Sei sulla pagina 1di 10

DOI 10.

1007/s11029-018-9724-x
Mechanics of Composite Materials, Vol. 54, No. 1, March, 2018 (Russian Original Vol. 54, No. 1, January-February, 2018)

TOUGHENING EFFECT OF MICROSCALE PARTICLES

ON THE TENSILE AND VIBRATION PROPERTIES OF S-GLASS-

FIBER-REINFORCED EPOXY COMPOSITES

A. Erkliğ,1 M. Bulut,2* and B. Fayzulla3

Keywords: particulate-filled composites, natural frequency, damping

The effect of borax, sewage sludge ash, silicon carbide, and perlite microparticles on the tensile, damping, and
vibration characteristics of S-glass/epoxy composite laminates was examined Their damping and vibration
properties were evaluated experimentally by using the dynamic modal analysis, identifying the response of the
fundamental natural frequency to the type and weight content of the particulates. The results obtained showed
that the introduction of specific amounts of such particulates into the matrix of S-glass/epoxy composite
noticeably improved its mechanical properties.

1. Introduction

Composites are material made of different constituent materials, and they possess various remarkable prop-
erties [1-3]. Particulate composites can contain nano- or microparticles of different types and shapes, which can be
oriented randomly or directionally [4,5].
In general, vibrations are classified into two categories — free and forced ones [6, 7]. The dynamic behavior of
a vibrating structural element can be characterized by three parameters, namely, the natural frequency, mode shape, and
damping factor. Damping is the inherent property of a material system to resist vibrations [8]. Bridges, aircraft wings,
wind turbines, and other structures may be exposed to external loadings that can lead to their resonance and catastrophic
failure when their vibration amplitude suddenly increases. Therefore, it is very important to determine the natural fre-

1
Gaziantep University, Faculty of Engineering, Mechanical Engineering Department, Gaziantep 27310, Turkey
2
Hakkari University, Faculty of Engineering, Mechanical Engineering Department, Hakkari 30000, Turkey
3
Erbil Polytechnic University, Erbil Technology Institute, Mechanical Department, Iraq
*
Corresponding author; e-mail: mehmetbulut@hakkari.edu.tr

Russian translation published in Mekhanika Kompozitnykh Materialov, Vol. 54, No. 1, pp. 171-184, January-
February, 2017. Original article submitted April 7, 2017; revision submitted July 5, 2017.

0191-5665/18/5401-0119 © 2018 Springer Science+Business Media, LLC 119


TABLE 1. Chemical Compositions of Fillers

Filler Chemical formula/Composition wt.%


Borax 99% sodium tetraborate decahydrate, Na2B4O7/(B2O3), (Na2O)
Perlite SiO2 (73), Al2O3 (15), Na2O3 (3.5), K2O (3), Fe2O3 (0.3), MgO (0.15), TiO2 (0.09), SO3 (0.1)
SiO2 (16.45), Al2O3 (5.93), Fe2O3 (10.71 ), CaO (2.24 ), MgO (2.49 ), SO3 (17.07 ), TiO2 (1.82),
Sewage sludge ash
P2O5 (6.33 ), ZnO (3.53), K2O (1.24), Na2O (1.20)
Silicon carbide SiC (100)

