. 7. Post-test photos of the helmets, showing the damages obtained from (A) frontal impact and (B) lateral impact.

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

In order to determine the extent of headform accelerations due


to ballistic impact, series of ballistic impact simulations with
explicit code are conducted using the nite element analysis (FEA)
software, Abaqus v6.10 (SIMULIA, RI, USA). The acceleration
values from the headform can then be correlated to various brain
traumatic injuries depending on the severity of shock/acceleration
transmitted to the headforms. Damage and energy absorbed by the
helmet and foams are also analyzed to look at the performance of
the components in attenuating impact shock.
2.3.1. Material properties
Since the ACH is made of KEVLAR which is an aramid
composite, its material properties depends on its ber directions. It
is a challenge to impose these material properties on the unique
shape of the helmet, ensuring that the X, Y and Z material direction
axes corresponds to the input material values. We have adopted the
technique employed by Gong et al. [27] whereby the helmet is
rstly broken into different plates before assigning the direction
perpendicular to the plates as the Z(3-) material axis, and directions
along the plates as X(1-) and Y(2-) material axes. Fig. 5 shows the
local coordinate systems applied to the various plates and the end
result of the material orientations dened to the helmet. As shown
in Fig. 5, the Z(3-) direction of the helmet is made to point
perpendicularly outwards from the helmets surface, while X(1-)
and Y(2-) material axes are points along the plane of the helmet.
The material properties of the ACH, foam padding, straps
netting, Hybrid III headform [28,29] and spherical projectile used
are shown in Tables 2e4. The ACH helmet properties are similar to
those used by Gong et al. [27]. For the strap netting, it is assumed to
be linearly elastic and its material properties are obtained from inhouse testing and also cross-referenced with those from Lee and
Gong [16]. The OA pads are tested in-house under dynamic
compression at different rates (Fig. 6). The experimental data are
then simplied and input into Abaqus through the use of the Low
Density Foam material model which is intended for low-density,
highly compressible elastomeric foams with signicant rate sensitive behavior, such as polyurethane foam. It assumes that the
Poissons ratio of the material is zero and allows the direct specication of uni-axial stressestrain curves at different strain rates for
both tension and compression. It shall be noted that reduced
integration with hourglass control is used for the formulation of the
foam elements. For the Hybrid III headform, the polymer skin is
assumed to be constantly in contact with the magnesium skull due
to the compressive forces exerted by the stretched polymer onto
the skull. Tie-constraints with a position tolerance of 0.1 mm are
used at the interfaces to secure the headform parts together whilst
other interfaces such as that between the helmet and interior

105

cushions as well as that between the cushions and headform are


imposed with the static-kinetic exponential decay frictional sliding
contact.
In view of the degree of inuence and importance of both interlamina and intra-lamina damages on the resulting helmet performance, both inter-ply delamination and intra-lamina damage of
the helmet laminate are modeled in our study so as to capture as
accurate and realistic as possible the mechanical response of the
helmet due to ballistic impact. In this study, the fabric-reinforced
aramid laminates of the helmet shell are modeled using the
Hashin Fabric Criterion that takes into account bi-directional
strength of the bers as woven into a fabric laminate. The failure
indices are as shown below.
For ber failure,
1st failure index: tensile ber 1 mode, s1 > 0


dft1

2  2  2
s
s
12 13
X1t
S12
S13

s1

2nd failure index: compressive ber 1 mode, s1 < 0

dfc1

2  2  2
s
s
12 13
X1c
S12
S13

s1

3rd failure index: tensile ber 2 mode, s2 > 0


dft2

2  2  2
s
s
12 23
X2t
S12
S23

s2

4th failure index: compressive ber 2 mode, s2 < 0


dfc2

2  2  2
s
s
12 23
X2c
S12
S23

s2

For matrix failure,


5th failure index: tensile matrix mode, s3 > 0

2  2  2  2
s
s
s
12 13 23
X3t
S12
S13
S23


dmt

s3

6th failure index: compressive matrix mode, s3 < 0


dmc

2  2  2  2
s
s
s
12 13 23
X3c
S12
S13
S23

s3

The above failure criteria, based in quadratic form, help to


account for tensile and compressive failure of the matrix, as well as,
ber buckling and ber breakage. The numerators of the terms are
the stresses obtained from analysis while the denominators are the
material strength values as provided in Table 3.

