JMRTEC-1218; No. of Pages 12
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
Available online at www.sciencedirect.com
www.jmrt.com.br
Original Article
Effect of graphene oxide coating on the ballistic performance of aramid fabric Anderson Oliveira da Silva ∗ , Ricardo Pondé Weber, Sérgio Neves Monteiro, Andreza Menezes Lima, Géssica Seara Faria, Wesley Oliveira da Silva, Suzane de Sant’ Ana Oliveira, Karollyne Gomes de Castro Monsores, Wagner Anacleto Pinheiro Military Institute of Engineering — IME, Department of Materials Science, Prac¸a General Tibúrcio 80, URCA, 22290-270 Rio de Janeiro, RJ, Brazil
a r t i c l e
i n f o
a b s t r a c t
Article history:
Twaron® brand aromatic polyamide, commonly known as aramid fabric, was reinforced
Received 24 October 2019
with a solution of graphene oxide (GO) followed by a heat treatment at 120 ◦ C. The objective
Accepted 19 December 2019
was to improve the ballistic properties of the material by increasing the friction coefficient of
Available online xxx
the fibers in the fabric. The GO produced by the modified Hummers method was character-
Keywords:
microscopy (SEM) and atomic force microscopy. Both the as-received fabric and fabric coated
Graphene oxide
with one and two filtrations of GO were characterized by the above-mentioned technique
ized by Fourier transform infrared and Raman spectroscopy as well as by scanning electron
Aramid
in addition to optical microscopy (OM). All samples were subjected to ballistic tests with 9
Nanocomposite
mm ammunition. The energy absorbed by the materials was calculated from the values of
Ballistic test
impact velocities and residual obtained by Doppler radar. The type of failure was characterized by means of SEM and OM visual analysis. The results show a better resistive force of the two filtrations GO coated fabric, up to 50%, when compared to the plain as-received fabric. © 2019 Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
1.
Introduction
The gradual increase of armed conflicts between police and criminals in large urban centers is today associated with escalating changes in the characteristics of used armaments. In a scenario where criminals have access to increasingly powerful armaments, it becomes necessary to obtain ever more effec-
∗
tive protection armor, as they become a key element in the personal and vehicular security of civil and military police [1]. Among the materials used in the manufacture of ballistic protection equipment, one can highlight those associated with fibers, both natural and synthetic. Natural fibers are a recent class investigated for possible substitution or hybridization with synthetic fibers in multilayered armor [2–7]. It is worth mentioning that, within this class, nanocellulose is a promising reinforcement for application requiring specials properties [8–14]. However, synthetic fiber-based armors are still the most preferred worldwide in ballistic vests for personal protection. This is particularly the case of aro-
Corresponding author. E-mail:
[email protected] (A.O. Silva). https://doi.org/10.1016/j.jmrt.2019.12.058 2238-7854/© 2019 Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/ licenses/by-nc-nd/4.0/). Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
2
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
Fig. 1 – Schematic deformation and cone formation in the back face of a ballistically tested textile material [7].
matic polyamide, also known as aramid, a synthetic fiber, which, has high strength and specific toughness, associated with low density and good mobility [15]. These specific fiber properties, added to the fabric architecture (weft and warp), contribute to the absorption of kinetic energy transferred from the projectile to the target in the form of elastic and plastic deformations [16–19]. The mechanisms of damage during the ballistic event coexist with these characteristics, contributing to the energy dissipation of the projectile in the material. Mechanism of cone formation on the back face of the target, failure by tensile of the primary fibers, secondary fibers deformation, cutting and friction of the fibers with the projectile, delamination, and matrix breaking, when applicable, are most relevant [20–23]. In general, when a projectile hits a textile (fabric) target, it stretches the fibers toward the back face of the compound, propagating transverse and longitudinal/radial waves that promote a conical deformation in the material. The fibers directly impacted by the projectile (primary fibers) are responsible for the penetration resistance, being subjected to higher tensions until reaching the failure by tensile rupture [20,23,24]. The other fibers, not directly affected by the projectile (secondary fibers), suffer a conical deformation Fig. 1, contributing to the dissipation of energy in the material. Since the impact effects are localized, the area of the secondary fibers is restricted to a radius rc equal to the distance traveled by the deformation waves generated from the impact [20]. During projectile penetration, the fabric intersections can promote lateral displacement of the fibers without breaking, but reducing the potential for energy dissipation and generating a perforation smaller than the original diameter of
the projectile [25,26,18]. The friction interactions at the intersections of the fibers under ballistic impact are improved when different contour conditions (for example, the addition of lubricants) are applied under a smooth surface fabric [25,27–30]. Oliveira [31] reported that aramid plates, whose fibers were treated with shear thickening fluid (STF) based on colloidal nano-silica, impacted by 9 mm ammunition, showed an improvement of about 25% in the absorbed energy, demonstrating an efficient interaction of a non-Newtonian fluid with aramid fibers in ballistic protection. With the emergence of graphene as a high strength and rigid material [32–34], researchers have been working to show that combining this material with aramid can significantly improve the mechanical and ballistic performance of the fiber. Recent results revealed that the application of graphene nanoplates on the surface of these fabrics improves tensile strength, Young’s modulus and tenacity of the nanocomposites [35,36]. Other studies [37,38] suggest that the oxide graphene (GO), consisting of regions interspersed with sp2-hybridized carbons containing carboxyl groups and sp3-hybridized carbons with hydroxyl and epoxy functional groups in their basal planes, improves the interaction of polymeric fiber surfaces. This improvement is attributed to the better mechanical fit at the nano-charge-matrix interface and its two-dimensional geometry. It was suggested that GO may be responsible for the higher stiffness obtained in the nanocomposite, resulting from the hexagonal and 2D arrangement of the atoms of covalently bound carbon in a honeycomb structure [20,33,38]. In this view, the present work evaluates the ballistic response of a did phase material composed of aramid fabric coated with GO, before and after deposition, through the energy dissipated by the impact of the 9 mm projectile and against the target as well as by the type of failure suffered by the material.