quencies and eliminate them from designs [9]. For this purpose, many researchers studied the effect of various fillers on
composite properties. Spiliotis et al. [10] used pet coke (PC) and sewage sludge (SSA) as admixtures to clay materials for
producing ceramics. Kutuk and Oguz [11] experimentally investigated the effect of sewage sludge ash on the mechanical
properties (tensile, flexural, and impact ones) of particle-filled composite materials. An unsaturated polyester resin was
used for producing composites. Patnaik et al [12] investigated the effects fly ash, alumina, and silicon carbide as filler
materials on the mechanical properties of glass/polyester composites. The study showed that the industrial filler fly ash
led to better their mechanical properties than alumina and silicon carbide. Asi [13] investigated the mechanical properties
of glass-fiber-reinforced epoxy composite filled with different proportions of Al2O3 particles. The results obtained showed
that the ultimate tensile strength and shear strength of the composites decreased with increasing content of Al2O3. Ozsoy
et al.  [14] studied the influence of micro- (Al2O3, TiO2, and fly ash) and nano- (Al2O3, TiO2, and clay) fillers on the me-
chanical properties of epoxy-based composites. They found that the tensile strength, flexural strength, and elongation at
break of the composites decreased, but the tensile modulus and flexural modulus increased with content of the micro- and
nanofillers. Pichi Reddy et al. [15] studied the effect of fly ash on the tensile strength of glass fiber/epoxy composites.
Agarwal et al. [16] investigated the effect of SiC content on thermomechanical properties (tensile, flexural, and impact
strengths) of chopped-glass-fiber-reinforced epoxy composites. Results showed that their mechanical properties increased
with content of SiC filler up to 10 wt.%, but then they decreased. Sankar et al. [17-18] revealed that the inclusion of rub-
ber particles in glass fabric/epoxy composites increased the loss factors in bending and shear, but decreased the storage
modulus. Datta and Włoch [19] found that the addition of 2 wt.% ZnO or TiO2 to an unsaturated polyester resin enhanced
the flexural strength and flexural moduli of composites with a glass fabric. A dynamic mechanical analysis showed a
growth in the storage and loss moduli. Pol et al. [20] revealed that an increasing volume fraction of nanoclay from 1 to
5% increased the natural frequency of beams made of an E-glass/epoxy hybrid composite from 10.88 to 11.66 Hz and its
damping coefficient from 0.041 to 0.055. However, when the volume fraction of nanoclay exceeded 7%, both the natural
frequency and damping ratio decreased. Kumar et al. [21] studied the mechanical and damping properties of a glass-epoxy
composite with carbon filler and found that the damping characteristics improved with growing weight ratio of the filler
up to 5 wt.%.
From the literature survey, it follows that the effect of various amounts of micro- and nanoscale particle fillers on
the mechanical properties of composites have been investigated extensively. However, no studies on the tensile, damping,
and vibration properties of S-glass/epoxy composite laminates with borax, sewage sludge ash, silicon carbide or perlite
particles can be found in the open literature. The present study aims to show variations in the tensile, damping, and vibra-
tion properties of S-glass/epoxy (GE) composite laminates with introduction of borax (Bx), silicon carbide (SiC), sewage
sludge ash (SSA), and perlite (Pr) particulates into them.

2. Materials and Methods

In this study, woven plain S-glass fiber plies with an areal density of 200 g/m2 were used as the reinforcement in
laminas. An epoxy resin (MOMENTIVE-MGS L285) with a hardener (MOMENTIVE-MGS H285) at a stoichiometric
ratio of 100:40 was used as the matrix. All materials were acquired from the Dost Kimya company, Istanbul. The chemical

120
compositions of the fillers are given in Table 1. The anhydrous borax (Etibor-68) was supplied by Etibor A.Ş., Bandırma,
Turkey, perlite — by Inper Perlit, Gaziantep, Turkey, SSA — by Gaziantep Büyükşehir Belediyesi, Turkey, and SiC — by
Esan Company, Turkey.
The particulates and epoxy resin were mixed in a bowl at constant speed of 800 rpm, and a hardener was
added in the ratio 0.285 of the mass of epoxy. The epoxy-particulate mixture was put on woven glass fibers with a
paint brush using a vacuum-supported production unit. Altogerher, 16 plies of impregnated fabrics with dimensions
of 220 × 320 mm were prepared and laid one on another layer by layer. The packs were first placed on flat molds,
subjected to a 0.3-MPa pressure for 1 h at 80°C for curing, and then cooled down to room temperature under the
pressure. The contents filler in the composite plates were chosen to be 0 (plain), 5, 10, 15, and 20 wt.% for SSA, Bx,
and SiC 1, 3, 5, and 10 wt.% for Pr. Tensile test specimens, according to the ASTM D638–10 [22] standard, were cut
from the plates. The vibration test specimens were 185 mm long, 12.7 mm wide, and 3.5 ± 0.3 mm thick. The free
beam length was kept at 138 mm.
The specimens were tested using a 300-kN Shimadzu AG-X series tensile testing machine. To determine the tensile
properties, five different specimens were tested at a crosshead motion speed of 2 mm/min up to failure.

