Download as pdf or txt
Download as pdf or txt
You are on page 1of 62

Journal of Materials Research and Technology

Characterization and Ballistic Performance of Hybrid Jute and Aramid Reinforcing


Graphite Nanoplatelets in High-Density Polyethylene Nanocomposites
--Manuscript Draft--

Manuscript Number: JMRT-D-23-09716R1

Article Type: Original article

Keywords: Jute; aramid; graphite nanoplatelet; high-density polyethylene; Hybrid


nanocomposite

Corresponding Author: Ulisses Oliveira Costa, Doctor


Military Institute of Engineering
Rio de Janeiro, Rio de Janeiro BRAZIL

First Author: Ulisses Oliveira Costa, Doctor

Order of Authors: Ulisses Oliveira Costa, Doctor

Fabio da Costa Garcia Filho, Doctor

Teresa Gómez-del Río, PhD

Édio Pereira Lima Júnior, Doctor

Sergio Neves Neves Monteiro, PhD

Lucio Fabio Cassiano Nascimento, Doctor

Abstract: Hybrid nanocomposites have emerged as a promising solution for engineering


applications associated with reduced costs and weight aiming to enhance performance
in special areas such as ballistic protection. This study delves into the development of
hybrid nanocomposites featuring a high-density polyethylene (HDPE) matrix modified
with graphite nanoplatelets (GNP) and reinforced by aramid and jute fabrics. The
incorporation of GNP significantly influences the crystalline structure of the GNP/HDPE
matrix, as evidenced by Raman and X-ray diffraction (XRD) analyses. It also plays an
important role in altering the dynamic mechanical behavior (DMA) of the
nanocomposites. Specifically, it increases stiffness, elevates the storage modulus, and
reduces the tan(δ) value. Additionally, substituting 5 layers of aramid for 5 layers of jute
fabric maintains statistically equivalent ballistic performance to the 20-layer aramid
nanocomposite. Furthermore, a hybrid nanocomposite consisting of 10 layers of jute
and 10 layers of aramid demonstrates impressive ballistic resistance against 9 mm
caliber ammunition. Scanning electron microscopy (SEM) analysis exposes the
intricate fracture mechanisms, encompassing phenomena such as crazing,
GNP/HDPE fibrillation, fiber rupture, and debonding. These mechanisms are significant
in the composite's energy absorption during a projectile impact. Moreover, a cost-
benefit analysis underscores the potential of hybrid nanocomposites in ballistic
protection by simultaneously reducing both helmet weight and cost in 7% and 40%,
respectively.

Suggested Reviewers: Dimitrios Papageorgiou, PhD


Senior Lecturer, Queen Mary University of London
d.papageorgiou@qmul.ac.uk
This professor has a lot of experience in the subject covered in the article.

Tuba Evgin, PhD


Professor, Dokuz Eylül University
tuba.evgin@deu.edu.tr
This professor has a lot of experience in the subject covered in the article

Eileen Harkin-Jones, PhD


Professor, Ulster University
e.harkin-jones@ulster.ac.uk
This professor has a lot of experience in the subject covered in the article.

Opposed Reviewers:

Response to Reviewers:

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Cover Letter

MINISTÉRIO DA DEFESA
EXÉRCITO BRASILEIRO
DEPARTAMENTO DE CIÊNCIA E TECNOLOGIA
INSTITUTO MILITAR DE ENGENHARIA
(Real Academia de Artilharia, Fortificação e Desenho/1792)

Rio de Janeiro, Nov 5, 2023.

Journal of Materials Research and Technology

Dear Editor in Charge,

The manuscript entitled " Characterization and Ballistic Performance of Hybrid Jute and
Aramid Reinforcing Graphite Nanoplatelets in High-Density Polyethylene Nanocomposites" is
being submitted for publication in Journal of Materials Research and Technology. On behalf of
my co-authors, I declare that there is no conflict of interest and the manuscript has not been
previously published or is currently under consideration for publication elsewhere. I also
confirm that all co-authors have read and approved the manuscript and, if accepted, it will not
be fully or partially published elsewhere without the written consent of the copyright holder.
The integration of natural lignocellulosic fibers and synthetic fibers for reinforcing polymeric
materials presents a compelling prospect for creating multifunctional hybrid materials and
structures with superior properties for advanced applications. Moreover, incorporating
graphite-based materials into these composites opens up opportunities for structural
functionalization and further property enhancement.
The present study focuses on optimizing the ballistic performance of newly developed
jute/aramid/high-density polyethylene hybrid nanocomposites by introducing graphite
nanoplatelets (GNP).
The number of citations for papers in this emerging field is encouraging, especially given its
novelty. One example is the graphene nanoplatelets-reinforced natural fiber composites, which
received 37 publications and over 245 citations between 2019 and 2022. This trend indicates a
rapidly expanding area of research that warrants further exploration and validates the
importance of the current investigation. Hence, this article has the potential to offer substantial
contributions to the field of Journal of Materials Research and Technology.

Best regards,
Ulisses Oliveira Costa
Military Institute of Engineering
Rio de Janeiro, Brazil
Response to Reviewers

Response to the Reviewers

Reviewer #1:
General comment: The manuscript has been written well but before acceptance
minor modification is required that has been given below:
Response: The authors would like to thank the Reviewers for the valuable
comments and suggestions on the structure and scientific aspects that contribute
to improve the manuscript. Amendments are provided accordingly. Responses to
each comment are listed below and all modifications/additions were marked as
Track changes in the revised version of the manuscript.

Comment (1): Authors need to thoroughly check the manuscript for grammatical
errors.
Response: The grammar of the manuscript is now revised in the improved
version of the manuscript. Whit the assistance of an English expert.

Comment (2): The introduction section has been written beautifully but need to
include recent published papers on nanocomposites like:

Effect of drilling process parameters on bearing strength of glass
fiber/aluminum mesh reinforced epoxy composites. Scientific Reports,
2023.

 Predicting kerf quality characteristics in laser cutting of basalt fibers


reinforced polymer composites using neural network and chimp
optimization. Alexandria Engineering Journal, 2022.

 Fabrication and Characterization of Nano-filled Polymer Composites. The


Egyptian International Journal of Engineering Sciences & Technology,
(2020).

 Preparation and characteristics of Cu-Al2O3 nanocomposite. Open


Journal of Metal, 2011.

Response: The authors would like to thank for the important background paper
suggestions for the improvement of the manuscript. In the revised version these
references are now included.

Comment (3): The discussion about reinforcement dispersion is not adequate.


The reviewer suggests the authors to discuss more about this. Some references
are suggested:

Predicting crystallite size of Mg-Ti-SiC nanocomposites using an adaptive
neuro-fuzzy inference system model modified by termite life cycle
optimizer. Alexandria Engineering Journal, 2023.

 Design and high efficient construction of bilayer NiCoO2/Poly(1-NA-co-oT)


nanocomposite absorber for X-band stealth applications. Vacuum, 2023.

 Mechanical performance of glass/epoxy composites enhanced by micro-


and nanosized aluminum particles. Journal of Industrial Textiles, 2021.

 The effects of nano-silica/nano-alumina on fatigue behavior of glass fiber-


reinforced epoxy composites. Journal of Composite Materials, 2017.

 Effect of through-the-thickness position of aluminum wire mesh on the


mechanical properties of GFRP/Al hybrid composites. Journal of Materials
Research and Technology, 2021.
Response: The authors agreed with the Reviewer’s comment. In the revised
version these references are now included, as well as a new discussion about
the dispersion of the nanofiller.

Comment (4): The length of Conclusion may be shorten.


Response: The Conclusion of the revised manuscript is now rewritten in
accordance to the Reviewer’s comment, which made it shot.

Reviewer #2:
General comment: This manuscript presented an interesting work about the
characterization and ballistic performance of hybrid HDPE nanocomposites. The
work has potential. However, some points listed below need to be improved.
Response: The authors would like to thank the Reviewers for the valuable
comments and suggestions on the structure and scientific aspects that contribute
to improve the manuscript. Amendments are provided accordingly. Responses to
each comment are listed below and all modifications/additions were marked as
Track changes in the revised version of the manuscript.
Comment (1): Abstract: add some numerical results in the abstract.
Response: The authors agreed with the Reviewer’s comment. The Abstract of
the manuscript is now rewritten with some numerical results.

Comment (2): Introduction: please clearer the novelty of this work.


Response: The authors fully agree with the Reviewer’s comment. In the new
version the innovative aspect of the present work is now included in the
introduction section.
Comment (3): Section 3.1: In XRD analysis I suggest add a table with the main
results (crystallinity, FWHM).
Response: The authors once again would like to thank the Reviewer for the
suggestion. In the revised version of the manuscript, the Section 3.1 is now with
a table highlighting the XRD parameters.

Comment (4): Section 3.1: in this section the authors observed an increase in
crystallinity but a decrease in crystal size after GNP addition. Please better
explain/discuss this interesting behavior in section 3.1.
Response: The authors understand the important point raised by the reviewer
and a better explanation in this regard is now included in the new version of the
manuscript.