2.
Materials and methods
2.1.
Synthesis of graphene oxide
The graphene oxide (GO) was fabricated from the graphite flakes, number 332461, supplied by Sigma-Aldrich, using the modified Hummers method [39–41]. In short, from flakes of graphite the following steps were performed: (i) intercalation, by the addition of sulfuric acid and sodium nitrate; (ii) oxidation, by the addition of potassium permanganate, with stirring for 7 days; (iii) exfoliation, by means of additions of dilute solutions of sulfuric acid and hydrogen peroxide, follows by washing and centrifugation cycles [42,43]. Fig. 2 illustrates these steps performed for the production of graphene oxide. Graphene oxide flakes suspended in the aqueous medium were produced at a concentration of 2 mg/ml.
2.2. Preparation of aramid fabrics coated with graphene oxide flake Aramid fabric circular plies having a diameter of 15 cm were cut from a Twaron® textile, with T750 weave and 460 g/m2 grammage, supplied by Teijin Aramid, with a linear density of
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
3
Fig. 2 – Steps performed for the production of graphene oxide.
3360 Dtex - 2000 filaments. These aramid plies were heated to 100 ◦ C to remove adsorbed moisture on the fibers. The filtration in this method was applied either once or twice, characterizing two different GO-coated aramid fabric pieces, with first and second filtration, respectively. The GO flakes suspended in the aqueous medium with a concentration of 2 mg/ml, were deposited on both surfaces of the aramid fabric pieces by the vacuum filtration method. The GO-modified pieces were then vacuum dried in a desiccator. After drying, the pieces were again heated to a temperature of 120 ◦ C for 30 min in an oven. This treatment promotes an increase in the adhesion of the GO to the fiber, related to the formation of the functional groups (−COOH) on the surface of the fiber and the greater interaction of these with the oxygenated groups in the GO [42]. Each ballistic sample was composed of 5 piece layers as further described. Three test groups were prepared for the study: Received Aramid (uncoated and dried at 100 ◦ C), Aramid + GO (1) (samples coated with GO on a first filtration, thermally treated at 120 ◦ C) and Aramid + GO (2) (GO-coated samples in a second filtration, thermally treated at 120 ◦ C). Because it was an experimental study, a small amount of GO (1 l) was used to coat the aramid fabric. Thus, the vacuum filtration methodology was performed either one or two times using the same chemical solution. The morphology of the samples was determined by atomic force microscopy (AFM), scanning electron microscopy (SEM) and optical microscopy (OM). AFM was performed on a Park System XE7 equipment in intermittent mode with a silicon tip, operating under the resonance frequency of 13 kHz and force of 0.18 N/m. SEM observation was conducted on a Quanta FEG 250 FEI microscope with secondary electrons operating under 10 and 30 kV at low vacuum. As the samples were not coated with a polymer resin to prevent removal of the GO impregnated into the aramid fabric, a sample of the GO-coated fabric was removed for manual manipulation for about 1 min. This was intended to check
the possible loss of tissue due to surface coating. SEM micrographs, Fig. 3, show a partial break of the GO sheets on the fiber surface. Since a significant amount of GO remained in the fiber, it was assumed that the adhesion of the GO into the fiber was efficient and that the manipulation of the samples during the ballistic test would not significantly affect the results of this work.
2.3.