2.1. Vibration tests

The dynamic characteristics of the composite laminates were measured using an experimental set-up. In the ex-
periments, a general-purpose PCB 352C03 ceramic shear ICP® accelerometer, a general-purpose PCB 086C03 modal impact
hammer, and a NI 9234 National Instrument product with a LABVIEW software were used for acquisition of the output
signal, the stimulus force signal, and data. The accelerometer was placed near the fixed edge. The frequency response func-
tion (FRF) and time signal graph were plotted on a computer by using the LABVIEW software, but the natural frequencies
were plotted on the screen from FRF curves by using the modal impact hammer. The natural frequency ωn , damping ratio
ξ , and the storage E ′ and loss E ′′ moduli were determined. The frequency responses were registered within the range
from 0 to 500 Hz. In the experiments, five specimens were excited in three different locations.
The half-power bandwidth method was used for determining the damping ratio. The maximum amplitude of the
first vibration mode was determined from the frequency response curve, and the damping ratio ξ was calculated from the
formula [23]
ω − ω1
ξ= 2 ,
2ωn

where ωn is the natural frequency of the first vibration mode and ω2 − ω1 is the bandwidth. The storage modulus E ′ was
obtained from Euler–Bernoulli beam theory [23], according to which

1.8752 E ′I
ωn = ,
2π L2 ρA
where L is the free length of the beam, r is the density of beam material, I is the second cross-sectional area moment of the
beam, and A is its cross section. The loss modulus E ′′ was found from the storage modulus using the relationship

E ′′(ω ) = 2 E ′(ω )ξ (ω ).

121
TABLE 2. Densities of Particle-Filled Fiber-Reinforced Composites

Composites wc ,% ρ, kg/m3
GE 0 1721 ± 12
GE-Bx 5 1664 ± 18
10 1667 ± 16
15 1617 ± 09
20 1628 ± 13
GE-SiC 5 1663 ± 07
10 1645 ± 15
15 1546 ± 12
20 1691 ± 10
GE-SSA 5 1592 ± 11
10 1632 ± 12
15 1632 ± 15
20 1661 ± 13
GE-Pr 1 1632 ± 12
3 1616 ± 16
5 1619 ± 18
10 1616 ± 15

3. Results and Discussion

3.1. Density variations

Table 2 shows the densities of composites with different types and content of particulates. As is seen, addition of the
fillers decreases the density, which may also decrease the weight of structures — the crucial property of structural materials.
Patnaik [12] reported that the density decreased linearly with growing content of filler not for all types of composites. In
addition, the density of a composite also depends on the relative proportion of matrix and reinforcing materials, and this is
one of the most important factors determining the properties of composites.

3.2. Results of tensile tests

The experimental tensile strength, modulus, and maximum strain of the composites considered are presented in
Table 3, and the stress–strain curves at different particulate ratios are illustrated in Fig. 1. These results show that the
tensile strength greatly depend on the type and content of filler. As seen in Fig. 1a, with Bx content growing to 10 wt.%,
the tensile strength increased, but a further increase in its content reduced the strength. This can be explained by the fact
that, at low concentrations, the particles in the matrix acted as barriers of stress transfer from one point to another, but
at higher ones, exceeding 10 wt.% they promoted this transfer. In addition, as the fiber/filler content increases, the bond-
ing area of surface also increases, resulting in a low bonding strength between a particulate and fiber. Because of the
insufficient amount of bonds between the three different constituents, the loads are not effectively transferred from one
point another, and the tensile strength of composite decreases. This phenomen is also reported in [15, 16]. In our case,
the maximum tensile strength, 430.07 MPa, was reached at 10 wt.% of Bx particulates, which by 10.5% exceeded that
of the unfilled composite.

122
TABLE 3. Tensile Properties of GE-Bx, GE-SiC, GE-SSA, and GE-Pr Composites

Composites wc, % E, GPa stu, МPa etu


GE 0 20 ± 2 389 ± 16 0.0229 ± 0.0015
GE-Bx 5 23 ± 3 418 ± 18 0.0264 ± 0.0011
10 21 ± 1 430 ± 17 0.0272 ± 0.0018
15 19 ± 2 401 ± 11 0.0280 ± 0.0013
20 19 ± 2 394 ± 15 0.0263 ± 0.0015
GE-SiC 5 19 ± 3 373 ± 16 0.0211 ± 0.0010
10 21 ± 1 394 ± 9 0.0218 ± 0.0015
15 20 ± 2 354 ± 12 0.0239 ± 0.0018
20 20 ± 2 343 ± 12 0.0223 ± 0.0019
GE-SA 5 18 ± 1 373 ± 10 0.0211 ± 0.0012
10 19 ± 1 422 ± 16 0.0271 ± 0.0017
15 19 ± 2 386 ± 15 0.0259 ± 0.0015
20 19 ± 2 370 ± 19 0.0218 ± 0.0016
GE-Pr 1 21 ± 3 443 ± 18 0.0261 ± 0.0014
3 20 ± 2 432 ± 12 0.0240 ± 0.0012
5 20 ± 1 413 ± 11 0.0231 ± 0.0013
10 20 ± 2 379 ± 17 0.0213 ± 0.0016