Comment (5): Section 3.5: the authors said that: "at room temperature (RT=25
°C), the E' values for composites with low GNP concentrations, such as 0.10 wt%,
reached the highest storage modulus value (2.53 GPa). This represents an
increase of over 30% compared to the 20J/HDPE composite (1.90 GPa)." I
suggest better describe this result. What cause this increase in E' value?
Response: The Reviewer raised another important point, which deserves
attention. In the revised version of the manuscript, this result is now better
discussed and explained.
Revised Manuscript File Click here to view linked References

Characterization and Ballistic Performance of Hybrid Jute and Aramid


Reinforcing Graphite Nanoplatelets in High-Density Polyethylene
Nanocomposites

Ulisses Oliveira Costa1,*, Fabio da Costa Garcia Filho1 , Teresa Gómez-del Río2,
Édio Pereira Lima Júnior1, Sergio Neves Monteiro1, Lucio Fabio Cassiano Nascimento 2
1
Materials Science Department of Military Institute of Engineering (IME), Rio de Janeiro 22290-270, Brazil;
fabiogarciafilho@gmail.com (F.d.C.G.F.); edio@ime.eb.br (É.P.L.Jr.); snevesmonteiro@gmail.com (S.N.M.);
lucio@ime.eb.br (L.F.C.N.)
2Durability
and Mechanical Integrity of Structural Materials Group (DIMME), School of Experimental Sciences
and Technology, Rey Juan Carlos University, C/Tulipán, s/n. Móstoles, 28933 Madrid, Spain;
mariateresa.gomez@urjc.es Field Code Changed
* Correspondence: ulissesolie@ime.eb.br; Tel.: +55 -21982418125 Field Code Changed

Abstract: Hybrid nanocomposites have emerged as a promising solution for


engineering applications associated with reduced costs and weight aiming to enhance
performance in special areas such as ballistic protection. This study delves into the
development of hybrid nanocomposites featuring a high-density polyethylene (HDPE)
matrix modified with graphite nanoplatelets (GNP) and reinforced by aramid and jute
fabrics. The incorporation of GNP significantly influences the crystalline structure of
the GNP/HDPE matrix, as evidenced by Raman and X-ray diffraction (XRD) analyses.
Furthermore, it assumes an important role in modifying the crystallization and glass
transition temperatures (Tc and Tg) and influencing the dynamic mechanical behavior.
Specifically, it increases viscoelastic stiffness, raises the storage modulus by more
than 30%elevates the storage modulus, and reduces the tan(δ) value. In addition,
replacing 5 layers of jute fabric with an equivalent number of aramid layers maintains
comparable ballistic performance to a 20-layer aramid nanocomposite. This
substitution yields a remarkable 659.41 J of absorbed energy and a limit velocity of
405.72 m/s. Additionally, a hybrid nanocomposite comprising 10 layers each of jute
and aramid showcases impressive ballistic resistance against 9 mm caliber
ammunition, achieving 419.84 J of absorbed energy and a limit velocity of 320.13 m/s.
Scanning electron microscopy (SEM) analysis exposes the intricate fracture
mechanisms, encompassing phenomena such as crazing, GNP/HDPE fibrillation, fiber
rupture, and debonding. These mechanisms are significant in the composite's energy
absorption during a projectile impact. Moreover, a cost-benefit analysis underscores
the potential of hybrid nanocomposites in ballistic protection by simultaneously
reducing both helmet weight and cost in 7% and 40%, respectively.

Keywords: Jute, aramid, graphite nanoplatelet, high-density polyethylene, hybrid


nanocomposite

1. Introduction
The technological development of polymers composite materials for engineering
applications serves a clear and fundamental purpose, namely, to reduce the cost and
weight of products [1-5], making them more competitive compared to metallic and
ceramic materials available in the market. This evolution benefits various industrial
sectors, such as construction, automotive, aerospace, and defense [1-5]. In particular,
the use of composites in defense materials, like ballistic armor for personal protection,
is crucial to ensure the safety and efficiency of vests and helmets [5].
In the specific context of ballistic vests and helmets, the predominant materials
are aramid and ultra-high molecular weight polyethylene (UHMWPE). Aramid is used
in the form of fabrics such as the commercial brands Kevlar® and Twaron®, produced
respectively by Dupont and Teijin. UHMWPE can be employed in the form of plates or
fabrics, resulting in products like Dyneema® or Spectra®. These materials are globally
recognized as the gold standard for protection against firearm shootings, blade stab,
explosion, and other impact events, thanks to their excellent mechanical properties,
high impact resistance, and energy absorption capacity [5-11]. Another favorable
characteristic of UHMWPE is its low density, making it floatable in water and,
consequently, providing greater comfort and agility to the user.
Additionally, a synthetic material like aramid fiber or its related fabric can undergo
various surface treatments and modifications to enhance properties and compatibility
with the polymeric matrices that bind to form composites. Some treatment methods
include chemical grafting (with acid or base), physical methods (like gamma radiation,
plasma, and ultraviolet), coating methods (using, for example, Polyethylene glycol-
PEG or epoxy resin), and nanostructural modifications (using nanoparticles like ZnO,
SiO2, TiO2, graphene, among others) [12-16].
Improving the surface of aramid fiber aims to enhance its chemical activity and
increase its roughness, enabling an excellent chemical bond and mechanical
interlocking between the fiber and the matrix. This process facilitates the interaction
between the fiber and the polymeric matrix, reducing the probability of failures such as
delamination, debonding, and slippage, while increasing the lifespan and safety of
composites [12].
However, it is essential to note that these cutting-edge materials come with a
relatively high cost, and additional treatments can further escalate production
expenses. Aware of this challenge, some researchers are focusing their efforts on
developing more cost-effective and environmentally viable hybrid composites, which
combine synthetic fibers with natural fibers for engineering applications [17-20].
A significant example is the study conducted by Meliande et al. [17], which
investigated the potential advantages of partial replacing aramid by curaua, a natural
lignocellulosic fiber (NLFs) in the tensile and impact properties of laminated hybrid
composites. Their results indicated that replacing 9 and 14 layers of aramid fabric by
2 and 3 layers of curauaá mat, resulted in a significantly advantageous reduction in all
tensile properties. Specifically, composites with approximately 33 vol.% volume of
aramid (10 layers) and 32 vol.% volume of curauaá mat (2 layers) in a matrix with about
35 vol.% volume of epoxy were considered the most promising for application in
ballistic helmets.
Similarly, the studies by Costa et al. [19] revealed that a hybrid nanocomposite
composed of 10 layers of NLFnatural jute fiber fabric (Corchorus capsularis) and 10 Formatted: Font: Italic
layers of aramid fabric, added with 0.10 wt.% of graphene nanoplatelets (GNP),
achieved a 2433% increase in mechanical toughness, 591% increase in tensile
strength, and a 462% reduction in ductility compared to plain composites containing
50% vol% of jute fiber reinforcing a high-density polyethylene (HDPE) matrix. These
results further highlight the potential of hybrid GNP added nanocomposites.
These novel hybrid nanocomposites have shown promising performance for
engineering applications [20]. NLFs, such as curaua and jute, are considered
promising candidates due to their many advantages, such as low density,
biodegradability, abundant availability, good damping properties, machinability, low
cost, ease of separation for recycling, enhanced energy recovery, and CO2 neutrality
[21]. Jute fiber or fabric is widely studied and recognized globally, making it a notable
choice for reinforcing polymer composites in various industries, including graphene-
based nanocomposites [20-24]. In this context, the present work investigated for the
first time the development of hybrid nanocomposites with a HDPE matrix functionalized
with GNP and reinforced with aramid and jute fabrics. The innovative aspect of the
present work is the combination of GNP functionalization, aramid, and jute
reinforcement within a HDPE matrix, exploring the potential synergies among these
elements. This exploration opens up new avenues for the production of hybrid
nanocomposites that are not only efficient in terms of ballistic protection but also
sustainable and cost-effective. The emphasis on sustainability is particularly
noteworthy, considering the environmental impact and cost challenges associated with
cutting-edge materials in the field of ballistic armor, specifically for helmets.

2. Methods
2.1. Materials and specimen preparation
From HDPE HE150 grade pellets sourced from Braskem [25], Brazil, and GNP
powder obtained from UCSGraphene, Brazil, a masterbatch (concentrate) was
produced through mechanical agitation, with an initial GNP concentration of 4.3 wt.%
by weight. Following this, three dilutions were conducted to acquire HDPE fractions
with GNP mass fractions of 0.10%, 0.25%, and 0.50 wt.%. Each dilution was executed
using the extrusion process in an interpenetrating and co-rotating twin-screw extruder,
Tecktril model DCT-2.
The extrusion conditions, as defined by Escocio et al. [26], were adopted,
including a screw rotation speed of 300 rpm, a feeder rotation speed of 15 rpm, and
tailored temperature profiles for the processing zones: 90°C for the first zone, 140°C
for the second to fifth zones, and 160-180°C for the sixth to ninth zones.
Subsequently, films with a uniform thickness of 300 μm were manufactured for
each condition using the hot compression molding technique. The processing took
place at a constant temperature of 170°C, employing a high-precision thermal press
Solab SL-11/15, Brazil. The films produced were associated to the following
nomenclature: 0.10%GNP/HDPE, 0.25%GNP/HDPE, and 0.50%GNP/HDPE.
In order to produce the hybrid nanocomposite plates, a laminating scheme was
adopted, where layers of fabrics were interleaved with HDPE films. For this purpose,
it was used an aramid fabric, commercially named Twaron®, provided by Teijin
Aramid, Brazil. The commercial grade used in the present work was the 410 g/m²
ballistic fabric, CT 736, with basket 2x2 weave. In addition, jute fabric was obtained
from the Brazilian manufacturer Sisalsul. According to Costa et al. [19], the jute fiber
fabric density was considered as 1.30 g/cm³. This parameter was important for
estimating the jute fiber fabric volume fraction. The as-received jute fabric consisted of
a simple weave, and it was cut to 120 × 120 mm dimensions and carefully dried in an
oven at 60 °C for 24 h to remove moisture. Consequently, it was manufactured hybrid
nanocomposites were manufactured by incorporating a combination of aramid and jute
fabrics reinforcement with a volume fraction of 50 vol.% content into the HDPE matrix,
resulting in plates measuring 120 x 120 x 10 mm. The process involved the use of 20
layers of fabric and 21 layers of HDPE films. During processing, the pressure was
gradually increased, adding one ton at a time, with each new pressure being
maintained for 1 min and followed by 30 s of degassing, until reaching a pressure of
13 tons. The cooling of the plates was conducted outdoors, allowing them to cool to
room temperature (RT) for demolding.
The Personnel Armor System for Ground Troops (PASGT) helmet was used as
a reference for the Jute/Aramid/GNP/HDPE hybrid nanocomposites. In the current
study, the number of layers corresponding to different reinforcements of synthetic and
natural fibers is a crucial aspect directly related to the hybrid composite design. The
traditional PASGT ballistic helmet has a final thickness of 8-10 mm [17,27], which is a
key criterion for achieving a density of 10 to 12 kg/m 2, equating to a stack weight of
approximately 200 g, and ensuring user comfort. Consequently, 20 layers of aramid
fabric were employed to simulate the PASGT helmet, representing a 100 vol.%
synthetic fiber composite.
Aiming to achieve cost-effectiveness, a strategic approach was adopted by
adjusting the number of aramid and jute layers in the hybrid nanocomposites as shown
in Table 1. The implementation followed a well-defined ratio: 20:0, 15:5, 10:10, and
5:15, corresponding to reduction fractions of 50%, 37.5%, 25%, and 12.5 vol.% by
volume of aramid in the composites, respectively. The nomenclature of the composites
is detailed in Table 1. Group 1 aims to analyze the influence of GNPs on the HDPE
matrix, while group 2 focuses on investigating the impact of reducing the aramid layers
in the hybrid nanocomposite. Table 1 the 0.X%p GNP indicates the best results
obtained from the group 1.