Preparation of ballistic samples
For the ballistic test, Fig. 4 shows both aramid fabric samples (Twaron® ) without GO coating and samples for each condition of GO-coated aramid fabric with one or two filtrations. These samples were prepared with the aid of an industrial sewing machine, in which for each ballistic sample, five pieces layers were sewn from each evaluation group. For fixing the samples to the test apparatus, a system displaying an aperture in the center with a diameter of 5 cm, was prepared with two aluminum plates. The samples were fixed between these plates with the id of screws, as illustrated in Fig. 5. The ballistic tests were carried out according to the NIJ Standard 0106.01 [44] level II, using 9 mm ammunition fired from a weapon (gun barrel) on targets 5 m away, at an incident angle of 0◦ . Fig. 6 illustrates schematically the ballistic test setup located at the Brazilian Army shooting range facility (CAEx), Rio de Janeiro, Brazil. The ballistic performance of the nanocomposite was evaluated considering the energy absorbed, calculated by Eqs. (1) and (2), using the residual (Vr ) and impact (Vi ) velocity obtained by Doppler radar model Weibel SL-520P, after complete perforation. Where EL is the absorbed energy, Ei the impact of energy and ER the residual energy. E = ½.m.(V2 )
(1)
EL = Ei −ER = ½.m.(Vi 2 −Vr 2 )
(2)
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
4
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
Fig. 3 – SEM of the manually manipulated sample (magnification of 1000 and 3000×).
Fig. 4 – Ballistic samples made from aramid fabrics reinforced with graphene oxide. (a) Aramid, (b) aramid + GO (1), (c) aramid + GO (2).
Fig. 5 – The system produced for fixing ballistic samples in the test apparatus.
Fig. 6 – (a) Ballistic test scheme [32,33] and (b) photograph the apparatus used by CAEx.
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
ARTICLE IN PRESS 5
j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
Analysis of variance ANOVA was performed to statistically validate the absorbed energy values, Eq. (2), by the different groups of samples: Received Aramid; Aramid + GO (1) and Aramid + GO (2) investigated. The damage produced by the projectile impact, in terms of the occurrence of stretching or cutting of the threads, was analyzed by visual inspection. The topography of the fracture surfaces of GO-coated and as-received aramid fabric was analyzed by SEM.
2.4.
Physico-chemical characterizations
Physical-chemical analyses were performed to evaluate the impregnation of GO onto the surface of aramid fabric and compared to untreated fabric and as-produced GO. Raman spectroscopy was performed in a NT-MDT NTEGRA spectrometer, with laser wavelength () 473 nm and radiation time of 100 s [43]. Fourier Infrared spectroscopy (FTIR) analysis was performed using a Frontier FT-IR/FIR spectrometer, in the region between 4000 cm−1 and 650 cm−1 , applying the attenuated total reflectance (ATR) technique. This analysis was used to verify the existence of changes in the structure of the samples, serving as a tool to identify possible variations in the wavelength and characteristic absorptions of the samples. The spectra were normalized based on the band at 821 cm−1 , attributed to stretching vibrations of the C H bonds of the aromatic ring present in the backbone of the aramid fiber. This band does not present significant variations of intensity when the fiber is subjected to thermochemical treatment [45].
3.
Results and discussion
3.1.
FTIR and Raman spectroscopy
Fig. 7 shows the spectra obtained by FTIR and Raman, characterizing the graphene oxide (GO) produced by the modified Hummers method similar to what was reported in other works [46–50]. The FTIR spectrum of GO in Fig. 7(a) shows the main characteristic absorption bands of the material. The large band in the spectral region of 3000–3700 cm−1 is attributed to stretching and vibrations of hydroxyl groups (O H) of water molecules adsorbed onto GO. This indicates that the sample produced has a high hydrophilic character [46,48]. The characteristic bands observed in the 2110 and 1633 cm−1 regions can be attributed, respectively, to epoxidized groups in the basal planes and to carbonyl groups (C O) and carboxylic groups ( COOH) [49]. The Raman spectrum in Fig. 7(b) shows a large and intense D-band at 1356 cm−1 resulting from sp3 carbon stretching and a visible G-band at 1590 cm−1 due to the stretching of sp2 carbons [51]. The acute aspect of the D band, in this case, implies the occurrence of sp3-type structural disturbance generated from the modification of the original carbon atoms of the sp2type graphene [46,48,49]. In the spectral region of 2924 cm−1 , there is another band characteristic of defects in the material, called 2D, which can be associated with the number of layers. When the 2D band exhibits an enlargement as seen in Fig. 7(b), it is an indicative that possibly graphene has four
Table 1 – Absorbed energy by the distinct aramid fabric samples. Evaluation group
A (Kg/m2 )
Energies absorbed (J)
Aramid Received Aramid + GO (1) Aramid + GO (2)
2,34 2,27 2,22
106,4 116,6 159,6
or more layers [52,53]. Thus, the sample reveals all the main characteristics of the GO.
3.2.
