You are on page 1of 29

Composite Interfaces

ISSN: 0927-6440 (Print) 1568-5543 (Online) Journal homepage: https://www.tandfonline.com/loi/tcoi20

Flexural properties of multiscale nanocomposites


containing multiwalled carbon nanotubes coated
glass fabric in epoxy/graphene matrix

Abu Mani, Pankaj Tambe & Ariful Rahaman

To cite this article: Abu Mani, Pankaj Tambe & Ariful Rahaman (2019): Flexural properties of
multiscale nanocomposites containing multiwalled carbon nanotubes coated glass fabric in epoxy/
graphene matrix, Composite Interfaces, DOI: 10.1080/09276440.2019.1569396

To link to this article: https://doi.org/10.1080/09276440.2019.1569396

Published online: 06 Feb 2019.

Submit your article to this journal

Article views: 6

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tcoi20
COMPOSITE INTERFACES
https://doi.org/10.1080/09276440.2019.1569396

ORIGINAL ARTICLE

Flexural properties of multiscale nanocomposites containing


multiwalled carbon nanotubes coated glass fabric in epoxy/
graphene matrix
Abu Mania, Pankaj Tambe b
and Ariful Rahamana
a
School of Mechanical Engineering, VIT University, Vellore, India; bSchool of Mechanical Engineering, VIT-
AP University, Andhra Pradesh, Amravati, India

ABSTRACT ARTICLE HISTORY


In this work, graphene is prepared by reduction of graphene Received 20 October 2018
oxide. The graphene obtained is used as the reinforcing nano- Accepted 8 January 2019
filler in an epoxy matrix. The incorporation of 0.1 wt.% loading KEYWORDS
of graphene in an epoxy matrix shows the maximum increase in Graphene; epoxy;
the flexural properties as compared to the pure epoxy. It is due nanocomposites; glass fabric
to good dispersion of graphene and interaction between
C-N group of epoxy and C=O group of graphene. The loading
of 0.1 of wt.% graphene in epoxy is used to prepare the glass
fabric reinforced multiscale nanocomposites. This multiscale
nanocomposite shows an enhancement in flexural properties.
It is due to the mechanical interlocking associate with random
dispersion of graphene in epoxy resin. In addition, glass fabric is
coated with polyvinylpyrrolidone (PVP) modified multiwalled
carbon nanotubes (MWNT) and its multiscale nanocomposites
show the enhanced flexural strength. The increase in flexural
strength is due to mechanical keying associated with the rough
surface obtained with the coating of MWNT over glass fabric.
Further, with an addition of 0.1 wt.% of graphene and MWNT
coated glass fabric in epoxy multiscale nanocomposites, flexural
properties enhancement are highest. It is due to the combined
effect of mechanical keying and mechanical interlocking asso-
ciated with graphene dispersion and rough surface of MWNT
coat over glass fabric. In addition, the toughness of the MWNT
coated glass fabric and graphene filled epoxy multiscale nano-
composites increases due to the mobile epoxy chains contribut-
ing towards the energy absorption. It is observed that the
epoxy chains in contact with PVP coated MWNT are mobile as
confirmed from dynamic mechanical analysis (DMA) studies.
Thus, due to mechanical interlocking, the mobility of epoxy
chains and mechanical keying; the flexural modulus, flexural
strength, flexural toughness, and storage modulus obtained
are highest for the graphene and MWNT coated glass fabric in
epoxy multiscale nanocomposites.

CONTACT Pankaj Tambe pankaj.tambe@vitap.ac.in


© 2019 Informa UK Limited, trading as Taylor & Francis Group
2 A. MANI ET AL.