Table 1. Hybrid nanocomposites configurations and associated nomenclatures


Group Composites Nomenclature
50 volvol.% Jute + HDPE 20J/HDPE
50 volvol.% Jute + 0.10%p GNP + HDPE 20J/0.10%GNP/HDPE
1
50 volvol.% Jute + 0.25%p GNP + HDPE 20J/0.25%GNP/HDPE
50 volvol.% Jute + 0.50%p GNP + HDPE 20J/0.50%GNP/HDPE

50 volvol.% Jute + 0.X%p GNP + HDPE 20J/0.X%GNP/HDPE


50 volvol.% Aramid + 0.X%p GNP + HDPE 20A/0.X%GNP/HDPE
2 37.5 volvol.% Jute + 12.5 volvol.% Aramid + 0.X%p GNP + HDPE 15J/5A/0.X%GNP/HDPE
25 volvol.% Jute + 25 volvol.% Aramid + 0.X%p GNP + HDPE 10J/10A/0.X%GNP/HDPE
12.5 volvol.% Jute + 37.5 volvol.% Aramid + 0.X%p GNP + HDPE 5J/15A/0.X%GNP/HDPE

2.2. X-ray diffraction analysis (XRD)


XRD analysis was used to determine by means of diffraction patterns, the
crystallinity, and the crystallite size of GNPs, HDPE, 0.10%GNP/HDPE,
0.25%GNP/HDPE, and 0.50%GNP/HDPE. Through the diffractograms, it is possible
to assess the quality of the polymer processing.
The equipment used for this analysis was a Panalytical X'pert Pro diffractometer
with a Co (cobalt) anode, and a scintillation counter detector (NaI), and operated at a
power of 40 mA x 40 kV. The scan range for the analysis was from 10° to 80°, using a
θ-2θ configuration, Rio de Janeiro, Brazil.
The broadening of the reflection profile in XRD pattern was used to calculate the
crystallite size, and the corresponding peak position is utilized to determine the
interlamellar spacing. The average crystallite size (Lc) (crystalline dimension along the
c-axis) of the GNPs was determined using the Scherrer equation [28,29], Eq.uation
(1), and the d002 spacing (d-spacing for the (002) plane) was calculated using Bragg's
Law [28,29], Equation Eq.(2).
𝐾𝜆
𝐿𝑐 = (1)
𝛽𝑐𝑜𝑠𝜃
where: Lc is the average crystallite size along the c-axis, K is the Scherrer constant
(typically taken as 0.9), λ is the X-ray wavelength used for the analysis, β is the full
width at half maximum (FWHM) of the diffraction peak in radians, θ is the Bragg angle
(diffraction angle) in radians.
𝑛𝜆 = 2𝑑𝑠𝑖𝑛𝜃 (2)
where: d is the spacing between crystallographic planes (in this case, the (002) plane),
n is the order of diffraction, λ is the X-ray wavelength used for the analysis, θ is the
Bragg angle (diffraction angle) in radians.
The crystallinity index was calculated by the Segal et al. [30] method [30],
considering the ratio of the crystalline area and the amorphous area of the
diffractogram.
2.3. Raman spectroscopy
Raman spectroscopy was conducted on GNPs, plain HDPE, and HDPE
nanocomposites reinforced with GNPs under different conditions: 0.10%GNP/HDPE,
0.25%GNP/HDPE, and 0.50%GNP/HDPE. The Raman spectra were obtained in
backscattering geometry at room temperatureRT using a Raman spectrometer
equipped with an Andor Shamrock spectrometer, a charge-coupled device (CCD) iDus
detector, a 488 nm (2.54 eV) laser, and an optical system, Rio de Janeiro, Brazil.
All measurements were taken using an optimal laser spot diameter of 1 μm and
a power of 1 mW. The spectral resolution of the spectrometer for this configuration was
determined using a silicon wafer peak at 520 cm −1, adjusted using a Gaussian line
shape with a full width at half maximum (FWHM) of 4 cm −1. The Raman spectra were
obtained using Origin Pro data analysis software.
2.4. Transmission electron microscopy (TEM)
The morphology and crystal structure of GNPs were analyzed using high-
resolution transmission electron microscopy (HR-TEM) and selected area electron
diffraction (SAED) techniques. The analyses were carried out using a JEOL 2100F
electron microscope, Brazil. The microscope was equipped with a CMOS camera and
operated with an acceleration voltage of 200 kV.
In HR-TEM mode, the electron microscope provides high-resolution images that
allow researchers to observe the detailed morphology and atomic arrangement of the
GNPs. On the other hand, SAED is a technique used to investigate the crystallographic
structure and lattice planes of materials by focusing the electron beam to a small area
on the sample surface, which produces diffraction patterns. These diffraction patterns
provide valuable information about the crystallographic orientation and lattice spacing
of the GNPs.
2.5. Scanning electron microscopy (SEM)
SEM technique was performed using a Quanta FEG 250 scanning electron
microscope, FEI, Brazil. The objective was to observe the relationship between the
ballistic performance of groups 10 and 21 with their microstructure, aiming to identify
the fracture mechanisms that occurred after ballistic and mechanical testing of the
structural composites and individual layers.
For the metallic coating of the samples, a high-vacuum metal film deposition
equipment from LEICA, model EM ACE600, was used.
2.6. Differential scanning calorimetry (DSC)
For the DSC analysis of the nanocomposite films, it was utilized an aluminum
crucible in a TA Instruments Q1000 calorimeter located in Brazil. The DSC equipment
operated under a nitrogen atmosphere with heating rates of 10ºC/min, within a
temperature range fromof 20 to 200ºC. Through this test, it was possible to calculate
the crystallization temperature (Tc) and melting temperature (Tm) of the
nanocomposites, as well as the degree of crystallinity of the composites, from the heat
flow versus temperature curves.
Aiming to determine the degree of crystallinity, Equation Eq.(3) was used,
following the approach described by [31,32].
∆𝐻𝑚
𝑋𝑐 = × 100 (3)
[∆𝐻0 × (1 − ∅)]
In which ΔH corresponds to the heat of fusion of the sample, ΔH 0 = 293 J/g refers
to the heat of fusion for 100% crystalline HDPE [32] and ϕ refers to the weight fraction
of the nanofiller in each sample. Additionally, a non-isothermal crystallization kinetics
analysis of HDPE and GNP/HDPE nanocomposites was carried out at the following
cooling rate: 10 °C/min, aiming to evaluate the effect of GNPs on the HDPE matrix.
2.7. Dynamic Mechanical Analysis (DMA)
DMA tests were performed according to ASTM D7028 – 08 [33], using a Q800
TA Instruments equipment. The test was conducted at a frequency of 1 Hz and a
temperature range from -140 to 150 °C with a heating rate of 5 °C/min under a nitrogen
atmosphere. Storage modulus (E'), loss modulus (E''), and tangent delta (tan δ) curves
were recorded for the samples listed in Table 1. DMA test used samples with
dimensions of 37 x 12 x 2 mm for the three-point bending mode. Through the E', E'',
and tan δ curves, it was possible to identify the viscoelastic behavior and glass
transition of the materials, as well as the influence of the GNPs [34].
2.8. Residual Velocity
To evaluate the ballistic performance of the nanocomposites and individual plates
forming the layers of the structural composites, a non-standard test was employed.
This test involves measuring the residual velocity at the Brazilian Army Evaluation
Center (CAEX). For this test, 9 mm caliber ammunition with a mass of 8.0 g, and 420
m/s impact velocity was used. The test utilizes a Doppler radar, consisting of two
optical barriers to measure impact and residual (after impact) velocities for each
shooting. The sample was positioned 2.75 m from the second radar barrier. With the
calculated velocities, it is possible to determine the variation in momentum (Δq) and
kinetic energy (ΔK) using the equations below:
∆𝑞 = 𝑚(𝑉𝑖𝑚𝑝 − 𝑉𝑟𝑒𝑠 ) (4)
1
∆𝐾 = 2 𝑚(𝑉𝑖𝑚𝑝 2 − 𝑉𝑟𝑒𝑠 2 ) (5)
where: m is the mass of the projectile (9 mm caliber ammunition) in kilograms (kg),
Vimp is the impact velocity of the projectile in meters per second (m/s),. This iswhile
corresponding to the velocity at which the projectile strikes the sample during the
ballistic test, Vres is the residual velocity after impact in meters per second (m/s)., This
iscorresponding to the velocity of the projectile after passing through the sample during
the ballistic test.
In these equations: Δq represents the change in momentum of the projectile due
to the impact and penetration through the sample, ΔK represents the change in kinetic
energy of the projectile during the ballistic event.
2.9. Cost-effectiveness
The assessment of cost-effectiveness involved a comprehensive analysis of
various factors, encompassing the expenses incurred in acquiring the materials for
fabricating the composite plates, the overall weight of these materials, and the specific
dimensions associated with the conventional PASGT helmet. Notably, the PASGT
helmet incorporates 19 layers of aramid fabric, resulting in a thickness range of 8 - 10
mm, an area of protection spanning 0.14 m², and an aerial density measuring 11.20
kg/m².
3. Results and discussions
3.1. XRD analysis
XRD analysis of GNPs, HDPE, and HDPE/GNP nanocomposites is depicted in
Figure 1 and highlighted in Table 2. In the XRD pattern of GNPs, it is observed sharp
peaks at 2θ ≈ 30.69°, along with relatively weaker peaks at 2θ ≈ 51.96° and 64.18°.
These correspond to carbon reflections (002), (100), and (004), respectively, as
detailed in reference [31]. The prominent peak at 2θ ≈ 30.69° unequivocally confirms
the presence of GNP hexagonal nanoplatelets, and its d-spacing was calculated as
0.338 Å using Bragg's law.