AFM, SEM and OM
SEM and AFM micrographs show that the GO synthesized by the Hummers method has a two-dimensional sheet-like structure [41]. The thickness profile extracted by AFM, Fig. 8(a), shows that the graphene flakes are approximately 3.05 nm. According to Lima et al. [43] and Chen et al. [53], a monolayer of GO has a thickness of approximately 0.8 nm. This indicates that the material produced has sheets of ∼4 layers. The application of these GO nanoplates onto the surface of aramid fabrics may show a significant improvement in the fracture toughness of the material, attributed to the best mechanical fit between the nano load-matrix interface and its two-dimensional geometry [38]. Fig. 9 shows images by OM and SEM of the three distinct samples. Both OM, Fig. 9(a), and SEM, Fig. 9(b), micrographics of Aramid Received sample display smooth surfaces suggesting, respectively, well-cleaned aramid fabric and related fibers. The OM Fig. 9(c), and SEM, Fig. 9(d), micrographics of Aramid + GO (1) show surfaces partially covered with relatively large GO flakes from a first filtration. Surfaces covered with GO from a second filtration are shown by OM, Fig. 9(e), and SEM, Fig. 9(f), micrographics for the Aramid + GO (2) sample. In particular, the aramid fiber shown by SEM in Fig. 9(f) reveals a surface almost completely covered with small GO flakes forming a veil-flake film, first mentioned by Lee et al. [34]. As further discussed, this more uniform GO coating might be responsible for the superior ballistic performance of the Aramid + GO (2) sample. Fig. 10 shows by AFM nanometric details on the surface of the three distinct samples. Aramid Received, Fig. 10(a), display a comparatively more smoother surface with a maximum roughness around 2 nm. The GO flakes covering the surface of the Aramid + GO (1) sample, Fig. 10(b), contribute to a sensibly rougher surface with thickness oscillation above 4 nm. By contrast, the more uniform graphene oxide veil-coated Aramid + GO (2) sample, Fig. 10(c), has a thickness that reaches about 7 nm.
3.3.
Ballistic absorbed energy
Table 1 shows the calculated values form Eq. (2) of the absorbed energy (EL ) by the studied materials in ballistic tests. Since the Doppler radar did not capture the impact velocity (Vi ) on the samples, it was assumed that the projectile velocity at the exit of the gun barrel to be (Vs ) [6,54,55]. The results showed that the ballistic resistance of the aramid fabric increases with the deposition of GO by values up to 50% higher than those for the as-received fabric. It is noted in Table 1 that the resistance varied significantly as a
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
6
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
Fig. 7 – Infrared spectroscopy (a) and Raman (b) of the graphene oxide obtained.
Fig. 8 – (a) Scanning electron microscopy (2500×) and atomic force — AFM (b) of synthesized graphene oxide flakes.
function of the areal density of the samples. In the case of the Aramid + GO (2) sample, the second filtration may have contributed to the GO flakes entering more into the fabrics, increasing the friction between them and the projectile. Consequently, the ballistic energy absorption of the material is significantly increased. These results, associated with those found by Gago et al. [56], Wetzel et al. [57] and Bizao et al. [32] support the feasibility of the development of new nanocomposites for ballistic applications. Fig. 11 shows the variation of the ballistic absorbed energy with the sample areal density, from the date in Table 1. In this figure is noteworthy the marked increase in absorbed energy for the second filtration Aramid + GO (2) sample with relatively smaller areal density. It is then suggested that the GO coating process might interfere with the space between yarns of the aramid fabric, reducing slightly its areal density. At the same time, a better GO coating significantly increases the absorbed energy owing to further discussed fracture mechanisms. The results show in Table 1 and Fig. 11 were validated by ANOVA. Table 2 presents the values of the ANOVA parameters F, F critical (Fc) and P. In this table, considering that 5% (0.05) is
Table 2 – Analysis of variance (ANOVA) parameters for ballistic absorbed energy of aramid fabric, as received and GO-coated. ANOVA parameters
F
Fc
P-value
Group of aramid fabric samples
115,06
3,182
0,0015
the level of significance, then those with F > Fc are associated with 95% of confidence for their difference. Moreover, if P < 0.05, a significant distinction might exist between values. As presented in Table 2, based on the ANOVA, the surface deposition of GO onto aramid fabric causes a statistically significant increase in the ballistic absorbed energy. In other words, the hypothesis that the values are the same can be rejected with 95% of confidence.
3.4.
Damage in fracture analysis
Figs. 12 and 13 show the macroscopic damage imposed to the ballistically tested samples. It can be seen in Fig. 12 that all tested samples were completely perforated, indicating that the initial energy of the projectile was greater than the kinetic
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
7
Fig. 9 – Microscopic images of the light field type with the magnification of 200 m and SEM (16,000×), of the Aramid Received (a) and (b), coated with of GO: Aramid + GO (1) (c) and (d), Aramid + GO (2) (e) and (f).
