You are on page 1of 13

This article was downloaded by: [2.137.131.

0]
On: 25 August 2015, At: 11:48
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: 5 Howick Place, London, SW1P 1WG

Composite Interfaces
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tcoi20

High impact velocity on multi-layered


composite of polyether ether ketone
and aluminium
a b a
D. García-González , M. Rodríguez-Millán , A. Vaz-Romero & A.
a
Arias
a
Department of Continuum Mechanics and Structural Analysis,
University Carlos III of Madrid, Avda. de la Universidad 30, 28911
Leganés, Madrid, Spain
b
Department of Mechanical Engineering, University Carlos III of
Madrid, Avda. de la Universidad 30, 28911 Leganés, Madrid, Spain
Click for updates Published online: 09 Jun 2015.

To cite this article: D. García-González, M. Rodríguez-Millán, A. Vaz-Romero & A. Arias (2015): High
impact velocity on multi-layered composite of polyether ether ketone and aluminium, Composite
Interfaces, DOI: 10.1080/09276440.2015.1051421

To link to this article: http://dx.doi.org/10.1080/09276440.2015.1051421

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [2.137.131.0] at 11:48 25 August 2015
Composite Interfaces, 2015
http://dx.doi.org/10.1080/09276440.2015.1051421

High impact velocity on multi-layered composite of polyether ether


ketone and aluminium
D. García-Gonzáleza, M. Rodríguez-Millánb, A. Vaz-Romeroa and A. Ariasa*
a
Department of Continuum Mechanics and Structural Analysis, University Carlos III of Madrid,
Avda. de la Universidad 30, 28911 Leganés, Madrid, Spain; bDepartment of Mechanical
Engineering, University Carlos III of Madrid, Avda. de la Universidad 30, 28911 Leganés,
Madrid, Spain
(Received 13 April 2015; accepted 12 May 2015)
Downloaded by [2.137.131.0] at 11:48 25 August 2015

Hybrid structures composed of multi-layer sheets of polymers and metals offer a


great potential for impact protection systems in automotive and aeronautics indus-
tries. This work presents an experimental and numerical investigation on the perfora-
tion behaviour of layered polymer/metal composites. Penetration tests have been
conducted on sandwich panels of aluminium 2024-T3 skins and polyether-ether-ke-
tone (PEEK) cores, using spherical projectiles. The perforation experiment covered
impact velocities in the range 250–500 m/s. The initial and residual velocities of the
projectile were measured, and the ballistic limit velocity was obtained. The impact
mechanical behaviour of PEEK core is compared with Ti6Al4V titanium core, for
the same areal density of protection. It has been shown high perforation efficiency
of PEEK polymer and a promised application for aeronautical protections. A
numerical modelling is presented and validated with experimental data.
Keywords: impact properties; hybrid structures; aluminium alloys; laminated
materials; composite materials

1. Introduction
Impact and blast threats exist in a wide range of engineering, security and defence
sectors. The protection of civil infrastructures and critical industrial facilities are the
topics of increasing relevance to defence agencies and governments. In the transport
industry, energy absorption and crashworthiness are key points in the design process of
vehicles, vessels and aircrafts. Development of protective composite structures capable
of sustaining an impact keeping the structural integrity is thus one of the main chal-
lenges of modern industry. In the design and development of lightweight structural
solutions suitable for energy absorption under impact loading, the material selection
represents a crucial decision.[1] Moreover, impact on composite metallic plates is a
complex and complete problem including dynamic behaviour, fracture, damage, contact
and friction. We observe an internal energy which is an irreversible thermodynamic
process due to transfer of kinetic energy, dynamic plastic flow, elastic and plastic wave
propagation and large plastic deformation at high strain rates inducing thermal soften-
ing responsible of instabilities. It has been observed during this kind of projectile-plate
impacts that the nose shape of the projectile used changes the energy absorbed, the

*Corresponding author. Email: aariash@ing.uc3m.es

© 2015 Taylor & Francis


2 D. García-González et al.

failure mode and the ballistic limit, which is a decisive variable for optimum design.[2]
This process is strongly coupled to hardening, strain rate and temperature of involved
material.
In recent years, metal–polymer–metal sandwich sheets show a high potential in
forming and design to be used in protective structures vs. monolithic metal plates.[3]
Of all possibilities for designing hybrid systems, three-layered metal–polymer–metal or
multi-layer sheets offer a great potential for automotive, construction, naval industries
and aeronautics. Examples are given by the well known, GLARE (aluminium layers
and Glass Fibre Reinforced Epoxy), which has been applied in aviation industry (e.g.
Airbus 380) or ARALL (Aramid Fibre Reinforced Aluminium).[4] The comparison
between different configurations of hybrid laminated polymer/metal with values of
similar areal density is an open research topic, in terms of energy absorption, critical
perforation velocity and failure mode.[5] In this context, structural polymers are often
used in applications where impact resistance is required. Though specific thermoplastic
such as polyether ether ketone (PEEK) shows a strong positive function of strain rate,
Downloaded by [2.137.131.0] at 11:48 25 August 2015