Figure 1. XRD chromatogram of GNPs and GNP/HDPE nanocomposites

For pure HDPE, the XRD revealed the presence of two prominent peaks at 2θ =
25.16° and 2θ = 28.10°, corresponding to the (110) and (200) reflections of the
orthorhombic HDPE phase. Additionally, there are two weaker peaks at 2θ = 35.21°
and 2θ = 42.51°, which can be attributed to the (210) and (020) reflection planes,
respectively, as detailed by Xiang et al [35]. The XRD patterns of GNP/HDPE
nanocomposites displayed intriguing trends. In fact,: the intensities of the peaks at 2θ
≈ 25.16° and 2θ ≈ 28.10° decreased as the GNP content increased, while the
intensities of the peaks at 2θ ≈ 35.21° and 2θ ≈ 42.51° showed an opposite trend. This
indicates a complex interplay between HDPE and GNPs in these nanocomposites.
Formatted: Indent: First line: 0"

Table 2. Summary of the pic’s parameters of the nanocomposites


Material Pic Intensity FWHM Crystallinity
2Theta (°) (counts) (%)
30.69 54678.5 0.27
GNP 51.96 1960.6 0.46 100.00
64.18 2334.8 0.49
25.20 78598.4 0.30
28.10 18590.6 0.38
HDPE 57.45
35.21 888.5 0.37
42.61 1554.6 0.44
25.30 84634.9 0.33
28.10 13365.3 0.40
31.12 4424.0 0.21
0.10%GNP/HDPE 58.60
35.32 695.1 0.33
42.61 1624.2 0.47
51.81 341.9 0.48
25.30 72471.8 0.36
28.10 10193.0 0.43
31.12 24231.9 0.22
0.25%GNP/HDPE 60.17
35.32 516.0 0.42
42.61 1255.8 0.48
51.81 330.0 0.51
25.30 90894.1 0.28
28.10 12733.2 0.36
31.12 43299.3 0.16
0.50%GNP/HDPE 35.32 735.1 0.33 65.43
42.61 1725.8 0.43
51.81 469.7 0.43
64.18 1305.5 0.30

To quantify the crystallinity, it was calculated the crystalline fraction by


determining the ratio of the area under the crystalline peaks to the total area under the
XRD spectrum for each nanocomposite, as described by Segal et al [30].
Consequently, the crystallinity fraction increased with higher GNP content from: HDPE
(57.45%), 0.10% GNP/HDPE (58.60%), 0.25% GNP/HDPE (60.17%), and 0.50%
GNP/HDPE (65.43%). These results suggest that GNPs serve as nucleating agents,
promoting nucleation in the vicinity of the GNPs.
To assess crystalline size, it was analyzed the broadening of the XRD pattern's
reflection profile and used the corresponding peak position to calculate the
interlamellar spacing. The average crystalline size (Lc) of GNPs using Scherrer's
Equation (1) was determined, and the d002 (spacing between (002) planes) was derived
from Bragg's law (Equation (2)).
Here, θ represents the Bragg angle, β is the full width at half maximum (FWHM)
(0.004712 rad), λ is the X-ray wavelength (0.1789010 nm), and κ is the shape factor
(0.91). The average size of the crystalline domains was found to be 35.83 nm, which
is notably larger than the value calculated by Albetran [28] but consistent with TEM
observations. Furthermore, based on the approach by Seehra et al. [29], the number
of layers along the c-axis (Nc) was calculated as the ratio of the crystal size to the
interplanar spacing. In this study, Nc was determined to be 106 layers, suggesting that
the majority of the particles consist of graphite-like structures rather than graphene.
However, XRD may not fully distinguish between various structures present in GNPs,
such as graphite, graphene, graphene oxide, rolled and semi-rolled sheets, etc. In the
case of HDPE and its nanocomposites, it was also calculated the crystallite sizes using
Equation Eq.(1). The results indicated that crystal sizes decreased with increasing
GNP content from: HDPE (31.11 nm), 0.10% GNP/HDPE (28.37 nm), 0.25%
GNP/HDPE (26.13 nm), and to 0.50% GNP/HDPE (33.32 nm). This observation
indicates that GNPs play a role in influencing the nucleation phase during the HDPE
crystallization process. Nevertheless, with an increase in GNP content, there is a
simultaneous escalation in agglomerate formation. This occurrence could hinder the
growth process of HDPE crystals by diminishing the superficial area and intensifying
steric hindrance from the GNPs. Consequently, this leads to the formation of a greater
number of smaller crystals [19,31].
3.2. Raman analysis
In Figure 2, the Raman spectral range spanning from 200 to 4000 cm −1 reveals
significant bands falling within the 1400 to 1480 cm −1 range. These bands are indicative
of methylene bending vibrations (δ(CH2)). Figure 2 provides further insight, clearly
demonstrating that the 1440 cm −1 band, initially identified as the D band in the
nanoplatelets, has undergone a noticeable shift compared to the same band when it
exists in isolation.

Figure 2. Raman Spectra of GNP and GNP/HDPE nanocomposites

Several references, including Fischer, Wallner, and Pieber et al. [36], Kelkar et
al. [37], Nabiyev et al. [38], and Silva and Wiebeck [39], have collectively emphasized
the importance of the 1418.7 cm −1 band, often referred to as the "crystallinity band."
This specific band serves as a distinctive marker for the degree of orthorhombic
crystallinity in HDPE samples, and its significance has been well-established by Xiang
et al. [35].
Within the spectral data, other notable bands include those at 1060 and 1367.3
cm−1, which are attributed to the amorphous phase. Additionally, Raman peaks at
1060.8 and 1127.7 cm−1 correspond to symmetric and asymmetric stretching vibrations
of C–C bonds, while peaks at 2849.7 cm−1 and 2883.8 cm−1 indicate stretching of (CH3)
groups. Furthermore, the bands at 2907.7 cm−1 and 3068.5 cm−1 are assigned to the
(C−H group) [35-38].
In the Raman spectra of graphite-derived materials, four primary bands are
observed. The first and most prominent is the G band, peaking at around 1580 cm−1.
This band represents the E2g vibrational mode associated with the stretching of C-C
bonds within crystalline planes, serving as a hallmark of its graphitic nature.
The second, the D band, with a peak near 1350 cm −1, corresponds to vibrational
modes of carbon atoms with sp3 hybridization. This feature signifies defects in the
hexagonal structure, encompassing vacancies, pentagon, and heptagon rings as well
as edge effects, and functional groups. The ratio between the D and G band intensities,
termed the ID/IG ratio, serves as a parameter to gauge structural disorder in
carbonaceous materials. For the studied material, a low ID/IG ratio of 0.10 is derived, in
line with the findings of Moriche et al. [40] and an indicative of minimal structural
defects.
3.3. Morphology analysis
SEM and TEM analyses were conducted on the GNPs to elucidate the nanofillers'
structure. In Figures 3(a) and (b), a SEM image of the GNPs is displayed, revealing
that the agglomerates' average lateral size measures below 10 μm. Furthermore, these
agglomerates comprise over 10 layers, thereby validating the graphitic nature of the
GNPs.

(a) (b)
Figure 3. SEM images of the GNPs: (a) Lower magnification 5000x, and (b) higher magnification
30000x.
Figure 4 presents TEM images depicting the crystalline domains of the GNPs.
Figure 4(a) reveals a 20 nm single crystal, aligning well with the average crystal size
calculated by XRD. Moreover, Figures 4(b) and (c) showcase the distinctive sheet
morphology, indicating varying degrees of rolling; ranging from semi-rolled to fully
rolled. These images provide affirmation of a mixed array of structures, as inferred from
XRD analysis by Costa et al. [19].
(a)

(b) (c)
Figure 4. HR-TEM images of the GNPs (a) agglomerate of GNPs; (b) presents the GNP
monocrystalline structure semi-rolled; (c) displays GNP completely rolled

Figures 5(a) to (d) illustrate the evolving surface characteristics of GNP/HDPE


nanocomposite films at varying nanofiller concentrations. Notably, the film roughness
demonstrates a proportional increase with the augmentation of GNP content. Despite
the incorporation of GNPs through a co-rotating twin-screw extruder process, the
dispersion of the nanofiller remains non-uniform. This significant morphological
variation holds potential implications for the dynamic behavior of the hybrid
nanocomposites, and can be attributed to steric hindrance induced by GNPs [41-46].
Importantly, these factors play an important role in influencing the interfacial adhesion
between the GNP/HDPE matrix and the reinforced hybrid structure, consequently
impacting mechanical and ballistic behavior [41].
(a) (b)

(c) (d)
Figure 5. SEM images of the GNP/HDPE nanocomposites: (a) HDPE, (b) 0.10%GNP/HDPE, (c)
0.25%GNP/HDPE, and (d) 0.50%GNP/HDPE.

3.4. DSC analysis


DSC analyses curves were performed on HDPE and GNP/HDPE composites
with varying GNP concentrations (0.10 to 0.50 wt.%) at a cooling rate of 10 °C/min.
The results revealed crystallization (Tc) and melting (Tm) temperatures, as shown in
Figure 6. It was observed that, as the GNP concentration increased in the composite,
Tc shifted on average by 1 °C to higher temperatures compared to pure HDPE. As
highlighted by Evgin et al. [31], this outcome indicates a nucleation effect caused by
the graphite load in the polymer matrix. On the other hand, the Tm did not change with
the GNP amount. Furthermore, the crystallization of the nanocomposites with GNPs is
influenced by the clustering of these nanoparticles, which might explain the observed
rise in the crystallization temperature as the graphite quantity increases. The crystalline
phase of these units, being less favored, hampers the regular packing of HDPE
polymer chains, as pointed out by Evgin et al. [31].
(a) (b)

Figure 6. DSC curves of 10°C/min: (a) cooling curve showing the exothermic pic representing the
crystalline temperature, and (b) heating curve showing the endothermic pic representing the melting
temperature.

Using the heating curves of the HDPE and GNP/HDPE films, the degree of
crystallinity for all composites was calculated using Equation (3). The obtained values
were as follows: HDPE (50.17%), HDPE/0.10%GNP (62.19%), HDPE/0.25%GNP
(63.71%), HDPE/0.50%GNP (75.10%). These results are closely aligned and follow
the same trend as the results obtained by XRD.