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
8
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
Fig. 10 – Microscopy atomic force (AFM) the Aramid Received (a), coated with of GO: Aramid + GO (1) (b) and Aramid + GO (2) (c).
energy that the target could absorb. The cone formed in the back face of the aramid fabric during the ballistic event is attributed to the propagation of transverse waves in the fibers [58], pulling all the yarns towards the point of impact [20]. In this case, the resistive force of the aramid fabric during the ballistic impact was attributed to the yarns and their fibers (primary fibers) directly affected by the projectile and respon-
sible for the greater absorption of kinetic energy during the event. The deformation of the other fabric yarns within the conic region (secondary fibers) added to this energy absorption. Fig. 13 shows SEM micrographs of the fracture surface of the distinct samples after the ballistic test. Fig. 13(a) reveals that there was a displacement towards the side of the fibers in the
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
9
of microfibrillation in the fracture of the fibers during the ballistic impact, contributing to the greater absorption of energy, shown in Fig. 13(d). Another point worth discussing is the nanometric scale of the GO, especially in the second filtration Aramid + GO (2). As indicated by Zeng et al. [38], the GO-coated aramid fibers might be considered a nanocomposite in which the nano-sized flakes of GO form a surface coating that reinforces the fibers. Indeed, the results in Table 1 and Fig. 11 revealed a ballistic reinforcement of GO-coated in comparison with non-coated as-received aramid fabric. Although the reinforcement effect of both graphene and GO on aramid fabric/KevlarTM has already been known [33,36,38], the present work is for the first time presenting a ballistic reinforcement caused by nanometric GO coating on aramid fabric. Fig. 11 – Ballistic absorbed energy of the distinct aramid fabric samples as a function of the fabric areal density.
4.
Conclusions
The results of the ballistic test on graphene oxide (GO) coating aramid fabric demonstrated that this GO-coated nanocomposite enables the development of novel materials for personal and vehicular ballistic protection. Aramid Received sample. This is probably due to the lower friction between the projectile and the target [26]. As a result, the portion of fibers struck by the projectile was small and, consequently, there was a lower energy absorption by the material. In Fig. 13(b) it is noted that there was fibrillation along the diameter of the aramid fibers due to the existence of a gradient of properties inherent to the manufacturing process. This mechanism promoted by the stretching of fibers is known as “peeling” [59,60]. On the other hand, Fig.13(c) shows that the aramid fabric coated with GO presented a greater amount of fibers ruptured after the ballistic test. Microscopically, there was a marked pull-out effect of the fibers, accompanied by fractures in a typical tensile morphology [60], Fig. 13(d), contributing significantly to the greater energy absorption of the material when compared to the as-received aramid fabric. It is proposed that GO coated on the surface of the aramid fabric promotes a higher interfacial resistance in the fibers, which causes a greater resistance to friction between projectile and target. In this case, the improvement of this resistive force associated with the GO coating friction promotes an amount
• The GO sheets, with up to 2 nm in thickness, produced by the modified Hummers method was found to have besides their main characteristics, a high hydrophilic character revealed by FTIR and Raman spectroscopy. • The thermal treatment at 120 ◦ C improved the adhesion of the GO sheets to the aramid fibers, disclosed by AFM and SEM, which could be attributed to the formation of functional groups on the surface of the fiber with the oxygenated groups in the graphene oxide. • The ballistic test on GO-coated aramid fabric showed a significant improvement in absorbed energy of up to 50% when compared to the uncoated aramid fabric, possibly due to higher friction between the fibers. ANOVA statistically validated these results. • The main mechanisms of damage observed were microfibrillation failure and primary fiber tensile rupture as well as secondary fiber deformation, friction between fiber and projectile due to cone formation in the back face of the nanocomposite.
Fig. 12 – Photographs of the samples ballistically tested with 9 mm ammunition. (a) Aramid Received, (b) Aramid + GO (1) and (c) Aramid + GO (2). Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
10
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
Fig. 13 – SEM fractures of tested ballistically samples with 9 mm ammunition. (a) The cone fracture surface of the Aramid Received sample, (b) “peeling” of the aramid fibers coated (4000×), (c) cone fracture surface in the Aramid + GO (2) and (d) microfibrillation of aramid fibers with GO after the ballistic test (8000×).
Conflicts of interest [2]
The authors declare no conflicts of interest.
Acknowledgments
[3]
The authors of this paper thank the Brazilian support agency CAPES for funding; to the Protective and Nanostructured Materials Coating Laboratory (PUC-RJ) for the contribution to the analyses, and to the Brazilian Army Evaluation Center (CAEx) for conducting ballistic tests.