, МPa , МPa
500
500

400 Bx SiC
400

300 300 5
3 4
200 200
5 3
4 2
100 1 100 1
2
 
0 0.01 0.02 0.03 0 0.01 0.02 0.03

, МPa , МPa
500
500 Pr
SSA
400
400
300 3
3 300
4 2
200 5
2 200 5
1 4
100 100 1
 
0 0.01 0.02 0.03 0 0.01 0.02 0.03

Fig. 1. Tensile stress–strain curves s– e of the composites investigated.

123
a b

20 мкм 20 мкм

c d

20 мкм 20 мкм

Fig. 2. SEM photos of the fracture surfaces of GE-Bx (a), GE-SiC (b), GE-SSA (c), and GE-Pr (d)
tensile specimens with the maximum tensile strength.

The inclusion of Pr led to a maximum tensile strength of 443 MPa at 1 wt.%, showing a 13.8% improvement
(Fig. 1d). There can be two reasons for the increased strength of these hybrid composites compared with that of the unfilled
GE. Fiest, the chemical reaction at the interface between filler particles and the matrix might be too strong to transfer the
tensile stress. Second, the corner points of the irregularly shaped particulates, due to their small size (< 35 µm), caused
lower stress concentrations in the polymeric matrix [24]. However, the tensile strength decreased with growing filler content.
The addition of SSA and SiC particulates, excluding the 10 wt.% one, affected the tensile strength negatively
(Fig. 1b, c) — it decreased by 1 and 8.4% from GE to GE-SiC and GE-SSA. The mechanical properties of the filler-rein-
forced composites were affected mainly by the interfacial adhesion between the fillers and matrix polymer. The decrease
in the tensile strength of the SiC- and SSA-filled composites may be due to the poor bonding, resulting in a low interface
adhesion between the filler and resin. The edge points of irregular shapes of SiC may produce stress concentrations in
the composites [25]. A reduction in the tensile strengths was observed at all ratios for GE-SSA, except for the GE-SSA
composites at 10 wt.%.
The lowest tensile strength was obtained at 20 wt.% of each of the filler considered, whose values were 394, 370,
and 343 MPa for Bx, SSA, and SiC, respectively. The decreasing of tensile properties at high percentages of the fillers
may be due to the low bonding strength between them and the epoxy matrix [13].
The strain at failure, in general, increased with all fillers, excluding SiC. The minimum failure strains were re-
corded at 20 wt.% of Bx or SSA, 10 wt.% of Pr, and 5 wt.% of SiC.
The scanning electron microscopy (SEM) images of the fracture surfaces of tensile specimens are shown in Fig. 2.
The surfaces were flat, without any necking, which indicated that specimens had failed in a brittle manner. The SEM
image of the unfilled composite showed that its damage mechanism was pull-out fibers from the matrix and their break.
The fracture surfaces of the composite specimens were examined at their maximum strength, i.e., GE-10 wt.% Bx, GE-10