[6,7] it has not been widely investigated for high impact applications. In the present
work, the impact behaviour of sandwich panels of 2024-T3 aluminium skin and PEEK
core is investigated and compared with sandwich panels of 2024-T3 aluminium skin
and Ti6Al4V core. Experimental data show similar ballistic limit for the same
areal density. Additionally, a numerical modelling is presented and validated with
experimental data.

2. Materials and experimental set-up


Multi-layered composites were composed of different combination of the following
plates: 2024-T3 aluminium with 1 mm of thickness, titanium alloy Ti6Al4V with 1 mm
of thickness and PEEK with 3 mm of thickness. According to Figure 1 and Table 1, three
different multi-layered composites configurations were selected to compare the similar
areal density: configuration I composed by four metal sheets (Al2024-Al2024-Al2024-
Al2024), configuration II composed by three metal sheets (Al2024-Ti6Al4V-Al2024) and
configuration III composed by metal and polymer sheet (Al2024-PEEK-Al2024). The
characteristics of these materials are shown next.

2.1. 2024-T3 aluminium


2024-T3 is an aluminium alloy, with Cu and Mg as the main alloying elements.
Applications of this material can be found in aircraft structural components, wing

Figure 1. Scheme of multi-layered system protection considered: (a) Configuration I:


Al–Al–Al–Al, (b) Configuration II: Al–Ti–Al and (c) Configuration III: Al-PEEK-Al.
Composite Interfaces 3

Table 1. Areal density and thickness of multi-layered system protection.

Areal density
Type Protection Thickness (mm) (kg/m2)
Configuration I 4 sheet of 2024-T3 Al 1-1-1-1 10.7
Configuration II 2 sheet of 2024-T3 Al + Ti6Al4V core 1-1-1 9.9
Configuration III 2 sheet of 2024-T3 Al + PEEK core 1-3-1 9.9

tension members, hardware, truck wheels, scientific instruments, veterinary and


orthopaedic braces and equipment; and in rivets because of its high strength, excellent
fatigue resistance and good strength-to-weight ratio.

2.2. Ti6Al4V titanium


Ti 6Al-4V is a titanium alloy considered in many applications due to its high strength
Downloaded by [2.137.131.0] at 11:48 25 August 2015

at low to moderate temperatures and light weight. The main applications are the aircraft
turbine engine components, aircraft structural components, aerospace fasteners,
high-performance automotive parts, marine applications, medical devices and sports
equipment.

2.3. PEEK polymer


As a typical high-performance semi-crystalline thermoplastic polymer, polyether-ether-
ketone, PEEK has received significant attentions in recent years. This is due to its high
mechanical strength and elastic modulus, good combination of thermal and mechanical
properties, chemical inertness, high toughness, easy processing, high wear resistance
and friction coefficient.[8] This thermoplastic has a melting point of 614 K and glass
transition of 417 K. It can be used continuously up to 520 K without any permanent
loss of mechanical properties. PEEK and its composites are good candidate materials
for a variety of structural applications in aerospace, biomechanics, automotive and
chemical industries. In space application, PEEK is applied for replacing aluminium
because of its superior performance at high temperatures. In addition, PEEK is also one
of a new generation of engineering polymers which has good cryogenic properties.
Due to glass fibres (GFs) have high strength, high flexural modulus and low expansion
rate, they are the most common fibre reinforcements in thermoplastics to reduce the
expansion rate and increase the flexural modulus of PEEK.[8,9]

2.4. Experimental set-up


Spherical projectile was used in perforation test with a diameter of 7.52 mm and a mass
of 1.7 g. In addition, the projectile underwent a heat treatment to increase their hard-
ness. To perform perpendicular impact tests on the multi-layered plates, a pneumatic
gas gun was used. It should be noticed that the diameter of the barrel was roughly
equal to the diameter of projectile. No sabot was required for guidance of projectile
inside the barrel, which helps to ensure the perpendicular of the impact. The impact
velocity and residual velocity were measured using lasers coupled to photodiodes and
with a high-speed record camera. Perforation experiments were performed with impact
velocities in the range from 250 to 500 m/s. It should be noted that, for all the tests
4 D. García-González et al.

performed, the projectiles showed an absence of plastic straining, damage or erosion


after the impact. Next, the experimental and numerical results are discussed.