[4]
[5]
references
[1] Nascimento LFC, Louro LHL, Monteiro SN, Gomes AV, Júnior É, Marc¸al RLSB. Ballistic performance in multilayer armor with epoxy composite reinforced with Malva fibers. Miner
[6]
Met Mater 2017:331–8, http://dx.doi.org/10.1007/978-3-319-52132-9 33. Benzait A, Trabzon L. A review of recent research on materials used in polymer-matrix composites for body armor application. J Compos Mater 2018;52(23):3241–63, http://dx.doi.org/10.1177/0021998318764002. Naveen J, Jawaid M, Zainudim ES, Sultan MT, Yahaya R. Evaluation of ballistic performance of hybrid Kevlar/cocos nucifera sheath reinforced epoxy composites. J Text Inst 2018;110:1179–89, http://dx.doi.org/10.1080/00405000.2018.1548801. Monteiro SN, Pereira C, Pereira Junior E, Weber RP, Assis F. Performance of plain-woven jute fabric-reinforced polyester matrix composite in multilayered ballistic system. Polymers 2018;10(3):230, http://dx.doi.org/10.3390/polym10030230. Assis FS, Pereira AC, da Costa Garcia Filho F, Lima EP Jr, Monterio SN, Weber RP. Performance of jute plain-woven mat reinforced polyester matrix composite in multilayered armor. J Mater Res Technol 2018;7(4):535–40, http://dx.doi.org/10.10110/j.jmrt.2018.05.026. Silva AO, de Castro Monsores KG, Oliveira SSA, Weber RP, Monteiro SN. Ballistic behavior of a hybrid composite
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21] [22]
reinforced with curaua and aramid fabric subjected to ultraviolet radiation. J Mater Res Technol 2018;7(4):584–91, http://dx.doi.org/10.1016/j.jmrt.2018.09.004. Costa UO, Nascimento LFC, Garcia JM, Monteiro SN, Luz FS, Pinheiro WA, et al. Effect of graphene oxide coating on natural fiber composite for multilayered ballistic armor. Polymers 2019;11:1356, http://dx.doi.org/10.3390/polym11081356. Abral H, Ariksa J, Mahardika M, Handayani D, Aminah I, Sandrawati N, et al. Transparent and antimicrobial cellulose film from ginger nanofiber. Food Hydrocolloids 2020;98:105266, http://dx.doi.org/10.1016/j.foodhyd.2019.105266. Ilyas RA, Sapuan SM, Ibrahim R, Atikah MSN, Atiqah A, Ansari MNM, et al. Production, processes and modification of nanocrystalline cellulose from agro-waste: a review. Intech Open 2019:3–32, http://dx.doi.org/10.5772/intechopen.87001. Halimatul MJ, Sapuan SM, Jawaid M, Ishak MR, Ilyas RA. Effect of sago starch and plasticizer content on the properties of thermoplastic films: mechanical testing and cyclic soaking-drying. Polimery 2019;64(6):422–31. Ilyas RA, Sapuan SM, Ibrahim R, Abral H, Ishak MR, Zainudin ES, et al. Effect of sugar palm nanofibrillated cellulose concentrations on morphological, mechanical and physical properties of biodegradable films based on agro-waste sugar palm (Arenga pinnata (Wurmb.) Merr) starch. J Mater Res Technol 2019;8(5):4819–30. Atiqah A, Jawaid M, Sapuan SM, Ishak MR, Ansari MNM, Ilyas RA. Physical and thermal properties of treated sugar palm/glass fibre reinforced thermoplastic polyurethane hybrid composites. J Mater Res Technol 2019;8(5): 3726–32. Ilyas RA, Sapuan SM, Atiqah A, Ibrahim R, Abral H, Ishak MR, et al. Sugar palm (Arenga pinnata [Wurmb.] Merr) starch films containing sugar palm nanofibrillated cellulose as reinforcement: water barrier properties. Polym Compos 2019:1–9. Ilyas RA, Sapuan SM, Ibrahim R, Abral H, Ishak MR, Zainudin ES, et al. Sugar palm (Arenga pinnata (Wurmb.) Merr) cellulosic fibre hierarchy: a comprehensive approach from macro to nano scale. J Mater Res Technol 2019;8(3):2753–66. Hussain S, Yorucu C, Ahmed I, Hussain R, Chen B, Khan MB, et al. Surface modification of aramid fibres by graphene oxide nano-sheets for multiscale polymer composites. Surf Coat Technol 2014;258:458–66, http://dx.doi.org/10.1016/j.surfcoat.2014.08.054. Crouch IG. The science of armour materials. Woodhead Publishing; 2016. Copyright©, Elsevier Ltd. All rights reserved. 2017, 754 p. Morye SS, Hine PJ, Duckett RA, Carr DJ, Ward IM. Modelling of the energy absorption by polymer composites upon ballistic impact. Compos Sci Technol 2000;60:2631–42. Abrate S, Ahmadi M. Dynamic deformation, damage and fracture in composite materials and structures. Woodhead Publishing; 2016. Copyright©, Elsevier Ltd. All rights reserved. Shi Y, Pinna C, Soutis C. Low-velocity impact of composite laminates: damage evolution. In: Abrate S, Ahmadi M, editors. Dynamic deformation, damage and fracture in composite materials and structures. Woodhead Publishing; 2016. Copyright©, Elsevier Ltd. All rights reserved. Naik NK, Shrirao P, Reddy BCK. Ballistic impact behavior of woven fabric composites: formulation. Int J Impact Eng 2006;32:1521–52. Cunniff PM. An analysis of the system effects in woven fabrics under ballistic impact. Text Res J 1992;62(9):495–509. Bhatnagar A. In: Bhatnagar A, editor. Lightweight ballistic composites military and law-enforcement applications.