Fig. 8. Failure mechanisms observed in post-test ACH using optical micrography. (A) The sagittal image depicting the front impact region. (B) The superior and coronal cross-section
images of the helmet at the lateral impact region.

106

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

Failure is assumed when either of the indices reaches the value


of one. However, the element is not deleted until the bers fail in
tension, that is, for either the rst or the third failure index to reach
one. This procedure is preferred since experimentally, the bers or
matrix are usually retained and continue to resist forces even when
they fail in compression. Matrix failure in tension is also not
considered as the criteria for element deletion because the bers
may still be intact to continue in resisting loads.

3. Results
3.1. Post-test failure analysis
After the ballistic tests, the damage to the helmet is documented
via visual observation, optical microscopy and CT scan. Fig. 7A show
the post-test observations made on the two tested ACHs after frontal
impact, while Fig. 7B show the pictures of the ACH helmets after

Fig. 9. Post-test comparison between the FE simulation and the CT scan images of the damaged helmet for (A) frontal impact and (B) lateral impact.

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

107

Fig. 10. Matrix compression and ber buckling damage on (A) helmet with strap-netting system (Helmet 1) and (B) helmet with interior OA foam padding (Helmet 2).

The crater or spatial extent of the impression is around 60 mm in


diameter, and a bulge can be seen at the backplane of the helmet.
Fig. 8 shows the optical micrographs depicting fabric laminate
bowing, yarn rupture and ply delamination. Bowing, which refers

lateral impact. Similar phenomenon has been observed in both


helmets, as shown in Fig. 7. In all four tests, there is no penetration of
projectile into the helmet. In general, a permanent dent of
12e18 mm on the helmet exterior is observed at the point of impact.
Table 5
Comparison between FE simulations and experimental tests for both models.
Impact orientation

Frontal impact @ 205 m/s

Specimen

Strap-netting
Test

Physical measurements
Helmet dent (mm)
(depth)
Permanent dent region (mm)
(Diameter)
Initial distance bet.
inner helmet surface
& head (mm)
Min. distance bet. inner
helmet surface
& head (mm)

Strap-netting

OA foam

Simulation

Test

Simulation

Test

Simulation

Test

Simulation

1895.0
676.7

753.3
362.3

1260.0
460.0

897.5
77.3

1435.0
127.0

1444.4
153.5

1580.0
240.0

24.5

15.0

15.0

39.0

10.0

26.0

251.3

284.4

248.4

248.7

X (Projectile
captured
by helmet
laminate)
285.3

278.9

288.6

283.9

21.3

14.6

21.5

14.8

31.9

19.4

32.3

18.0

Front cushion
compressed
during impact

Front cushion
compressed
during impact

Foam pad
compressed
during impact

Foam pad
compressed
during impact

Delaminated
helmet shell
just touching
skin

Delaminated
helmet shell not
touching skin

Delaminated
helmet shell
almost touching
skin

Delaminated
helmet shell
not touching
skin

15.0

6.4

12.6

6.5

18.7

10.4

13.1

9.0

42.0

32.0

46.0

27.5

42.0

32.0

42.0

32.0

19.5

21.9

29.9

30.0

14.3

15.9

18.6

19.8

1346.7
Peak helmet strains (m3)
Peak accel. along impact
439.2
direction (G)
High-speed camera measurements
Rebound velocity
21.0
of projectile (m/s)

Energy absorbed by
helmet/head (J)
(Eballimpact  Eballrebound)
Est.dynamic deection
(mm) (radius)
Observations from
HS images