124
TABLE 4. Natural Frequencies and Damping Ratios of GE-Bx, GE-SiC, GE-SSA, and GE-Pr Composites

Composites wc, % ω1, Hz Damping ratio E′, GPa E′′, GPa


GE 0 87 ± 1.4 0.276 ± 0.012 19 ± 0.5 10 ± 0.22
GE-Bx 5 102 ± 2.5 0.355 ± 0.011 22 ± 0.9 15 ± 0.29
10 102 ± 2.8 0.299 ± 0.016 21 ± 0.6 12 ± 0.25
15 104 ± 2.2 0.216 ± 0.015 18 ± 0.5 8 ± 0.150
20 107 ± 1.9 0.271 ± 0.012 18 ± 0.7 10 ± 0.12
GE-SiC 5 92 ± 2.4 0.300 ± 0.013 17 ± 0.9 10 ± 0.14
10 104 ± 1.7 0.103 ± 0.014 20 ± 0.9 4 ± 0.050
15 108 ± 1.5 0.245 ± 0.016 16 ± 1.0 7 ± 0.020
20 107 ± 2.3 0.290 ± 0.011 19 ± 0.8 11 ± 0.06
GE-SSA 5 106 ± 3.3 0.202 ± 0.015 17 ± 0.7 7 ± 0.09
10 102 ± 2.8 0.405 ± 0.015 18 ± 0.5 14 ± 0.50
15 107 ± 1.9 0.262 ± 0.017 19 ± 0.8 10 ± 0.90
20 110 ± 1.5 0.167 ± 0.012 18 ± 0.4 6 ± 0.080
GE-Pr 1 103 ± 2.4 0.183 ± 0.014 20 ± 0.8 7 ± 0.050
3 104 ± 2.1 0.309 ± 0.013 19 ± 0.2 11 ± 0.09
5 105 ± 2.4 0.330 ± 0.010 19 ± 0.6 12 ± 0.50
10 107 ± 2.5 0.200 ± 0.016 19 ± 0.9 7 ± 0.070

wt.% SiC, GE-10 wt.% SSA, and GE-1 wt.% Pr composites. The images show that the glass fibers are pulled out from
the fracture surface and microcracks have occurred in plies in the matrix around filler particles [26]. These microcracks
interact with adjacent particles, growing through and around them. The particles bridge the microcracks and absorb a
significant amount the energy dissipating prior to the main failure by the subsequent plastic deformation of the particles
plastic deformation, leading to an increased fracture strength of particulate-filled polymer composites [26, 27].

3.3. Vibration tests

The natural frequencies of the first vibration mode and the damping ratios of the composite laminates considered
are given in Table 4. As is seen, the increasing elastic modulus of composites increased their natural frequencies [28].
Also, the particles positively affected their damping ratios. The natural frequencies of composite specimens increased
with filler content. These results are in agreement with previous studies [11, 14]
In order to figure out the damping properties of the particle-filled composite laminates, their loss and storage
moduli were identified using Eq. 3, and the results are presented in Table 4. It is clearly seen that filler content does not
clearly correlate with the loss modulus — the modulus can increase or decrease with filler concentration [19]. The ad-
dition of 5 wt.% Bx particulates increased the storage modulus by about 14.1% and the loss modulus by about 46.4%.
The addition of 10 wt.% SiC particulates increased the storage modulus by about 5.2%, but 20 wt.% increased the loss
modulus by about 5.94%. The addition of 15 wt.% SSA particulates increased the storage modulus by about 1.48%, but
10 wt.% increased the loss modulus by about 38.6%. The addition of 1 wt.% Pr particulates increased the storage modulus
by about 3.2%, but 5 wt.% increased the loss modulus by about 18.5%. This can be explained by the good dispersion of
the microfillers and their good interaction with the polymeric matrix [19]. In contrary, SSA particulates showed a nega-
tive effect at all their concentrations.

125
The loss modulus is the measure of lost energy [30] and characterizes the energy dissipating capacity of a material.
A high loss modulus means a high degree of energy dissipation before failure [31]. The addition of 5 wt.% Bx slightly
increased the storage modulus because of an improved interaction between the epoxy, borax, and fiber, indicating a better
load transfer between the particles, fiber, and matrix. The greatest increase in the loss modulus was observed at 5 wt.% of Bx.
The highest natural frequencies, of 107, 108, 110, and 107 Hz, were found at the particle content of 20 wt.% for
the GE-Bx, and GE-SiC, and GE-SSA composites (but at 10 wt.% for GE-Pr), with increments of 23.5, 24.7, 27.0, and
23.5%, respectively. The particle type did not significantly affect the first-mode natural frequency of the composites. It is
also found that the damping ratio of particulate-filled fiber-reinforced polymer-matrix composites depends on interfacial
interactions in polymer/filler systems [29]. Both increasing and decreasing damping ratios are observed for particle-
modified GE composite specimens [20, 21]. With addition of 10 wt.% SSA, the damping ratio increased by 46.5%. For
other filler types, the maximum damping ratios were observed at a particulate content of 5 wt.%, namely, 0.355 (28.6%),
0.3 (8.69%), and 0.33(19.5%) for the GE-Bx, GE-SiC, and GE-Pr composites, respectively. The greatest damping ratio
was obtained by adding 10 wt.% of SSA.