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35] [36]
[37]
[38]
[39]
11
Abington Hall, Abington, Cambridge CB1 6AH, England: Woodhead Publishing Limited; 2006. p. 364–97. Pasquali M, Terra C, Gaudenzi P. Analytical modelling of high-velocity impacts on thin woven fabric composite targets. Compos Struct 2015;131:951–65, http://dx.doi.org/10.1016/j.compstruct.2015.06.078. Nilakantan G, Wetzel ED, Bogetti TA, Gillespie JW. Finite element analysis of projectile size and shape effects on the probabilistic penetration response of high strength fabrics. Compos Struct 2012;94(5):1846–54, http://dx.doi.org/10.1016/j.compstruct.2011.12.028. Duan Y, Keefe M, Bogetti TA, Cheeseman BA. Modeling friction effects on the ballistic impact behavior of a single-ply high-strength fabric. Int J Impact Eng 2005;31:996e1012. Briscoe BJ, Motamedi F. The ballistic impact characteristics of aramid fabrics: the influence of interface friction. Wear 1992;158:229–47. Duan Y, Keefe M, Bogetti TA, Cheeseman BA, Powers B. A numerical investigation of the influence of friction on the energy absorption by a high strength fabric subjected to ballistic impact. Int J Impact Eng 2006;32:1299–312. Das S, Jagan S, Shaw A, Pal A. Determination of inter-yarn friction and its effect on ballistic response of para-aramid woven fabric under low velocity impact. Compos Struct 2015;120:129–40, http://dx.doi.org/10.1016/j.compstruct.2014.09.063. Chu Y, Chen X. Finite element modeling effects of inter-yarn friction on the single-layer high-performance fabrics subject to ballistic impact. Mech Mater 2018;126:99–110, http://dx.doi.org/10.1016/j.mechmat.2018.08.003. Wang Y, Chen X, Young R, Kinloch I. Finite element analysis of effect of inter-yarn friction on ballistic impact response of woven fabrics. Compos Struct 2016;135:8–16. Oliveira AM. Development of nanomodified compounds using aged aramid fibers: an alternative to ballistic protections (in Portuguese). Dissertation (146 f.). Master’s thesis (Mechanical Engineering). Federal University of Minas Gerais - MG; 2016. Available in: http://hdl.handle.net/1843/BUBD-AXVL9P. [Access to: 6 March 2019]. Bizao RA, Machado LD, de Sousa JM, Pugno NM, Galvao DS. Scale effects on the ballistic penetration of graphene sheets. Sci Rep 2018;8(1), http://dx.doi.org/10.1038/s41598-018-25050-2. Martirosyan KS, Zyskin M. Modeling and simulation of the elastic properties of kevlar reinforced by graphene. In: Aliofkhazraei M, Ali N, Milne WI, Ozkan CS, Mitura S, Gervasoni JL, editors. Graphene science handbook: siza-dependet properties. CRC Press; 2016, 505 p. Lee J-H, Loya PE, Lou J, Thomas EL. Dynamic mechanical behavior of multilayer graphene via supersonic projectile penetration. Science 2014;346(6213):1092–6. Mukhopadhyay P, Gupta RK. Graphite, graphene, and their polymers nanocomposites. CRC Press; 2013, 589 p. Martirosyan KS, Zyskin M. Simulation of the elastic properties of reinforced kevlar-graphene composites. Int J Nanosci 2012;11(3):1–11. Rodríguez-Uicab O, Avilés F, Gonzalez-Chi PI, Canché-Escamilla G, Duarte-Aranda S, Yazdani-Pedram M, et al. Deposition of carbon nanotubes onto aramid fibers using as-reicevedand chemically modified fibers. Appl Surf Sci 2016;385:379–90. Zeng L, Liu X, Chen X, Soutis C. Surface modification of aramid fibres with graphene oxide for interface improvement in composites. Appl Compos Mater 2018;25:843, http://dx.doi.org/10.1007/s10443-018-9718-9. Rourke JP, Pandey PA, Moore JJ, Bates M, Kinloch IA, Young RJ, et al. The real graphene oxide revealed — stripping the
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058
JMRTEC-1218; No. of Pages 12
12
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
ARTICLE IN PRESS j m a t e r r e s t e c h n o l . 2 0 1 9;x x x(x x):xxx–xxx