Lateral impact @ 220 m/s


OA foam

108

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

to the misalignment of orthogonal yarns, is observed when yarns


are pushed aside instead of being broken by projectile perforation,
and is induced by transverse deformation of the fabric laminate
which distorts the alignment between orthogonal yarns. Yarn
rupturing is also observed and occurs when yarns are broken due to
over-stretching. This results in ber breakage and fraying. It is
noted that, from our ballistic tests, yarn rupture and bowing are
observed to be more signicant in lateral impact than frontal
impact. Matrix compression is also a predominant mechanism
observed in the ber-(epoxy) matrix composite whereby impact
energy is absorbed in the form of permanent matrix deformation of
the laminate plies without failure of the KEVLAR yarns at the inner
layers.
3.2. FE Model validation with experimental data
As seen in Fig. 9, the FE simulated results of the impacts are
comparable to those through observations of the postexperimental CT scans. The damages on the helmet for both
frontal and lateral impact analysis, as well as, deformation experienced during impact are shown in Fig. 9. The predicted region of
matrix compression damage, ber buckling and delamination
corresponds well with post-test observations from CT and optical
images (Figs. 7e10).
In order to ascertain the robustness and accuracy of predictions
by the FE models, certain quantiable parameters such as
maximum deection of the helmet, the depth and diameter of
impression, and the projectiles rebound velocity are being
compared. Table 5 shows the comparison of parameters obtained
from both experiments and simulations for front and side ballistic
impacts to the assembly. It is observed that the acceleration values
and projectiles rebound velocities are quite well correlated.
Fig. 11A and B depict the overlay plots of the acceleration proles
of the headform for both experiment and simulation for frontal
and lateral impacts respectively. It demonstrates that the FE
models can give a reasonable prediction of impact accelerations as
compared to the experimental acceleration (G) values. The peak
strains from numerical analyses have generally higher values
compared to those obtained from experimental data and this is
believed to be due to the adhesive and plastic backing below the
gauges that effectively damped the experimental strain responses
(Table 5).
In general, there is reasonably good correlation between the
simulation results and experimental data in terms of headform
acceleration and observed failure modes. Good correlation of the
type and spatial extent of helmet damage between experiment and
FE results further authenticates the quality of the numerical model
in predicting the correct transmitted impact force to the head.
3.3. High speed camera photography
Observations made from high-speed camera images, post-test
inspections and CT-scans reveal that both inter-lamina and intralamina damages to the helmets are extensive. The high speed
camera images for the ballistic impact to the front and side of the
helmet-cushion-headform assembly are shown as a series of
pictures in Figs. 12 and 13. These images are selected to illustrate
the sequence of events from initial impact to the deformation of the
helmet shell and then to rebound of the projectile and subsequent
elastic recovery of the helmet shell.
The use of the high speed camera allows the inspection of the
dynamic response of the components which is otherwise difcult
to observe with the human eye. The series of pictures show the
deformation undergone by the helmet via impact of the projectile
at 205 m/s (for frontal) or 220 m/s (for lateral). Upon impact, paint

Fig. 11. Graphs of head acceleration against time for (A) front impact and (B) lateral
impact.

aks from the region of the helmet closest to the projectile are
discharged. The projectile then compresses and deforms the local
helmet shell region where it is in contact with. This eventually
causes a localized protrusion/bulge of the shell at the helmet rim.
For the frontal impact, the dynamic deformation of the helmet
eventually compresses the front foam padding. The estimated foam
loading rate is between 5 m/s to 7 m/s. At about 0.82 ms after
impact, the projectile starts to reverse its direction of motion and
rebound off the shell. This is followed simultaneously with the
elastic recovery of the shell and foam. For the lateral impact, the
helmet deforms signicantly, causing the interior helmet shell to
just touch the underlying headform. Upon impact, the helmet shell
deforms extensively inwards. Local exing of the helmet rim occurs
subsequently due to the different localized velocities between the
impact region and the helmet rim. Instead of rebounding, as in the
frontal case, the projectile is embedded into the helmet, followed
by elastic recovery of the helmet shell.
4. Discussion
4.1. Effects of impact orientation and different cushion systems
Observations from both experiments and simulations are discussed in this section. Firstly, it is noted that there is no penetration
of the projectile into the helmet for both frontal and lateral impacts
at the projectile velocity of approximately 200 m/s. The post-test
analysis of the ACHs reveals delamination of the laminate plies,
ber rupture and matrix compression damage. The majority of the