3. Numerical model
3.1. Mesh definition and boundary conditions used during numerical simulations
The numerical model was developed using the explicit finite element code Abaqus/Ex-
plicit.[10] There were implemented three numerical models which represent the differ-
ent combinations of multi-layered structures tested. Therefore, the numerical model
consisted of various solids: the multi-layered composite composed by three or four
plates (depending of the sandwich selected) and the projectile. Rusinek et al. [11] have
noticed the convenience of using 3D approach to simulate particular aspects of the
problem, mainly related to petalling, since this failure mode is a non-symmetric pro-
cess. In the numerical simulations reported in the present work, also 3D numerical
approach was used. For each plate, two kinds of elements were used in order to get a
Downloaded by [2.137.131.0] at 11:48 25 August 2015

good damage-dependent evolution of the zone. As Figure 2 shows, each plate is


divided into three zones: zone A is the area in contact with the projectile. The four-
node linear tetrahedral element (C3D4), which allows greater freedom of damage prop-
agation, was implemented in the impact region of each plate defined as an area of
14.5 mm diameter (twice the diameter of the projectile), zone B corresponds to the
transition zone between the impact zone and the rest of the plate. It was meshed with
eight-node brick hexahedral elements with one integration point (C3D8R), and zone C
corresponds to peripheral region of the plate. In this zone, the sensitivity of the mesh
was reduced by a gradually reduction of the element size. There were defined four ele-
ments along the thickness for plates of both metal alloys and ten elements for the plate
of PEEK. This is in agreement with the recommendations reported [12] where it is sug-
gested that at least four elements should be used through the thickness when modelling
any structures carrying bending loads. This optimum configuration has been obtained
from a convergence study using different mesh densities. The sandwich plate was

Figure 2. Mesh corresponding to sandwich configuration divided into three zones.


Composite Interfaces 5

constrained by all its lateral faces, and the rigid body was constrained not to move in
the X and Y directions.

3.2. Modelling of thermoviscoplastic behaviour


In order to define the stress–strain constitutive relationship of the semi-crystalline poly-
mer PEEK, a phenomenological Johnson–Cook (JC) model [13] was used as the mate-
rial model according to strain rate and temperature sensitivity viscoplastic behaviour of
PEEK reported in [6]. Since the JC model is frequently used to define the material
behaviour in ductile metals,[14,15] it has been also employed to analyse the dynamic
behaviour of polymer materials.[16,17] Besides, this model has been recently used in
dynamic problems to define the thermoviscoplastic behaviour of PEEK and the parame-
ters of the model for this material are reported in [12]. The JC model defines the effec-
tive flow stress as a function of equivalent plastic strain ep , equivalent plastic strain rate
e_ p and temperature T:
  
Downloaded by [2.137.131.0] at 11:48 25 August 2015

e_ p
r _
 
ðep ; ep ; T Þ ¼ ½A þ B  ep  1 þ C  ln
n
 ½1  THm  (1)
e_ 0
where TH is the homologous temperature defined as
T  T0
TH ¼ (2)
Tmelt  Troom
where T is the current material temperature, Tmelt is the melting temperature, and Troom
is the room temperature define as 296 K for all cases. The different constants are
defined in Table 2.
It is important to consider a thermal softening caused by a great location of plastic
deformation. In addition, this constitutive relation coupled to the heat equation,
Equation (3), allows us to obtain the temperature increase ΔT in adiabatic conditions:
Z
b
DT ðep ; e_ p ; T Þ ¼ ðep ; e_ p ; T Þdep
r (3)
q  Cp
where β is the percentage of plastic work transformed to heat (Quinney Taylor
coefficient), Cp is the heat capacity, and ρ is the density. The constants which define
the thermal evolution due to adiabatic heating effects are shown in Table 3.
Regarding both metal alloys used as composite skins (aluminium 2024) and as a
metal core (Ti6Al4V), several approaches may be used including physical approxima-
tions like the JC model defined before.[13,18]

Table 2. General material properties of Ti6Al4V,[13] aluminium 2024-T3 [18] and polyether
ether ketone.[9,12]

Elastic properties Other physical constants


Material E (GPa) ν ρ (kg/m3) β Cp (J/kg K)
Ti6Al4V 109.8 0.31 4428 0.9 560
2024-T3 aluminium 70 0.3 2700 0.9 900
PEEK 3.6 0.4 1304 0.9 2180
6 D. García-González et al.