3.5. DMA analysis


DMA tests were conducted for all the composites. The curves for storage
modulus (E’), loss modulus (E”), and tangent delta (tan (δ)) are shown in Figure 7 for
20J/HDPE and 20J/GNP/HDPE composites with different GNP loadings. Figure 7(a)
illustrates the variation of E’ with temperature for the investigated composites. For both
E’ results, without GNP and with GNP, the values decreased with temperature.
However, for the composites with GNP incorporated into the HDPE matrix, they
remained higher than those with for pure HDPE composites. This indicates that GNP
reinforcement improved the viscoelastic stiffness of the composite across a
temperature range of -140 to 150 °C.

(a) (b)
(c)

Figure 7. DMA test results for the 20J/HDPE and 20J/GNP/HDPE composites, depicting the variation
of (a) storage modulus (E'), (b) loss modulus (E"), and (c) tan δ with temperature.

Furthermore, at room temperature (RT=25 °C), the E’ values for composites with
low GNP concentrations, such as 0.10 wt.%, reached the highest storage modulus
value (2.53 GPa). This represents an increase of over 30% compared to the 20J/HDPE
composite (1.90 GPa). The observed outcome can be ascribed to the reinforcing
impact of GNP, which induces a limitation in the mobility of HDPE chains, ultimately
leading to a strengthened interphase [40]. Additionally, the E’ values obtained in this
study align with those reported by other authors previously [40,41]. Lastly, it can be
observed that with an increase in GNP loading in the HDPE matrix, the E’ values tend
to decrease, approaching those of pure HDPE.
Another notable aspect is the sudden drops in E’ values, indicated by peaks in
Figure 7(b), occurring at -110 and 60 °C for all the curves. The first peak corresponds
to γ relaxation, marking the end of the glassy behavior region of HDPE. This drop could
potentially be linked to the transition to a rubbery amorphous region, characteristic of
increased mobility of chain segments in the amorphous phase of the polymer matrix,
i.e., the Tg. The second abrupt drop, α relaxation, is related to the thermal stability of
the composites, namely, the mobility of segments within the crystalline phase, which
becomes more thermally resistant with GNP incorporation [41].
In Figure 7(b), it is worth noting that for all HDPE composites reinforced with GNP,
the E” values were higher than those of the 20J/HDPE composite and were slightly
shifted towards higher temperatures. Indeed, E” represents a viscous response of the
material and is indicative of its tendency to dissipate applied mechanical energy [47].
In the case of composites, E” is often described as internal friction and should be
greater as the volume fraction of reinforcement increases [48]. This results in a larger
interfacial surface area between the load/polymer matrix. According to Mohanty et al.
[47], the molecular movement of the polymer due to structural heterogeneities like the
load interface would directly increase internal friction and, consequently, the value of
E”. However, with the increase in the amount of GNP in the HDPE matrix, the formation
of more agglomerates reduces the surface area between reinforcement and matrix. As
can be observed, the highest E” values were obtained for composites with low GNP
quantities (≥ 0.25 wt.%).
When it comes to composites, E” is often described as internal friction and tends
to be higher with a larger volume fraction of reinforcement [41]. This results in a greater
interfacial surface area between the load and polymer matrix. As Mohanty et al. [47]
indicate, the molecular motion of the polymer due to structural heterogeneities like the
load interface would directly increase internal friction and, consequently, the value of
E”. However, with the increase in the amount of GNP in the HDPE matrix, more
agglomerates are formed, which reduces the surface area between reinforcement and
matrix. Notably, the highest E” values were observed for composites with lower GNP
quantities (≥ 0.25 wt%).
Figure 7(c) depicts the variation of tanδ with temperature for the studied
composites. The tanδ represents the relationship between viscous energy dissipation
and elastic energy storage per cycle (tanδ = E′′/E′), and it is's a dimensionless number
linked to the material's mechanical damping. As Saba et al. [48] explain, a high tanδ
value indicates a material with significant deformation capacity.
Moreover, the temperature at the tanδ peak is considered the material's dynamic
Tg. According to the inset of Figure 7(c), the tanδ value displayed an increasing trend
as the GNP concentration in the HDPE matrix increased. It shifted from -113.80 °C to
-110.22 °C for the composite with 0.50% GNP. This highlights the stiffer behavior of
composites with GNP incorporated into the HDPE matrix.
In comparison to the 20J/HDPE composite, the GNP-functionalized composites
exhibited peaks of tanδ with lower magnitudes, indicating a decreasing trend as the
incorporated GNP quantity increased. This behavior suggests that the 20J/HDPE
composite has a weak interfacial bond between the fibers and the matrix, allowing it to
dissipate more energy, resulting in a higher damping peak magnitude compared to a
material with a strongly bonded interface. However, after HDPE functionalization with
GNP, the magnitude of tanδ peaks decreased, signifying an improvement in the
interfacial bond and a subsequent reduction in energy dissipation capacity.
Figure 8(a) illustrates that as the number of layers of Twaron®™ aramid fabric
increases, the E’ values at 35 °C decrease. This could be due to the greater flexibility
of aramid fabric compared to jute fabric [47]. For instance, comparing the E’ values of
20J/0.10%GNP/HDPE nanocomposites and 20A/0.10%GNP/HDPE nanocomposites,
a reduction of over 70% can be calculated. Additionally, the same two relaxation
phenomena observed in the E’ curves of Figure 7(a) were observed.

(a) (b)
(c)

Figure 8. DMA test results for the J/A/GNP/HDPE composites, showing the variation of (a) storage
modulus (E'), (b) loss modulus (E"), and (c) tan δ with temperature.

However, for the E” values found in Figure 8(b), a considerable increase is seen
compared to the non-aramid composites of Figure 7(b). The hybrid
10J/10A/0.10%GNP/HDPE nanocomposite reached a peak E” around 60 °C of 230.32
MPa, representing an increase of 16.47% compared to the 20J/0.10%GNP/HDPE
nanocomposite (197.75 MPa). Also, a peculiar behavior around 15 °C is observed for
the hybrid nanocomposites 10J/10A/0.10%GNP/HDPE and 20A/0.10%GNP/HDPE.
This might have occurred due to slipping or accommodation between the fabric layers
during the test.
Regarding the tan(δ) values shown in Figure 8(c), it is noticeable that concerning
the maximum value, the aramid-containing composites showed an increase compared
to the non-aramid nanocomposites of group 1. This points to an improvement in the
energy dissipation capacity of the HDPE/GNP/Jute/Aramid composites. In relation to
Tg, the hybrid 10J/10A/0.10%GNP/HDPE nanocomposite presented the highest value
at -110.40 °C, similar to the 20J/0.10%GNP/HDPE composite (-110.22 °C).
Additionally, with an increase in the number of aramid fabric layers, there was a
decreasing trend in Tg values for the hybrid nanocomposites.
3.6. Residual velocity results
Table 3 provides a summary of all the properties obtained from the residual
velocity test for all the composite conditions analyzed in this study. These properties
encompass the projectile's impact velocity, the residual velocity after impact, the
energy absorbed by the composite, and the limit velocity.

Table 3. Summary of properties obtained from the residual velocity test for all the composites studied
in this work.
Vimp VRes Eabs VLim
Material
(m/s) (m/s) (J) (m/s)
20J/HDPE 423.40±6.10 402.17±7.56 69.50±5.53 130.79±5.54
20J/0.10%GNP/HDPE 421.61±2.87 401.26±3.28 66.98±3.39 129.37±3.25
20J/0.25%GNP/HDPE 419.86±2.00 401.41±2.44 62.94±7.44 124.78±6.38
20J/0.50%GNP/HDPE 421.47±2.80 402.91±3.14 64.66±2.92 127.12±2.88
15J/5A/0.10%GNP/HDPE 421.87±5.74 377.66±6.13 141.36±9.81 187.90±6.50
10J/10A/0.10%GNP/HDPE(*) 422.59±2.20 262.94±74.93 419.84±146.45 320.13±55.60
5J/15A/0.10%GNP/HDPE(*) 418.54±5.25 158.01±42.24 659.41±55.83 405.72±17.54
20A/0.10%GNP/HDPE(**) 422.48±2.16 0 713.96±7.29 422.48±2.16
(*) Partial perforation occurred, (**) No perforation occurred

Regarding this last parameter, the absorbed energy values, they were utilized
along with the kinetic energy equation to determine the velocity required to prevent
composite perforation.
Table 3 also presents the results of the nanocomposites absorbed energy (Eabs) Formatted: Subscript
after the impact of a 9 mm caliber projectile for the nanocomposites without aramid
fibers. However,. In fact, no evidence of modifications caused by GNP incorporation in
HDPE was found. As a result, the ANOVA analysis demonstrated that the values of
residual velocities are statistically equivalent with a 95% confidence level, as the
calculated F value of 3.14 < critical F value of 3.24.
This finding suggests that composites composed of 20 layers of jute fabric are
not suitable for applications in ballistic helmets. This is due to the fact that the amount
of absorbed energy is significantly lower in comparison to the impact energy, rendering
them insufficient to provide the required protection.
However, as aramid fabric layers were incorporated, a notable increase in energy
absorption capacity was observed, as clearly evidenced in Table 3. This trend was
further corroborated by ANOVA, where the calculated F value (86.97) significantly
exceeded the critical F value (2.87). According to the Tukey test, Table 4, this
significant difference was observed only for hybrid nanocomposites with more than 10
layers, with an average difference (d.m.s) of 137.43 J.