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

109

Fig. 12. High-speed camera images for frontal ballistic impact on the helmet with OA foams (Helmet 2).

damage occurs within the exterior half of the helmet thickness.


Secondly, our study also indicates that the impact forces experienced by the headform in frontal and lateral impacts are higher for
Helmet 1 (with strap-netting) and for Helmet 2 (with OA foam)
respectively (i.e. Helmet 1: 439 G for frontal impact and 77 G for
lateral impact; Helmet 2: 362 G for frontal impact and 154 G for
lateral impact) (Table 5 & Fig. 14A). The OA foam paddings in
Helmet 2 reduces acceleration for frontal impact, mainly because of
the superior capability in absorbing impact shock, compared to the
stiffer front cushion of Helmet 1 (Fig. 15). Further parametric
analyses, although the results are not presented in this paper, show
that softer cushions are more capable in attenuating impact shock,

through foam deformation (which allows distribution of impact


forces to larger surface areas) and plateau characteristics of the
foam stressestrain curves (which deforms the foam without
transmitting high contact impact forces to the head). A similar
conclusion on the effect of foam stiffness in mitigating ballistic
head impact has been reported by Lee and Gong [16]. In the event of
lateral impact, the presence of a large air gap between the interior
surface of the lateral side of the Helmet 1 (with strap-netting) and
the headform allows the helmet to ex considerably without
having to contact the headform immediately upon impact.
Furthermore, the impact energy due to lateral impact is primarily
absorbed through large deection of the helmet shell, whereas for

110

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

Fig. 13. High-speed camera images for lateral ballistic impact on the helmet with strap-netting system (Helmet 1). Notice the large gap between the side of the helmet shell to
the headform.

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

111

spines, resulting in higher linear acceleration during frontal


impacts. The results herein have also been compared with prior
dynamic tests on human cadavers and dummies. It is found that the
present results are in line with Jones and Mohan [30]s observations
in a series of dynamic baseball impact tests on dummies, which
concluded that the frontal impacts produce higher mean head
acceleration than lateral impacts. The present results also agree
well with the dynamic impact study of Got et al. [31] in 1978 using
helmeted cadavers, which reported that head accelerations were
50% higher in frontal impacts than lateral impacts. However, this
does not infer that frontal impact is more severe than lateral impact
as the tolerance limit is lower for lateral impact than frontal
impact [32].
4.2. Comparison with injury criteria

Fig. 14. (A) Bar chart showing the peak accelerations of the two designs in both frontal
and lateral impacts; (B) acceleration responses (indicated in the legend) for the
helmets with two interior cushion designs, in relation to other published criteria (from
Shewchenko et al. [36]).

frontal impact, a portion of the energy is absorbed by the paddings


and that the maximum deection of the helmet shell is lesser than
that observed for lateral impact (Table 5).
Thirdly, for both cushioning systems, the resultant G forces
experienced from a frontal impact on the helmet-cushionheadform are higher than that from lateral impact, even when
the latter has higher projectile speeds and produced greater impact
energy. This may be attributed to the Hybrid III headeneck complex
being stiffer in the front-back direction which may also be physiologically suggested by the anatomical positioning of the cervical