oxidative debris from the graphene-like sheets. Angew Chem Int Ed 2011;50:3173–7. Alam SN, Sharma N, Kumar L. Synthesis of graphene oxide (GO) by moddified hummers method and its thermal reduction to obtain reduce graphene oxide (rGO). Graphene 2017;6:1–18, http://dx.doi.org/10.4236/graphene.2017.61001. Zaaba NI, Foo KL, Hashim U, Tan SJ, Liu W-W, Voon CH. Synthesis of graphene oxide using modified hummers method: solvent influence. Proc Eng 2017;184:469–77. Silva AO, Menezes AL, Farias GS, Monsores KGC, Mark MFV, Pine WA, et al. Study of graphene oxide adhesion in an aromatic polyamide fabric by heat treatment. 72nd ABM Annual Congress, Sao Paulo 2017:1024–33, http://dx.doi.org/10.5151/1516-392X-30388. ISSN: 2594-5327. Lima AM, Faria GS, Flowers JOH, Cross LRO, Pine WA. Evaluation of graphene oxide crystallite size by different techniques. 72nd ABM Annual Congress, Sao Paulo 2017:553–8, http://dx.doi.org/10.5151/1516-392X-30266. ISSN: 2594-5327. NIJ Standards-0101.06. Ballistic resistance of body armor. U.S. Department of Justice/Office of Justice Programs — National Institute of Justice; 2008. Jia C, Chen P, Li N, Wang Q, Lu C, Yu Q. Effects of Twaron fiber surface treatment by air dielectric barrier discharge plasma on the interfacial adhesion in fiber reinforced composites. Surf Coat Technol 2010;204:3668–75. Marcano DC, Kosynkin DV, Berlin JM, Sinitskii A, Sun ZZ, Slesarev A, et al. Improved synthesis of graphene oxide. ACS Nano 2010;4(8):4806–14. ˇ y´ M, Berger J, Redondo J, Telychko M, Procházka P, Konecn ˇ Vondrácek M, et al. Simple device for the growth of monocrystalline graphene of micrometer size in SiC (0001). J Vac Sci Technol A 2018;36(3):031401, http://dx.doi.org/10.1116/1.5008977. Alam SN, Sharma N, Kumar L. Synthesis of graphene oxide (GO) by modified hummers method and its thermal reduction to obtain reduced graphene oxide (rGO). Graphene 2017;6:1–18, http://dx.doi.org/10.4236/graphene.2017.61001. Hussain S, Yorucu C, Ahmed I, Hussain R, Chen B, Khan MB, et al. Surface modification of aramid fibres by graphene oxide nano-sheets for multiscale polymer composites. Surf
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58] [59]
[60]
Coat Technol 2014;258:458–66, http://dx.doi.org/10.1016/j.surfcoat.2014.08.054. Noh YJ, Joh HI, Yu J, Hwang SH, Lee S, Lee CH, et al. Ultra-high dispersion of graphene in polymer composite via solvent free fabrication and functionalization. Sci Rep 2015;5:9141, http://dx.doi.org/10.1038/srep09141. Ferrari AC, Meyer JC, Scardaci V, Casiraghi C, Lazzeri M, Mauri F, et al. Raman spectrum of graphene and graphene layers. Phys Rev Lett 2006;97(18), http://dx.doi.org/10.1103/physrevlett.97.187401. Wang Y ying, Ni Z hua, Yu T, Shen ZX, Wang H min, Wu Y hong, et al. Raman studies of monolayer graphene: the substrate effect. J Phys Chem C 2008;112(29):10637–40, http://dx.doi.org/10.1021/jp8008404. Chen J, Yao B, Li C, Shi G. An improved Hummers method for eco-friendly synthesis of graphene oxide. Carbon 2013;64:225–9, http://dx.doi.org/10.1016/j.carbon.2013.07.055. Braga FO, Bolzan LT, Luz FS, Lopes PHLM, Lima EP Jr, Monteiro SN. High energy ballistic and fracture comparison between multilayered armor systems using non-woven curaua fabric composites and aramid laminates. Mater Res Technol 2017;6(4):417–22, http://dx.doi.org/10.1016/j.jmrt.2017.08.001. Yahaya R, Sapuan SM, Jawaid M, Leman Z, Zainudin ES. Mechanical performance of woven Kenaf-Kevlar hybrid composites. J Reinf Plast Compos 2014;33(24):2242–54. Gago I, del Río M, Carretero J, Ibarra I, León G, Miguel B. Graphene-based nanocomposites for bulletproof vests: ballistic limits and properties in tension and impact. V National Congress of R&D in Defense and Security 2017 (In Portuguese). Wetzel ED, Balu R, Beaudet TD. A theoretical consideration of the ballistic response of continuous membranes. J Mater Phys Solids 2015;82:23–31, http://dx.doi.org/10.1016/j.jmps.2015.05.008. Meyers MA. Dynamic behavior of materials. New York, USA: John Wiley & Sons; 1994. Naglis MMM, D’Almeida JARM. Evaluation of failure modes under the impact of fiber-reinforced polymer matrix composites. Polym Sci Technol 1998:54–60 (In Portuguese). Sawyer LC, Grubb DT. Polymer microscopy. 2ed Cambridge: Chapman & Hall.; 1994, 303 p.
Please cite this article in press as: Silva AO, et al. Effect of graphene oxide coating on the ballistic performance of aramid fabric. J Mater Res Technol. 2019. https://doi.org/10.1016/j.jmrt.2019.12.058