4. Conclusion

In this study, a S-glass fiber/epoxy composite modified by microscale particles of four types was prepared and
investigated. In the experimental part, the hand lay-up method was employed at a 0.3-MPa pressure and 80°C temperature.
The effects of perlite, sewage sludge ash, silicon carbide, and borax particulates on the tensile strength and vibration prop-
erties (the natural frequency, damping ratio, and the storage and loss moduli) of the composite were explored for different
filler ratios. The following main conclusions can be drawn from the results obtained.
• The tensile properties of composite were significantly influenced by the type and content of fillers.
• The maximum tensile strength was reached at 10 wt.% of Bx, SiC, and SSA and 1 wt.% of Pr fillers, with improve-
ments of 10.5, 1.2, 8.4, and 13.8%, respectively.
• The lowest tensile strength was obtained at 20 wt.% for the GE-Bx, GE-SSA, and GE-SiC composites and at 10 wt.%
for the Pr-filled composites.
• The strain at failure increased at all filler types and ratios. Their minimum value was found at 20 wt.% of Bx and
SSA and at 10 wt.% of Pr, with exception of SiC, which exhibited the lowest value at 5 wt.%
• The natural frequency increased with increasing filler content of all types of particles ratio. The highest natural fre-
quencies, of 107, 108, 110, and 107 Hz were found at 20 wt.% of Bx, SiC, SSA fillers and 10 wt.% of Pr, with increments
of 23.5, 24.7, 27.0, and 23.5%, respectively.
• Both increasing and decreasing damping ratios of the particulate-filled composite specimens were obtained. The
maximum value of 0.405 was found at 10 wt.% of SSA, showing a 46.5% increase. For the other filler types, the maximum
damping ratios was seen at 5 wt.%, namely, 0.355, 0.3, and 0.33, with increments of 28.6, 8.69, and 19.5%, for Bx, SiC,
and Pr, respectively.
• The storage and loss moduli of the particulate-filled S-glass/epoxy composites exhibited the highest values at 5
wt.% of Bx, showing increments of 15 and 50%, respectively.

REFERENCES

1. S. Sreenivasulu and A. Chennakesava Reddy, “Mechanical properties evaluation of bamboo-fiber-reinforced compos-


ite,” Int. J. Eng. Res., 3, 187-194 (2014).
2.Advances in Diverse Industrial Applications of Nanocomposites, Intech, India, 113-138 (2011).
3. D. Verma, P.C. Gope, A. Shandilya, and A. Gupta, “Coir fiber reinforcement and application in polymer composites:
A review,” J. Mater. Env. Sci., 4, 263-276 (2013).