Holbourn (1943) [33] suggested that the two main causes of


brain injury; linear acceleration causing focal brain damage
through skull deformation (bending and fracture) while angular
acceleration generating shear strains and diffuse brain injury. It is
generally believed that rotational kinematics is more damaging to
the brain than linear acceleration even though there have been
controversies over the mechanism of brain injury in regards to
whether it is caused mainly by linear acceleration or rotational
acceleration over the decades [34]. However, the present study is
limited to the analysis of only linear acceleration values due to the
difculties in measuring rotational acceleration in the current
setup. Therefore, the present peak linear accelerations are
compared with some established injury criteria such as Wayne
State Tolerance Curve (WSTC) and Federal Motor Vehicles Safety
Standards (FMVSS) 218 criterion.
According to the WSTC, for linear acceleration loading, the risk
of brain damage due to non-penetrating impacts is determined by
both the magnitude and duration of the acceleration pulse. Short
duration, high acceleration impacts (2 ms, 200 G) lead to similar
injury risk as long duration, low acceleration impacts (20 ms, 55 G).
This is known as WSTC criterion [35e37]. Another commonly used
criterion of head injury is the FMVSS 218 criterion, which forbids
peak acceleration exceeding 400 G [36,38].
From the acceleration proles provided in Fig. 11, the pulse
duration for the peak accelerations are between 0.1 and 0.2 ms,
although the headform took about 1.5 ms for the impact to dampen
out and for accelerations to reach below 50 G. It can be concluded
that for frontal impact, Helmet 1 (with strap-netting), which
experiences up to 439 G impact acceleration, fails both the WSTC
and FMVSS 218 criteria, whilst Helmet 2 (with OA foam), of 362 G
acceleration, barely passes the criteria (Fig. 14B). For lateral ballistic
impact, the peak acceleration values of both helmets are observed
to be within the WSTC and FMVSS 218 limits. However, it shall be
noted that the human tolerance for lateral impacts is lower than
that of frontal impacts even though lower G is experienced from
side impacts [32].
5. Conclusion

Fig. 15. Graph of energy absorbed by all the OA foams against time for frontal impact.

In this study, ballistic analysis using FEM as well as experimental


impact tests has been carried out to evaluate the performance of
the ACH as well as the effectiveness of its interior cushioning
systems. The FE results, in terms of the mechanical behavior of the
assembly, extent of various forms of damage to the helmet laminate
and the experimental parameters, correlated well with those obtained from ballistic experiments. Frontal impact is observed to
have G value than lateral impact. The use of OA foams helps to
reduce frontal impact G forces and thus offers better protection
from frontal impacts than the strap-netting system. Both