Table 3. Johnson–Cook model parameters of Ti6Al4V,[13] aluminium 2024-T3 [18] and


polyether ether ketone.[12]

Material A (MPa) B (MPa) n C m Tmelt (K) e_ 0


Ti6Al4V 1098 1092 0.93 0.014 1.1 1878 1.0
2024-T3 aluminium 352 440 0.42 0.0083 1.7 775 3.33 × 10−4
PEEK 132 10 1.2 0.034 0.7 614 0.001

3.3. Modelling of the fracture model


Since the ductility of semi-crystalline polymers has been reported dependent on the
initial triaxiality (η) and strain rate (e_ p ),[12] it was included a JC fracture model [2]
available in Abaqus/Explicit as an erosive procedure inducing element deletion when
an imposed plastic strain level (defined for each time increment by Equation (4)) is
reached.
  
e_ p
Downloaded by [2.137.131.0] at 11:48 25 August 2015

ep ¼ D1 þ D2  expðD3  gÞ  1 þ D4  ln ½1 þ D5  TH  (4)


e_ 0
where η is the stress triaxiality, and Di are constants.
This JC fracture model has been commonly used to define the fracture criterion of
the metal alloys employed in this work, so it was selected to define the failure of both
metal materials.[14,15] The constants for each material are collected in Table 4 accord-
ing to a calibration by experimental data reported in [12,14,15]. Regarding interlaminar
damage, some authors propose using tied interfaces or tie break interfaces to simulate
delamination based on a critical shear stress.[19] In the present work, the effect of
adhesion between layers can be neglected. Therefore, the contact between layers has
been defined by a normal behaviour based on the ‘hard contact’ option which allows to
adjust automatically the stiffness to minimise penetration without adversely affecting
the time increment. In relation to frictional effects, it has been used a constant friction
coefficient value μ = 0.2 to define the contact metal skins/PEEK core.[12] In case of
the contact between metal layers, a constant friction coefficient value μ = 0.26 has been
employed as reported in [11]. The potential dependence of the friction coefficient on
the temperature and the sliding velocity is not taken into account. The constant value
used for the friction coefficient is based on the assumption of a constant pressure along
the projectile-plate and skin layer–core layer contact zone.[12]

4. Results
4.1. Experimental results
Three different sandwich configurations have been tested at impact velocities from 250
to 500 m/s: 2024Al-PEEK-2024 denoted configuration I, 2024Al-Ti6Al4V-2024Al

Table 4. Constants used to define the J-C damage criterion of Ti6Al4V,[14] aluminium 2024-T3
[15] and PEEK according experimental data.[12]

Material D1 D2 D3 D4 D5
Ti6Al4V −0.09 0.27 0.48 0.014 3.87
2024-T3 aluminium 0.112 0.123 1.5 0.007 0.0
PEEK 0.05 1.2 −0.254 −0.009 1.0
Composite Interfaces 7

denoted configuration II and 2024Al-PEEK-2024 denoted configuration III. Figure 3,


Figure 4 and Figure 5 show experimental observations of the final stage of perforation
process for configurations I, II and III, at impact velocities of 435, 305.8 and 298.7 m/
s, respectively. Plastic instabilities formation and progression are identified as the cause
of behind the target collapse for all the impact tests. For configuration I, composed by
four aluminium layers, a progressive damage from shear to petalling can be observed
from top layer to bottom layer (Figure 3). However, when the material selected for the
core layer is titanium in configuration II, the failure mechanism is based on cracking
propagation following the rolling direction and the perpendicular one (Figure 4). For
configuration III, using PEEK as core, the failure is based on ductile behaviour [12]
absorbing large amounts of energy (Figure 5).
It can be noticed that the core layer leads the energy absorption capability of the
multi-layered configuration and its failure mode varies depending on the material
employed. Regarding to material density of the core layer, it is well known that the
polymer layer slows down the impact waves compared with multi-layered metal
Downloaded by [2.137.131.0] at 11:48 25 August 2015

targets,[3] and consequently, a larger volume of the target gets involved in the
deformation process.
The experimental results of impact velocities have been analysed. Figure 6 shows
the residual velocity vs. impact velocity (Vr − V0) curves obtained for sandwich tested.
The ballistic limit is the maximum value of the initial impact velocity which induces a
residual velocity equal to zero. The ballistic limit of 2024Al-PEEK-2024 sandwich was
found Vbl ≈ 360 m/s to be just slightly greater than corresponding to 2024Al-Ti6Al4V-
2024Al sandwich, Vbl ≈ 332 m/s both of them with the same areal density (9.9 kg/m2).
Therefore, it has been shown high efficiency perforation of sandwich with PEEK core.
The results shown in Figure 6 have been fitted via the expression proposed by Recht
and Ipson [20]:

Figure 3. Final stage of perforation process for configuration I, at impact velocity equal to
435 m/s.

Figure 4. Final stage of perforation process for configuration I, at impact velocity equal to
305.8 m/s.
8 D. García-González et al.

Figure 5. Final stage of perforation process for configuration I, at impact velocity equal to
298.7 m/s (corresponding to arrest of projectile).
Downloaded by [2.137.131.0] at 11:48 25 August 2015

Figure 6. Residual velocity vs. impact velocity for the three multi-layered composite considered
in this work (experimental, numerical and analytical data).
Composite Interfaces 9

Vr ¼ ðV0k  Vblk Þ1=k (5)

where k is a fitting parameter. The value of k was determined for each configuration
corresponding to k = 1.55 for 2024Al-PEEK-2024Al; k = 1.55 for 2024Al-2024Al
-2024Al -2024Al; and k = 1.85 for 2024Al-Ti6Al4V-2024Al.

4.2. Numerical results


A qualitative approach for experimental and numerical data was obtained by the model
for residual velocity (Figure 6) and strain distribution at failure (Figure 7). An agreement
for residual velocity and a prediction of the residual velocity with an error of about 15%
was obtained by numerical model. These differences can be explained for some causes:
the identification of material based only on compressive tests and the Mises yield surface
used in the Johnson–Cook model for PEEK material. Besides, it is important to note that
the failure criterion applied involves element deletion. Therefore, special attention has to
Downloaded by [2.137.131.0] at 11:48 25 August 2015

be paid to the energy balance and its preservation during numerical simulations. In this
type of analysis,[12] the elimination of elements was restricted to the crack propagation
stage, and therefore, the approach is non-conservative.
Furthermore, it has been found a good agreement between numerical and
experimental results in terms of layer separation for the three configurations. Related to
the target deformation, the absorbed energy during the impact test is controlled by the
addition of the energy associated to each layer, being the deformation influenced by the
ductility of the next layer.[21] Therefore, at high impact velocity processes, the ten-
dency of separation between aluminium and core is significant and the aluminium layer
can dissipate more energy through membrane deformation (Figure 7). This free
deformation results in superior impact performance.[22] The use of PEEK as core layer
involves higher energy absorption which derives in lower values for residual velocity.

Figure 7. Numerical simulations of impact process for the three configurations of multi-layered
composites. Impact velocity equal to 380 m/s.
10 D. García-González et al.

5. Conclusions
High impact test in gas gun has been conducted on multi-layered composite of PEEK
and aluminium sheets using spherical projectiles from 250 to 500 m/s initial velocity.
The experimental and numerical results support the following major conclusions:

• In the full range of impact velocities, multi-layered composite with PEEK core is
more efficient for energy absorption in comparison with metal multi-layered sys-
tems. For metal skin, plastic instabilities formation and progression are identified
as the cause of behind the failure of sandwich structure. During experiments, the
PEEK core was observed to behave in a ductile manner without evidence of brit-
tle failure. This effect involves a larger volume of the target in the deformation
process and consequently a greater delamination between layers. Therefore,
higher energy absorption derives in lower values for residual velocity.
• Finite element simulations of experiments were performed using a plasticity
model typically used for ductile metals. The simulation was in good agreement
Downloaded by [2.137.131.0] at 11:48 25 August 2015

with the experimental data, confirming the predominantly ductile response of


PEEK under high strain rate. A qualitative agreement for residual velocity with a
maximum error of 15% was obtained by numerical model.

In conclusion, multi-layered composite of PEEK polymer appears to be an attractive


candidate for impact applications.

Acknowledgements
The researchers are indebted to LATI Company for PEEK material supplied. The authors express
their thanks to Mr Sergio Puerta, Mr David Pedroche and Ms Ascensión Aynat for their technical
support.

Disclosure statement
No potential conflict of interest was reported by the authors.