Table 4. Tukey Test Results for Absorbed Energy Values of Hybrid Nanocomposites.
20J/
15J/5A/0.10%GN 10J/10A/0.10%G 5J/15A/0.10%GN 20A/0.10%GN
0.10%GNP/
P/HDPE NP/HDPE P/HDPE P/HDPE
HDPE
20J/
0.00 74.39 352.86 592.43 646.98
0.10%GNP/HDPE
15J/5A/0.10%GN
74.39 0.00 278.47 518.05 572.60
P/HDPE
10J/10A/0.10%G
352.86 278.47 0.00 239.57 294.12
NP/HDPE
5J/15A/0.10%GN
592.43 518.05 239.57 0.00 54.55
P/HDPE
20A/0.10%GNP/H
646.98 572.60 294.12 54.55 0.00
DPE

Furthermore, partial perforation was observed in some ballistic shootings for


hybrid nanocomposites containing 10 and 15 layers of aramid. On the other hand, none
of the shootings resulted in perforation for the 20A/0.10%GNP/HDPE composites with 20
layers of aramid, as indicated in Table 3. In this context, compared to the
20J/0.10%GNP/HDPE composite, the 10J/10A/0.10%GNP/HDPE composite exhibited
a remarkable 526.81% increase in absorbed energy.
As a result, the limit velocity values of the hybrid nanocomposites, as seen in
Table 3, exhibited an increasing trend as the number of aramid layers increased.
Compared to the 20J/0.10%GNP/HDPE composite, the 10J/10A/0.10%GNP/HDPE
composite registered a significant 147.35% increase in limit velocity. Furthermore, the
hybrid nanocomposite 5J/15A/0.10%GNP/HDPE also stood out as being statistically
comparable to 20A/0.10%GNP/HDPE in terms of ballistic performance.
To comprehend the fracture mechanisms associated with hybrid nanocomposites
when subjected to a 9 mm firearm shooting, the plates are shown before and after the
shootings in Figure 9. In Figure 9(a), the appearance of the plates before the shootings
can be observed. Subsequently, in Figures 9(b) to (d), the plates after the shootings
are displayed. The 10J/10A/0.10%GNP/HDPE hybrid nanocomposite, depicted in
Figure 9(b), exhibited significant delamination at the interface between the aramid
layers and the jute layers. This was a result of multiple shootings, particularly near the
plate edges. However, the analysis of the velocity profile obtained from the equipment
indicates that the initial shooting, occurring at the center, did not result in complete
perforation.

Figure 9. Fracture surfaces of hybrid nanocomposites observed via SEM:


(a) HDPE matrix fracture; (b) HDPE microfibrils; (c) Surface of an aramid layer with detached fibers;
(d) Broken jute fibers with remnants of HDPE.

As mentioned earlier, when analyzing hybrid nanocomposites with over 10 layers


of aramid, complete perforation was not observed, as depicted in Figures 9(c) and (d).
However, clear delamination can be noticed in the 5J/15A/0.10%GNP/HDPE
composite, shown in Figure 9(c). This delamination arises due to the differences in
mechanical properties of the involved fibers. Jute fibers, being more fragile and weaker
compared to aramid fibers, cannot withstand significant deformations. Consequently,
layer separation and jute fiber fracture occur while the aramid fibers remain intact.
Through fracture surface SEM analysis of the fracture surfaces of the hybrid
nanocomposites, illustrated in Figure 10, it was possible to observe the regions
primarily affected by delamination. In Figure 10(a), a fracture mechanism known as
crazing can be identified, characterized by multiple microcracks within the composite
matrix. However, the formation of GNP/HDPE microfibrils contributes to stabilizing the
coalescence of these microcracks, as shown in Figure 10(b) [49,50].

Figure 10. Appearance of hybrid nanocomposites before and after the shootings: (a) Composite plates
from Group 2 before the shootings; (b) HDPE/GNP/10Jute/10Aramid plate after multiple shootings; (c)
HDPE/GNP/5Jute/15Aramid plate after a single shooting; and (d) HDPE/GNP/20Aramid plate after a
single shooting. (Author).

Additionally, in the delamination areas, the ease with which the fabric loses its
structure, resulting in the opening of the weave, can be observed, as depicted in Figure
10(c). This occurrence can be attributed to the lesser adhesion of aramid fibers with
into the polymer matrix compared to jute fibers combined with the GNP/HDPE matrix,
as exemplified in Figure 10(d).
Finally, it is relevant to compare the limit velocity (VLim) values in Table 3 achieved Formatted: Subscript
by the hybrid nanocomposites with the ballistic levels established by the NIJ 0106.01
standard [51], in order to demonstrate the potential of each composite for ballistic
helmet applications. As specified in the mentioned standard, considering an 8 g, 9 mm
caliber projectile, the hybrid nanocomposites with 10 layers reached ballistic level II-A,
while those with 15 and 20 layers achieved level IIIA, even after multiple shootings.
3.7. Cost effectiveness analysis
Considering the provided existing information, a cost and weight analysis per
helmet was conducted, using the PASGT helmet as a reference, with an area of 0.14
m² and an aerial density of 11.2 kg/m², as per Folgar [52]. In this context, Table 54 was
generated, containing the recorded costs and weights of each helmet produced under
the conditions of the hybrid nanocomposites studied in this work. The densities and
prices of each material were obtained from respective commercial companies, such as
Braskem, Sisalsul, Teijin Aramid, and UCSgraphene.
For this purpose, nanocomposite plates with 20 layers of fabric were considered,
with a matrix of 50 vol.% volume of HDPE and 50 vol.% volume of fabric reinforcement.
This approach allowed estimating the weight and cost of each nanocomposite based
on their specific configurations, in accordance with the PASGT model, as described by
Folgar [52].
Table 54. Cost and weight estimates for helmets manufactured using the hybrid nanocomposites
studied in this work.
Cost/unit (USD)
20A/0.10 5J/15A/0.10 10J/10A/0.1 15J/5A/0.10 20J/0.10
Mate Price Weight/
Density %GNP/ %GNP/ 0%GNP/ %GNP/ %GNP/
rial (USD)/kg helmet (kg)
HDPE HDPE HDPE HDPE HDPE
HDP 0.948
2.64 0.66 1.75 1.75 1.75 1.75 1.75
E (g/cm³)
320
Jute 8.70 0.90 0.00 1.95 3.90 5.85 7.80
(g/m²)
Ara 410
60.00 1.15 68.88 51.66 34.44 17.22 0.00
mid (g/m²)
GNP 1.993.30 7.00E-04 1.40 1.40 1.40 1.40 1.40
Total 72.03 56.76 41.49 26.21 10.94
Decrease
- 21.21 42.41 63.61 84.81
in Cost (%)
Weight/
1.81 1.75 1.69 1.62 1.56
helmet (kg)
Decrease in
3.60 7.47 11.64 16.15
weight (%)

Based on the data presented in Table 54, it was observed that the inclusion of 5
layers of jute fabric results in a cost reduction of over 20%. Additionally, the addition of
5 layers of jute fabric leads to a reduction of over 3% in the weight of the
nanocomposites. This outcome underscores the feasibility of using these materials in
ballistic helmet applications, as the hybrid nanocomposite 5J/15A/0.10%GNP/HDPE
maintains the same level of protection as the 20A/0.10%GNP/HDPE nanocomposite.
Consequently, the hybrid nanocomposite 10J/10A/0.10%GNP/HDPE presents a
cost reduction of over 40%. Similarly, it can be observed that the hybrid nanocomposite
10J/10A/0.10%GNP/HDPE demonstrates a weight reduction of over 7% compared to
helmets produced with 20A/0.10%GNP/HDPE. This emphasizes the significant
potential of these hybrid materials, particularly the GNP incorporated nanocomposites,
for applications in ballistic helmets.

4. Conclusions

 Raman spectroscopy detected D and G bands in the GNP/HDPE nanocomposite,


indicating structural changes. The ID/IG ratio suggested minimal structural
defects associated with GNPs. In addition, DSC and XRD results indicated
increased crystallinity with higher GNP content, influencing the crystallization
process.
 DMA revealed enhanced viscoelastic stiffness in GNP-reinforced
nanocomposites, with the highest storage modulus (E’) of 2.53 GPa observed at
0.10 wt,% GNP concentration (2.53 GPa).
 Ballistic results highlight revealed partial perforation on hybrid nanocomposites
with 10 and 15 layers of aramid. It was found a remarkable 526.81% increase in
absorbed energy in 10J/10A/0.10%GNP/HDPE compared to
20J/0.10%GNP/HDPE nanocomposite. There is an Iincreasing trend in limit
velocity as aramid layers increase, with 147.35% increase for
10J/10A/0.10%GNP/HDPE.
 Delamination and fracture in J/A/GNP/HDPE due to weaker jute fibers compared
to aramid fibers. SEM analysis revealed crazing fracture mechanism with
microcracks, as well as GNP/HDPE microfibrils stabilized microcrack
coalescence.
 Using 5 layers of jute fabric results in a cost reduction of over 20% and reduces
the weight of the nanocomposites by over 3%. This cost-effectiveness and weight
reduction make themse promising materials promising for enhancing the
affordability and comfort of ballistic helmets. Additionally, the hybrid
nanocomposite 10J/10A/0.10%GNP/HDPE stands out, with a cost reduction
exceeding 40% and a weight reduction surpassing 7% compared to the
20A/0.10%GNP/HDPE counterparts.

Author Contributions: Conceptualization, data curation, investigations, and


methodology, U.O.C.; Formal analysis, validation, writing—original draft preparation,
writing—review and editing, U.O.C., F.C.G.F., T.G.R.; Prepared testing analyzed data
and wrote the paper, U.O.C. and L.F.C.N.; Performed the tests, E.P.L.Jr; Formal
analysis validation and visualization - F.C.G.F., E.P.L.Jr; T.G.R; Funding acquisition,
writing-review, and editing, S.N.M. All authors have read and agreed to the published
version of the manuscript.
Funding: This research received no external funding
Data Availability Statement: All data underlying the results are available as part of
the article and no additional source data are required.
Acknowledgments: The authors would like to thank the support to this investigation
by the Brazilian agencies: CNPq, CAPES and FAPERJ. As well as the UCSGraphene
for providing the graphite nanoplatelets; the Teijin Aramid for supplying the aramid
fabric; the Macromolecules Institute Professor Eloisa Mano – IMA, for their support in
the thermal analysis; and the Durability and Mechanical Integrity of Structural Materials
Group (DIMME), Rey Juan Carlos University, School of Experimental Sciences and
Technology – RJCU for their support in the dynamic mechanical tests, and the Brazilian
Army Evaluation Center (CAEX) for their support in the ballistic tests.
Conflicts of Interest: The authors declare no conflict of interest
References