112

L.B. Tan et al. / International Journal of Impact Engineering 50 (2012) 99e112

experiments and FE simulations have shown that softer foams with


lower stiffness are more effective as shock absorbing materials
against ballistic impacts under the condition of intensity below NIJ
Level 2 [39].
Moreover, a novel methodology in employing radiological
images, namely the CT, to rapidly generate a 3D FE assembly of
combat helmet, interior cushions, and the Hybrid III headform, is
also presented in the present study. This proposed methodology
will gain popularity and usage due to the time and cost savings
involved in physical testing and prototyping. More importantly,
with the numerical model setup and validated, subsequent nite
element analyses (FEA) can be performed better to design better
interior cushioning system for the ACH helmet, to implement better
protective helmet designs, and in the process, protect both military
personnel and civilians from traumatic head injury.
Acknowledgements
This work was supported by the Defence Science Organization
(DSO) of Singapore. Special thanks to their researchers for
providing the Advanced Combat Helmets (ACHs) and HybridIII
headform. We also acknowledge Dr. Lee Shu Jin (Division of Plastic,
Reconstructive and Aesthetic Surgery, National University Health
System) and Mohd Fairus Zam Zam (National University Health
System) for the usage of CT equipment and Mr. Low Chee Wah
(Mechanical Engineering Department, National University of
Singapore) for helping out with the ballistic tests.
References
[1] Mathers CD, Lopez AD, Murray CJL. The burden of disease and mortality by
condition: data, methods and results for 2001. In: Lopez A, Mathers C,
Ezzati M, Jamison DT, Murray CJL, editors. Global burden of disease and risk
factors. Washington, DC: The International Bank for Reconstruction and
Development/The World Bank Group; 2006. p. 45e240.
[2] Carroll AW, Soderstrom CA. A new nonpenetrating ballistic injury. Annals of
Surgery 1978;188:753e7.
[3] Alley Jr RH. Head and neck injuries in high school football. The Journal of the
American Medical Association 1964;188:418e22.
[4] Patrick LM. Head impact protection. In: Caveness WF, Walker AE, editors.
Head injury conference. Philadelphia: Lippincott; 1966. p. 41e48.
[5] Ewing CL, Irving AM. Evaluation of head protection in aircraft. Aerospace
Medicine 1969;40:596e9.
[6] Rayne JM, Maslen KR. Factors in the design of protective helmets. Aerospace
Medicine 1969;40:631e7.
[7] Haug E, Choi H-Y, Robin S, Beaugonin M. Human models for crash and impact
simulation. In: Ayache N, editor. Handbook of numerical analysis. Elsevier;
2004. p. 231e452.
[8] Goldsmith W, Khalil TB. Effect of a protective device in the reduction of head
injury. In: Int. Conf. on Biokinetics of Impacts. Amsterdam, The Netherlands;
1973.
[9] Khalil TB. Impact on a model head-helmet system. PhD Dissertation. Berkeley:
University of California; 1973.
[10] Khalil TB, Goldsmith W, Sackman JL. Impact on a model head-helmet system.
International Journal of Mechanical Sciences 1974;16:609e25.
[11] Van Hoof J, Deutekom MJ, Worswick MJ, Bolduc M. Experimental and
numerical analysis of the ballistic response of composite helmet materials. In:
18th International symposium on ballistics. San Antonio, TX, USA; 1999.
[12] Van Hoof J, Cronin DS, Worswick MJ, Williams KV, Nandlall D. Numerical head
and composite helmet models to predict blunt trauma. In: 19th International
symposium on ballistics. Interlaken, Switzerland; 2001.
[13] Baumgartner D, Willinger R. Finite element modelling of human head injuries
caused by ballistic projectiles. In: NATO Research & Technology Organisation
(RTO) specialist meeting. Koblenz, Germany; 2003.