126
4. S. Ray, Doctoral dissertation, National Institute of Technology Rourkela (2009).
5. R. N. Rothon (Ed.), Particulate-Filled Polymer Composites, 2nd Edition, Smithers Rapra Publishing (2003). 
6. M. Anuara, A. Mat Isaa, and Z. Ummi, “Modal characteristics study of CEM-1 single-layer printed circuit board using
experimental modal analysis,” Proc. Eng., 41, 1360-1366 (2012).
7. I. Lee and D. Kim, “Natural frequency and mode shape sensitivities of damped systems: Part I, Distinct natural fre-
quencies,” J. Sound Vib., 223, 399-412 (1999).
8. S. Vikas Hoysala, P. Kiran Kumar, and T. Madhusudhan, “Static and dynamic behavior of jute-reinforced epoxy
composites with and without silicon dioxide as epoxy modifier: A review,” Int. Res. J. Eng. Tech., 2, 480-484 (2015).
9. M. Bulut, Major Project Report for the Master Degree in Mechanical Engineering, University of Gaziantep, Graduate
School of Natural and Applied Sciences, Turkey (2013).
10. X. D. Spiliotis, K. Ntampegliotis, V. Karayannis, and G. A. Papapolymerou, “Physico-mechanical properties of ex-
truded and sintered ceramics using pet coke and sewage sludge as admixtures,” J. Cer. Process. Res., 16, 1-17 (2015). 
11. M. Kutuk and Z. Oguz , “A Research on Effect of Sewage Sludge Ash on the Mechanical Properties of Composite
Material,” World Congress on Civil, Struct., and Env. Eng. (CSEE’16), 30-31 (2016).
12. A. Patnaik, A. Satapathy, S. Mahapatra, and R. Dash, “Implementation of Taguchi design for erosion of fiber-reinforced
polyester composite systems with SiC filler,” J. Reinf. Plast. Comp., 27, 1093-1111 (2008).
13. O. Asi, “Cracks in a powder vibrating sieve disc,” J. Reinf. Plast. Comp. 28, 2861-2867 (2009).
14. I. Ozsoy, A. Demirkol, A. Mimaroglu, H. Unal, and Z. Demir, “The Influence of micro- and nano-filler content on the
mechanical properties of epoxy composites,” J. Mech. Eng., 61, 601-609 (2015).
15. S. Pichi Reddy, P. Chandra Sekhar Rao, A. Chennakesava Reddy, and G. Parmeswari, “Tensile and flexural strength of
glass fiber epoxy composites,” Frict., 2, 354-364 (2014).
16. G. Agarwal, A. Patnaik, R. Kumar Sharma, and J. Agarwal, “Effect of stacking sequence on physical, mechanical and
tribological properties of glass-carbon hybrid composites,” Int. J. Adv. Struct. Eng. (IJASE), 5, 1-8 (2014).
17. R. Sankar, P. Krishna, V. Rao, and P. Babu, “The effect of natural rubber particle inclusions on the mechanical and
damping properties of epoxy-filled glass fibre composites,” Proc. Inst. Mech. Eng., Part L: J. Mater.: Des. and App.,
224, 63-70 (2010).
18. R. Sankar, R. Srikant, P. Krishna, V. Rao, and P. Babu, “Estimation of the dynamic properties of epoxy glass fabric
composites with natural rubber particle inclusions,” Int. J. Aut. Mech. Eng., 7, 968 (2013).
19. J.Datta and M. Włoch, “Influence of selected submicron inorganic particles on mechanical and thermo-mechanical
properties of unsaturated polyester/glass composites,” J. Reinf. Plas. Comp., 33, 935-941 (2013).
20. M. Pol, A. Zabihollah, S. Zareie, and G. Liaghat, “Effects of nano-particles concentration on dynamic response of
laminated nanocomposite beam,” Mechanika, 19, 53-57 (2013).
21. P. Kumar, K. Karthik, and T. Raja, “Vibration damping characteristics of hybrid polymer matrix composite,” Int. J.
Mech. Mechat. Eng., 15, 42-47 (2015).
22. ASTM Amer. Soc. for Testing and Mater. Standard test method for tensile properties of plastics. D 638-10. Philadel-
phia, PA (2010).
23. M. Bulut, A. Erkliğ, and E. Yeter, “Experimental investigation on influence of Kevlar fiber hybridization on tensile and
damping response of Kevlar/glass/epoxy resin composite laminates,” J. Comp. Mat., 50, 1875-1886 (2016).
24. B. Hulugappa, M. V. Achutha and B. Suresha, “Effect of fillers on the mechanical properties and fracture toughness of
glass-fabric-reinforced epoxy composites,” J. Min. Mat. Charac. Eng., 4, 1-14 (2016).
25. K. S. Ahmed, V. Mallinatha, and S. J. Amith, “Effect of ceramic fillers on mechanical properties of woven jute fabric
reinforced epoxy composites,” J. Reinf. Plast. Comp., 30, 1315-1326 (2011).
26. L. Wang, J. Zhang, X. Yang, C. Zhang, W. Gong, and J. Yu, “Flexural properties of epoxy syntactic foams reinforced
by fiberglass mesh and/or short glass fiber,” Mat. Des., 55, 929-936 (2014).
27. D. Stevanovic, S. Kalyanasundaram, A. Lowe, and P.Y.B. Jar, “Mode I and mode II delamination properties of glass/
vinyl-ester composite toughened by particulate modified interlayers,” Comp. Sci. Tech., 63, 1949-64 (2003).
28. A. Alva and S.Raja, “Damping characteristics of epoxy-reinforced composite with multiwall carbon nanotubes,, Mech.
Adv. Mat. Struct., 21, 197-206 (2014).

127
29. A. Erkliğ, M. Bulut, and E. Yeter, “The effect of hybridization and boundary conditions on damping and free vibration
of composite plates,” Sci. Eng. Comp. Mat., 22, 565-571 (2015).
30. B. Cappella, Manuals in Polymer Science, Springer laboratory (2016).
31. S. Gupta, P. R. Mantena, and A. Al-Ostaz, “Dynamic mechanical and impact property correlation of nanoclay- and
graphite-platelet-reinforced vinyl ester nanocomposites,” J. Reinf. Plast. Comp., 29, 2037-2047 (2010).

128

Potrebbero piacerti anche