1. Seif A, Fathy A, Megahed AA. Effect of drilling process parameters on bearing strength of glass
fiber/aluminum mesh reinforced epoxy composites. Scientific Reports, 2023. 13(1), 12143.
2. Najjar IMR, Sadoun AM, Abd Elaziz M, Abdallah AW, Fathy A, Elsheikh AH. Predicting kerf quality
characteristics in laser cutting of basalt fibers reinforced polymer composites using neural network
and chimp optimization. Alex. Eng. J., 2022. 61(12), 11005-11018.
3. Shehata F, Fathy A, Megahed M, Morsy D. Fabrication and characterization of nano-filled polymer
composites. EIJEST, 2019. (28), 33-38.
4. Shehata F, Abdelhameed M, Fathy A, Elmahdy M. Preparation and characteristics of Cu-Al 2 O
3 nanocomposite. J. Met., 2011. 1(02), 25.
1.5. Jagadeesh P, Puttegowda M, Boonyasopon P, Rangappa SM, Khan A, Siengchin S. Recent
developments and challenges in natural fiber composites: A review. Polym. Compos., 2022.
43(5), 2545-2561.
2.6. Melaibari A, Wagih A, Basha M, Lubineau G, Al-Athel K, Eltaher MA. Sandwich composite
laminate with intraply hybrid woven CFRP/dyneema core for enhanced impact damage resistance
and tolerance. J. Mater. Res. Technol., 2022. 21, 1784-1797.
3.7. Yuan Z, Ma W, Xu W, Sun Y, Gu B, Chen X. A numerical study on stress wave propagation in
quasi-isotropic stacks of Dyneema® compliant composite laminates. Compos. Struct., 2023. 312,
116869.
4.8. Haque BZ, Gillespie Jr JW. Depth of penetration of Dyneema® HB26 hard ballistic laminates. J.
Thermoplast. Compos. Mater., 2023. 36(4), 1361-1381.
5.9. Chinnapandi M, Katiyar A, Nandi T, Velmurugan R. High-Velocity Impact Studies on Dyneema
Fabric with and without STF-Experimental and Theoretical Studies. In Composite Materials for
Extreme Loading: Proceedings of the Indo-Korean workshop on Multi Functional Materials for
Extreme Loading 2021, Springer Singapore. 2022. (pp. 269-291)
6.10. Weerasinghe D, Bambach MR, Mohotti D, Wang H, Hazell PJ. High-velocity projectile impact
response of rubber-coated aramid Twaron fabrics. Int. J. Mech. Sci., 2022. 229, 107515.
7.11. Xu YJ, Zhang H, Huang GY. Ballistic performance of B4C/STF/Twaron composite fabric.
Compos. Struct., 2022. 279, 114754.
8.12. Zhang B, Jia L, Tian M, Ning N, Zhang L, Wang W. Surface and interface modification of aramid
fiber and its reinforcement for polymer composites: A review. Eur. Polym. J., 2021. 147, 110352.
9.13. Nasser J, Lin J, Steinke K, Sodano HA. Enhanced interfacial strength of aramid fiber reinforced
composites through adsorbed aramid nanofiber coatings. Compos. Sci. Technol., 2019. 174, 125-
133.
10.14. Gonzalez-Chi PI, Rodríguez-Uicab O, Martin-Barrera C, Uribe-Calderon J, Canché-Escamilla G,
Yazdani-Pedram M, Avilés F. Influence of aramid fiber treatment and carbon nanotubes on the
interfacial strength of polypropylene hierarchical composites. Compos. B Eng., 2017. 122, 16-22.
11.15. Nasser J, Lin J, Sodano H. High strength fiber reinforced composites with surface fibrilized aramid
fibers. J. Appl. Phys., 2018. 124(4).
12.16. Xu T, Qi Z, Yin Q, Jiao Y, An L, Tan Y. Effects of Air Plasma Modification on Aramid Fiber Surface
and Its Composite Interface and Mechanical Properties. Polym., 2022. 14(22), 4892.
13.17. Meliande NM, Silveira PHPMD, Monteiro SN, Nascimento LFC. Tensile Properties of Curaua–
Aramid Hybrid Laminated Composites for Ballistic Helmet. Polym., 2022. 14(13), 2588.
14.18. Meliande NM, Oliveira MS, Silveira PHPMD, Dias RR, Marçal RLSB, Monteiro SN, Nascimento
LFC. Curaua–Aramid Hybrid Laminated Composites for Impact Applications: Flexural, Charpy
Impact and Elastic Properties. Polym., 2022. 14(18), 3749.
15.19. Costa UO, Garcia Filho FDC, Río TGD, Rodrigues JGP, Simonassi NT, Monteiro SN, Nascimento
LFC. Mechanical Properties Optimization of Hybrid Aramid and Jute Fabrics-Reinforced
Graphene Nanoplatelets in Functionalized HDPE Matrix Nanocomposites. Polym., 2023. 15(11),
2460.
16.20. Tomasi TB, Vargas N, Rodrigues DR, Miranda PI, Rossa BLV, Lavoratti A, Zattera AJ. Influence
of the addition of graphene nanoplatelets on the ballistic properties of HDPE/aramid multi-laminar
composites. Polym.-Plast. Technol. Mater., 2022. 61(4), 363-373.
17.21. Salman SD. The influence of kenaf contents and stacking sequence on drop-weight impact
properties of hybrid laminated composites reinforced polyvinyl butyral composites. J. Ind.
Text., 2022. 51(5_suppl), 8645S-8667S.
18.22. Sarker F, Karim N, Afroj S, Koncherry V, Novoselov KS, Potluri P. High-performance graphene-
based natural fiber composites. ACS Appl. Mater. Interfaces., 2018. 10(40), 34502-34512.
19.23. Sarker F, Potluri P, Afroj S, Koncherry V, Novoselov KS, Karim N. Ultrahigh performance of
nanoengineered graphene-based natural jute fiber composites. ACS Appl. Mater.
Interfaces., 2019. 11(23), 21166-21176.
20.24. Karim N, Sarker F, Afroj S, Zhang M, Potluri P, Novoselov KS. Sustainable and Multifunctional
Composites of Graphene‐Based Natural Jute Fibers. Adv. Sustain. Syst., 2021. 5(3), 2000228.
21.25. Braskem. High-density polyethylene technical data sheet. Available online:
https://www.braskem.com.br/eu-826 rope/product-search?p=228, (accessed on 29 April 2023).
22.26. Escocio VA, Visconte LL, Cavalcante ADP, Furtado AMS, Pacheco EB. Study of mechanical and
morphological properties of bio-based polyethylene (HDPE) and sponge-gourds (Luffa-cylindrica)
agroresidue composites. AIP Conf. Proc. (Vol. 1664, No. 1). AIP Publishing. 2015.
23.27. Walsh SM, Scott BR, Jones TL, Cho K, Wolbert J. A materials approach in the development of
multi-threat warfighter head protection. US Army Res. Lab., Aberdeen Proving Ground, EUA:
Harford County, MD, USA. 2008.
24.28. Albetran HM. Structural characterization of graphite nanoplatelets synthesized from graphite
flakes. Preprints. 2020.
25.29. Seehra MS, Geddam UK, Schwegler-Berry D, Stefaniak AB. Detection and quantification of 2H
and 3R phases in commercial graphene-based materials. Carb., 2015. 95, 818-823.
26.30. Segal LGJMA, Creely JJ, Martin Jr AE, Conrad CM. An empirical method for estimating the
degree of crystallinity of native cellulose using the X-ray diffractometer. Text. Res. J., 1959.
29(10), 786-794.
27.31. Evgin T, Turgut A, Hamaoui G, Spitalsky Z, Horny N, Micusik M, Omastova M. Size effects of
graphene nanoplatelets on the properties of high-density polyethylene nanocomposites:
morphological, thermal, electrical, and mechanical characterization. Beilstein J. Nanotechnol.,
2020. 11(1), 167-179.
28.32. McNally T, Pötschke P, Halley P, Murphy M, Martin D, Bell SE, Quinn JP. Polyethylene
multiwalled carbon nanotube composites. Polym., 2005. 46(19), 8222-8232.
29.33. American Society for Testing and Materials. ASTM D7028 - 08: Standard test method for glass
transition temperature (dma tg) of polymer matrix composites by dynamic mechanical analysis
(dma). West Conshohocken, PA, 2015. 13 p.
30.34. Kaftelen‐Odabaşı H, Odabaşı A, Özdemir M, Baydoğan M. A study on graphene reinforced
carbon fiber epoxy composites: Investigation of electrical, flexural, and dynamic mechanical
properties. Polym. Compos., 2023. 44(1), 121-135.
31.35. Xiang D, Wang L, Tang Y, Zhao C, Harkin‐Jones E, Li Y. Effect of phase transitions on the
electrical properties of polymer/carbon nanotube and polymer/graphene nanoplatelet composites
with different conductive network structures. Polym. Int., 2018. 67(2), 227-235.
32.36. Fischer J, Wallner GM, Pieber A. Spectroscopical investigation of ski base materials. Macromol.
Symp., Weinheim: WILEY‐VCH Verlag. 2008. 265(1), pp. 28-36.
33.37. Kelkar VP, Rolsky CB, Pant A, Green MD, Tongay S, Halden RU. Chemical and physical changes
of microplastics during sterilization by chlorination. Water Res., 2019. 163, 114871.
34.38. Nabiyev AA, Islamov AK, Maharramov AM, Nuriyev MA, Ismayilova RS, Doroshkevic AS, Kuklin
AI. Structural Studies of dielectric HDPE+ ZrO2 polymer nanocomposites: filler concentration
dependences. J. Phys.: Conference Series. IOP Publishing. 2018. 994(1), p. 012011).
35.39. Silva DJD, Wiebeck H. Predicting LDPE/HDPE blend composition by CARS-PLS regression and
confocal Raman spectroscopy. Polímeros, 2019. 29.
36.40. Moriche R, Prolongo SG, Sánchez M, Jiménez-Suárez A, Sayagués MJ, Ureña A. Morphological
changes on graphene nanoplatelets induced during dispersion into an epoxy resin by different
methods. Compos. B. Eng. 2015. 72, 199-205.
37.41. Salleh FM, Hassan A, Yahya R, Azzahari AD. Effects of extrusion temperature on the rheological,
dynamic mechanical and tensile properties of kenaf fiber/HDPE composites. Compos. B. Eng.,
2014. 58, 259-266.
42. Ahmadian H, Zhou T, Abd Elaziz M, Al-Betar MA, Sadoun AM, Najjar IMR, Yu Q. Predicting
crystallite size of Mg-Ti-SiC nanocomposites using an adaptive neuro-fuzzy inference system
model modified by termite life cycle optimizer. Alex. Eng. J., 2023. 84, 285-300.
43. Fouly A, Kuppusamy S, Kulandaivel A, Abdullaev S, Fathy A, Hassan A, Mohanavel V. Design
and high efficient construction of bilayer NiCoO2/Poly (1-NA-co-oT) nanocomposite absorber for
X-band stealth applications. Vacuum, 2023. 112792.
44. Megahed M, Fathy A, Morsy D, Shehata F. Mechanical performance of glass/epoxy composites
enhanced by micro-and nanosized aluminum particles. J. Ind. Text., 2021. 51(1), 68-92.
45. Fathy A, Shaker A, Hamid MA, Megahed AA. The effects of nano-silica/nano-alumina on fatigue
behavior of glass fiber-reinforced epoxy composites. J. Compos. Mat., 2017. 51(12), 1667-1679.
46. Sadoun AM, Abd El-Wadoud F, Fathy A, Kabeel AM, Megahed AA. Effect of through-the-
thickness position of aluminum wire mesh on the mechanical properties of GFRP/Al hybrid
composites. Journal of Materials Research and Technology, 2021. 15, 500-510.
38.47. Mohanty S, Verma SK, Nayak SK. Dynamic mechanical and thermal properties of MAPE treated
jute/HDPE composites. Compos. Sci. Technol., 2006. 66(3-4), 538-547.
39.48. Saba N, Jawaid M, Alothman OY, Paridah MT. A review on dynamic mechanical properties of
natural fibre reinforced polymer composites. Constr. Build. Mater., 2016. 106, 149-159.
40.49. Garcia Filho FC, Luz FS, Oliveira MS, Pereira AC, Costa UO, Monteiro SN. Thermal behavior of
graphene oxide-coated piassava fiber and their epoxy composites. J. Mater. Res.
Technol., 2020. 9(3), 5343-5351.
41.50. Tarani E, Chrysafi I, Kállay-Menyhárd A, Pavlidou E, Kehagias T, Bikiaris DN, Chrissafis K.
Influence of graphene platelet aspect ratio on the mechanical properties of HDPE
nanocomposites: Microscopic observation and micromechanical modeling. Polym., 2020. 12(8),
1719.
42.51. United State Department of Justice. NIJ (National Institute of Justice) Standard for Ballistic
Helmets. 1981.
43.52. Folgar F. Thermoplastic matrix combat helmet with carbon-epoxy skin for ballistic performance.
In Advanced Fibrous Composite Materials for Ballistic Protection. Woodhead Publ., 2016. pp.
437-456.
Declaration of Competing Interest