[14] Aare M, Kleiven S. Evaluation of head response to ballistic helmet impacts


using the nite element method. International Journal of Impact Engineering
2007;34:596e608.
[15] Tham CY, Tan VBC, Lee HP. Ballistic impact of a KEVLAR helmet: experiment
and simulations. International Journal of Impact Engineering 2008;35:
304e18.
[16] Lee HP, Gong SW. Finite element analysis for the evaluation of protective
functions of helmets against ballistic impact. Computer Methods in Biomechanics and Biomedical Engineering 2010;13:537e50.
[17] Yang J, Dai J. Simulation-based assessment of rear effect to ballistic helmet
impact. Computer-Aided Design and Applications 2010;7:59e73.
[18] Tan VBC, Ching TW. Computational simulation of fabric armour subjected to
ballistic impacts. International Journal of Impact Engineering 2006;32:
1737e51.
[19] Rao MP, Duan Y, Keefe M, Powers BM, Bogetti TA. Modeling the effects of yarn
material properties and friction on the ballistic impact of a plain-weave fabric.
Composite Structures 2009;89:556e66.
[20] Mamivand M, Liaghat GH. A model for ballistic impact on multi-layer fabric
targets. International Journal of Impact Engineering 2010;37:806e12.
[21] Feli S, Yas MH, Asgari MR. An analytical model for perforation of ceramic/
multi-layered planar woven fabric targets by blunt projectiles. Composite
Structures 2011;93:548e56.
[22] Ha-Minh C, Boussu F, Kanit T, Crpin D, Imad A. Analysis on failure mechanisms of an interlock woven fabric under ballistic impact. Engineering Failure
Analysis 2011;18:2179e87.
[23] Ha-Minh C, Kanit T, Boussu F, Imad A. Numerical multi-scale modeling for
textile woven fabric against ballistic impact. Computational Materials Science
2011;50:2172e84.
[24] Abu Talib AR, Abbud LH, Ali A, Mustapha F. Ballistic impact performance of
kevlar-29 and Al2O3 powder/epoxy targets under high velocity impact.
Materials & Design 2012;35:12e9.
[25] Mills NJ. Finite element modelling of foam deformation. In: Mills NJ, editor.
Polymer foams handbook: engineering and biomechanics applications and
design guide. Oxford: Butterworth-Heinemann; 2007. p. 115e45.
[26] Goel R. Study of an advanced helmet liner concept to reduce TBI: experiments
& simulation using sandwich structures. M.Sc Dissertation. 130 p. Berkeley:
Department of Aeronautics and Astronautics, Massachusetts Institute of
Technology (MIT); 2010. 130.
[27] Gong SW, Lee HP, Lu C. Computational simulation of the human head
response to non-contact impact. Computers & Structures 2008;86:758e70.
[28] Yang KH, Le J. Finite element modeling of Hybrid III headeneck complex. In:
36th Stapp car crash conference. Society of Automotive Engineers (SAE), SAE
paper no. 922526 Seattle, WA, USA; 1992. p. 777e789.
[29] Chang C-Y. Development of a coupled impact model to study of motorcycle
helmet protecting effect. M.Eng Dissertation. 101 p. Chiayi, Taiwan, Republic
of China: Institute of Mechanical Engineering; College of Engineering,
National Chung Cheng University; 2000. 101.
[30] Jones IS, Mohan D. Head impact tolerance: correlation between dummy
impacts and actual head injuries. Report in TRB of National Academies. 32 p.
Arlington, VA, US: Insurance Institute for Highway Safety; 1984.
[31] Got C, Patel A, Fayon A, Tarriere C, Walsch G. Results of experimental head
impact on cadavers: the various data obtained and the relation to some
measured physical parameters. In: 22nd Stapp car crash conference, Society of
Automotive Engineers (SAE), SAE Paper No. 780887, Ann Arbor, USA; 1978. p.
57e99.
[32] Allsop D, Kennett K. Skull and facial bone trauma. In: Nahum AM, Melvin J,
editors. Accidental injury: biomechanics and prevention. 2nd ed. New York,
US: Springer-Verlag; 2002. p. 577.
[33] Holbourn A. Mechanics of head injuries. Lancet 1943;2:438e41.
[34] King AI, Yang KH, Zhang L, Hardy W, Viano D. Is head injury caused by linear
or angular acceleration? In: International IRCOBI conference on the biomechanics of impacts, Lisbon, Portugal; 2003. p. 1e12.
[35] Gurdjian ES, Webster JE, Lissner HR. Observations on the mechanism of brain
concussion, contusion, and laceration. Surgery, Gynecology and Obstetrics
1955;101:680e90.
[36] Shewchenko N, Withnall C, Keown M, Gittens R, Dvorak J. Heading in football.
Part 1: development of biomechanical methods to investigate head response.
British Journal of Sports Medicine 2005;39(Suppl. 1):i10e25.
[37] Gurdjian ES, Lissner HR, Patrick LM. Protection of the head and neck in sports.
The Journal of the American Medical Association 1962;182:509e12.
[38] Federal Motor Vehicle Safety Standards (FMVSS), Standard No. 218: motorcycle helmets, U.S.D.o. Transportation; 1998.
[39] National Institute of Justice (NIJ) Standard, NILECJ-STD-0106.00: ballistic
helmets, U.S.D.o. Justice, Washington, DC, US; September 1975.

You might also like