Funding
This work was supported by the Ministerio de Ciencia e Innovación de España under [grant
number DPI/2011-24068].

References
[1] Arias A, López-Puente J, Loya J, et al. Analysis of high-speed impact problems. In:
Constitutive relations under impact loadings. Vienna: Springer; 2014. p. 137–207.
[2] Arias A, Rodríguez-Martínez JA, Rusinek A. Numerical simulations of impact behaviour of
thin steel plates subjected to cylindrical, conical and hemispherical non-deformable
projectiles. Eng. Fract. Mech. 2008;75:1635–1656.
[3] Sokolova O, Carradò A, Palkowski H. Metal-polymer-metal sandwiches with local metal
reinforcements: a study on formability by deep drawing and bending. Compos. Struct.
2011;94:1–7.
[4] Carradò A, Faerber J, Niemeyer S, et al. Metal/polymer/metal hybrid systems: towards
potential formability applications. Compos. Struct. 2011;93:715–721.
Composite Interfaces 11

[5] Rodríguez-Millán M, Vaz-Romero A, Rusinek A, et al. Experimental study on the


perforation process of 5754-H111 and 6082-T6 aluminium plates subjected to normal impact
by conical, hemispherical and blunt projectiles. Exp. Mech. 2014;54:729–742.
[6] Rae P, Brown E, Orler E. The mechanical properties of poly(ether-ether-ketone) (PEEK)
with emphasis on the large compressive strain response. Polymer. 2007;48:598–615.
[7] Millett J, Bourne N, Stevens G. Taylor impact of polyether ether ketone. Int. J. Impact Eng.
2006;32:1086–1094.
[8] Chu X, Wu Z.X., Huang R, et al. Mechanical and thermal expansion properties of glass
fibers reinforced PEEK composites at cryogenic temperatures. Cryogenics. 2010;50:84–88.
[9] LATI. Lati peek materials properties data table. Technical report; 2014. Available from:
http://www.lati.com.
[10] Hibbitt H, Karlsson B, Sorensen P. Abaqus v6.12 documentation—ABAQUS analysis user’s
manual. Providence (RI): Dassault Systèmes Simulia; 2012.
[11] Rusinek A, Rodríguez-Martínez JA, Zaera R, et al. Experimental and numerical study on
the perforation process of mild steel sheets subjected to perpendicular impact by hemispheri-
cal projectiles. Int. J. Impact Eng. 2009;36:565–587.
[12] Garcia-Gonzalez D, Rusinek A, Jankowiak T, et al. Mechanical impact behavior of
polyether-ether-ketone (PEEK). Compos. Struct. 2015;124:88–99.
Downloaded by [2.137.131.0] at 11:48 25 August 2015

[13] Wang X, Shi J. Validation of Johnson-Cook plasticity and damage model using impact
experiment. Int. J. Impact Eng. 2013;60:67–75.
[14] Lee W, Chen C. High temperature impact properties and dislocation substructure of Ti–6Al–
7Nb biomedical alloy. Mater. Sci. Eng. 2013;576:91–100.
[15] Wierzbicki T, Bao Y, Lee Y, et al. Calibration and evaluation of seven fracture models. Int.
J. Mech. Sci. 2005;47:719–743.
[16] Louche H, Piette-Coudol F, Arrieux R, et al. An experimental and modeling study of the
thermomechanical behavior of an ABS polymer structural component during an impact test.
Int. J. Impact Eng. 2009;36:847–861.
[17] Duan Y, Saigal A, Greif R. A uniform phenomenological constitutive model for glassy and
semicrystalline polymers. Polym. Eng. Sci. 2001;41:1322–1328.
[18] Teng X, Wierzbicki T. Numerical study on crack propagation in high velocity perforation.
Comput. Struct. 2005;83:989–1004.
[19] Chai Gin Boay, Manikandan Periyasamy. Low velocity impact response of fibre-metal
laminates – a review. Compos. Struct. 2014;107:363–381.
[20] Recht R, Ipson T. Ballistic perforation dynamics. J. Appl. Mech. 1963;30:384–390.
[21] Pozuelo M, Carreño F, Ruano O. Delamination effect on the impact toughness of an
ultrahigh carbon–mild steel laminate composite. Compos. Sci. Technol. 2006;66:2671–2676.
[22] Sadighi M, Alderliesten R. Impact resistance of fiber-metal laminates: a review. Int. J.
Impact Eng. 2012;49:77–90.

You might also like