Declaration of interests

☒The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:
Figure 1 Click here to access/download;Figure;Imagem1.png
Figure 2 Click here to access/download;Figure;Graph3.png
Figure 3a Click here to access/download;Figure;Imagem3.jpg
Figure 3b Click here to access/download;Figure;Imagem3b.jpg
Figure 4a Click here to access/download;Figure;Imagem4a.jpg
Figure 4b Click here to access/download;Figure;Imagem4b.jpg
Figure 4c Click here to access/download;Figure;Imagem4c.jpg
Figure 4d Click here to access/download;Figure;Imagem4d.jpg
Figure 5a Click here to access/download;Figure;Imagem5a.png
Figure 5b Click here to access/download;Figure;Imagem5b.png
Figure 5c Click here to access/download;Figure;Imagem5c.png
Figure 6a Click here to access/download;Figure;Imagem6a.png
Figure 6b Click here to access/download;Figure;Imagem6b.png
Figure 7a Click here to access/download;Figure;Imagem7a.png
Figure 7b Click here to access/download;Figure;Imagem7b.png
Figure 7c Click here to access/download;Figure;Imagem7c.png
Figure 8a Click here to access/download;Figure;Imagem8a.png
Figure 8b Click here to access/download;Figure;Imagem8b.png
Figure 8c Click here to access/download;Figure;Imagem8c.png
Figure 9a Click here to access/download;Figure;Figure 9a.jpeg
Figure 9c Click here to access/download;Figure;Figure 9c.jpeg
Figure 9d Click here to access/download;Figure;Figure 9d.jpeg
Figure 9b Click here to access/download;Figure;Figure 9b.jpeg
Figure 10a Click here to access/download;Figure;Figure 10a.tif
Figure 10b Click here to access/download;Figure;Figure 10b.tif
Figure 10c Click here to access/download;Figure;Figure 10c.tif
Figure 10d Click here to access/download;Figure;Figure 10d.tif
Table 5 Click here to access/download;Table;Table 5.docx

Table 5. Cost and weight estimates for helmets manufactured using the hybrid nanocomposites
studied in this work.
Cost/unit (USD)
Price 20A/0.10 5J/15A/0.1 10J/10A/0. 15J/5A/0.1 20J/0.10
Mat Densit Weight/
(USD)/k %GNP/ 0%GNP/ 10%GNP/ 0%GNP/ %GNP/
erial y helmet (kg)
g HDPE HDPE HDPE HDPE HDPE
HDP 0.948
2.64 0.66 1.75 1.75 1.75 1.75 1.75
E (g/cm³)
320
Jute 8.70 0.90 0.00 1.95 3.90 5.85 7.80
(g/m²)
Ara 410
60.00 1.15 68.88 51.66 34.44 17.22 0.00
mid (g/m²)
GNP 1.993.30 7.00E-04 1.40 1.40 1.40 1.40 1.40
Total 72.03 56.76 41.49 26.21 10.94
Decrease
- 21.21 42.41 63.61 84.81
in Cost (%)
Weight/
1.81 1.75 1.69 1.62 1.56
helmet (kg)
Decrease in
3.60 7.47 11.64 16.15
weight (%)
Table 2 Click here to access/download;Table;Table 2.docx

Table 2. Summary of the pic’s parameters of the nanocomposites


Pic
Intensity Crystallinity
Material 2Theta FWHM
(counts) (%)
(°)
30.69 54678.5 0.27
GNP 51.96 1960.6 0.46 100.00
64.18 2334.8 0.49
25.20 78598.4 0.30
28.10 18590.6 0.38
HDPE 57.45
35.21 888.5 0.37
42.61 1554.6 0.44
25.30 84634.9 0.33
28.10 13365.3 0.40
31.12 4424.0 0.21
0.10%GNP/HDPE 58.60
35.32 695.1 0.33
42.61 1624.2 0.47
51.81 341.9 0.48
25.30 72471.8 0.36
28.10 10193.0 0.43
31.12 24231.9 0.22
0.25%GNP/HDPE 60.17
35.32 516.0 0.42
42.61 1255.8 0.48
51.81 330.0 0.51
25.30 90894.1 0.28
28.10 12733.2 0.36
31.12 43299.3 0.16
0.50%GNP/HDPE 35.32 735.1 0.33 65.43
42.61 1725.8 0.43
51.81 469.7 0.43
64.18 1305.5 0.30
Table 3 Click here to access/download;Table;Table 3.docx

Table 2. Summary of properties obtained from the residual velocity test for all the composites
studied in this work.
Vimp VRes Eab VLim
Material
(m/s) (m/s) (J) (m/s)
20J/HDPE 423.40±6.10 402.17±7.56 69.50±5.53 130.79±5.54
20J/0.10%GNP/HDPE 421.61±2.87 401.26±3.28 66.98±3.39 129.37±3.25
20J/0.25%GNP/HDPE 419.86±2.00 401.41±2.44 62.94±7.44 124.78±6.38
20J/0.50%GNP/HDPE 421.47±2.80 402.91±3.14 64.66±2.92 127.12±2.88
15J/5A/0.10%GNP/HDPE 421.87±5.74 377.66±6.13 141.36±9.81 187.90±6.50
10J/10A/0.10%GNP/HDPE(*) 422.59±2.20 262.94±74.93 419.84±146.45 320.13±55.60
5J/15A/0.10%GNP/HDPE(*) 418.54±5.25 158.01±42.24 659.41±55.83 405.72±17.54
20A/0.10%GNP/HDPE(**) 422.48±2.16 0 713.96±7.29 422.48±2.16
(*) Partial perforation occurred, (**) No perforation occurred
Table 4 Click here to access/download;Table;Table 4.docx

Table 4. Tukey Test Results for Absorbed Energy Values of Hybrid Nanocomposites.
20J/
15J/5A/0.10%G 10J/10A/0.10%G 5J/15A/0.10%G 20A/0.10%GN
0.10%GNP
NP/HDPE NP/HDPE NP/HDPE P/HDPE
/HDPE
20J/
0.10%GNP/HDP 0.00 74.39 352.86 592.43 646.98
E
15J/5A/0.10%G
74.39 0.00 278.47 518.05 572.60
NP/HDPE
10J/10A/0.10%G
352.86 278.47 0.00 239.57 294.12
NP/HDPE
5J/15A/0.10%G
592.43 518.05 239.57 0.00 54.55
NP/HDPE
20A/0.10%GNP/
646.98 572.60 294.12 54.55 0.00
HDPE
Table 1 Click here to access/download;Table;Table 1.docx

Table 1. Hybrid nanocomposites configurations and associated nomenclatures


Group Composites Nomenclature
50 vol% Jute + HDPE 20J/HDPE
50 vol% Jute + 0.10%p GNP + HDPE 20J/0.10%GNP/HDPE
1
50 vol% Jute + 0.25%p GNP + HDPE 20J/0.25%GNP/HDPE
50 vol% Jute + 0.50%p GNP + HDPE 20J/0.50%GNP/HDPE

50 vol% Jute + 0.X%p GNP + HDPE 20J/0.X%GNP/HDPE


50 vol% Aramid + 0.X%p GNP + HDPE 20A/0.X%GNP/HDPE
2 37.5 vol% Jute + 12.5 vol% Aramid + 0.X%p GNP + HDPE 15J/5A/0.X%GNP/HDPE
25 vol% Jute + 25 vol% Aramid + 0.X%p GNP + HDPE 10J/10A/0.X%GNP/HDPE
12.5 vol% Jute + 37.5 vol% Aramid + 0.X%p GNP + HDPE 5J/15A/0.X%GNP/HDPE

You might also like