1. Introduction
Glass fiber reinforced polymer (GFRP) composites were of under commercial exploita-
tion in the aerospace industry for many decades [1,2]. The glass fiber acts as
a reinforcing material and the polymer is used as a matrix material. In the aerospace
industry, glass fiber is generally used in the form of fabric. For a fabric preparation, the
multiple glass fibers are grouped together in the form of yarn and converted into
a laminate for composites manufacturing [1]. These laminates are used to prepare
composites using various polymer matrix composites manufacturing processes. The
various types of polymeric materials are used as a matrix material. Among various
polymers used as a matrix material, epoxy is one of the widely used material in the
aerospace industry [3–5]. The properties of processed GFRP composites depends on the
strong interaction between the fiber and the epoxy matrix [6–8]. Since the glass fiber is
the main load bearing element of GFRP composites. There are two main reasons for
GFRP composites failure. The first major reason is delamination occurrence at the
fiber-matrix interface [7,8]. The other main reason for the fracture of GFRP composites
is the matrix phase cracking [8].
In view of matrix phase cracking, nanofillers has attracted the attention of materials
scientist to manufacture advanced GFRP composites. There were many studies carried out
using different nanofillers like carbon nanotubes (CNTs), graphene and clay mixed with the
epoxy phase of GFRP composites. A brief review of the recent literature in this area of
research is tabulated in Table 1. The majority of the paper concentrates on multiwalled
carbon nanotubes (MWNT) filled epoxy and its GFRP nanocomposites [9–28]. Gojny et al.
[9] used amine functionalized double-walled carbon nanotubes (NH2-DWNT) in the
epoxy matrix to improve the mechanical properties like tensile properties and fracture
toughness (K1c) at a low content of NH2-DWNT of GFRP nanocomposites. There is a 41%
enhancement in K1c with an addition of 0.3 wt.% NH2-DWNT in GFRP nanocomposites.
Warrier et al. [12] and Zeiler et al. [25] achieved ~10% enhancement in K1c with the
addition of pristine and surface modified MWNT in GFRP nanocomposites. DWNT is
more effective in K1c enhancement as compared to MWNT, but the cost of DWNT is more.
Table 1 clearly reveals that the tensile, flexural properties enhance with addition of
Table 1. Literature review.
Wt.% of Nanofillers in Epoxy Modification of Glass
Sr. No Matrix Fiber Properties Enhancement (%) Remarks References
1 MWNT - IFSS – 27% Wettability studied [12]
2 MWNT (0.5%) - K1c – 10% Sizing with MWNT/epoxy [13]
3 MWNT (0.1%) - TM – 25%, TS – 41%, FM – 49%, Dispersion using high shear homogenizer [14]
FS – 74%
4 MWNT (0.3%) - ILSS – 16% Dispersion using mini calender [15]
5 MWNT G1c – 76% Aligned MWNT [19]
6 MWNT - IFSS – 100% Optimum amine-epoxide ratio [22]
7 MWNT (0.1%) FM – 20%, FS – 32% [27]
8 DWNT-NH2 (0.3%) - TM – 13%, TS – 5%, K1c – 41% Anisotropic electrical conductivity observed [9]
9 O-MWNT (2%) - ILSS – 33% [10]
10 MWNT-NH2 (0.3%) - FM – 77%, FS – 37%, Dispersion using 3-roll mill [16]
11 MWNT-OH (0.5%) - TM – 4%, Toughness – 12% Tensile Modulus decreases at 0.5 wt.% of MWNT-OH [23]
12 f-MWNT (0.5%) - K1c – 9.5, G1C – 8%, GIIC – 6% [25]
13 NH2-MWNT (0.5%) - FM – 24% [28]
14 MWNT (1.4 vol.%)+ INK (5.6 vol.% - FM – 13%, FS – 105% Reduced the weight of composite structures [24]
15 GNP (0.25%) - FS – 30% 25% increase in Fracture toughness [26]
16 Clay (1%) - ILSS – 44%, FS – 24%, K1c – 23% S2-glass fiber [21]
17 OMMT - TM – 12.5%, TS – 7.5% Bidirectional laminate [20]
18 Clay (3%) FM – 21%, FS – 27%, IS- 31% Unidirectional GF used [11]
19 Clay (5%) - FM – 10% Solution mixed clay better dispersed in epoxy [17]
20 OMMT (6%) - FM – 12.5% Modified clay [18]
21 - GO-g-GF ILSS – 44% Grafting of GO over GF [29]
22 - GO-c-GF ISS – 219% EPD of GO over GF [30]
23 - Gr+CNT - EPD coating for damage detection [31]
24 - MWNT coated over GF - EPD and dip coated MWNT on GF. EPD coated MWNT on GF showed [32]
improved properties of fibers.
25 - MWNT coated over GF - Interfacial sliding observed at MWNT coated over GF in an epoxy [33]
matrix.
26 o-MWNT (0.3wt%) o-MWNT-g-GF - Flexural toughness increases with modification of GF from o-MWNT [34]
and epoxy+o-MWNT matrix.
27 - ZnO nanowires - TM increases and interfacial strength increases by 109% [6]
28 - f-MWNT coating over - ILSS, TS and FS improved with surface coating of functionalized MWNT [35]
MWNT over GF.
COMPOSITE INTERFACES

DWNT-NH2 – Double-walled carbon nanotubes functionalized with amine, MWNT – Multi-walled carbon nanotubes, O-MWNT – Oxidized MWNT, MWNT-NH2 – Multi-walled carbon nanotubes
functionalized with amine, MWNT-OH – Multi-walled carbon nanotubes functionalized with hydroxyl group, f-MWNT – Multi-walled carbon nanotubes functionalized, GNP – Graphene
nanoplatelets, GF – Glass fiber, GO – Graphene oxide, Gr – Graphene, GO-g-GF – GO grafted GF, GO-c-GF – GO coated over GF, EPD – Electrophoretic deposition, TM – Tensile Modulus, TS –
Tensile Strength, FM – Flexural Modulus, FS – Flexural Strength, ILSS – Interlaminar shear strength, IFSS – Interfacial shear strength, K1c – Fracture Toughness.
3
4 A. MANI ET AL.

functionalized or surface modified CNTs in GFRP nanocomposites [9–23]. Interlaminar


shear strength (ILSS) enhance by 33% with an addition of surface modified MWNT in
GFRP nanocomposites [10]. Haque et al. [21] achieved 23% enhancement in K1c and 44%
in ILSS with an addition of 1 wt.% clay in GFRP nanocomposites. Thus, the inclusion of
nanoparticles in epoxy matrix restricts the matrix phase cracking of the GFRP.
The delamination occurrence at the fiber-matrix interface of GFRP is a major
challenge for effective commercial utilization. The delamination occurrence at the fiber-
matrix interface can be restricted with tailoring strong interaction between fiber and
matrix by depositing a layer of nanoparticles over the glass fabric. The CNTs are
extensively used for the modification of glass, carbon and Kevlar fabric [1]. The
CNTs are deposited using three methods viz. (a) Electrophoresis deposition (EPD)
(b) CVD growth and (c) Sizing method. In the CVD method of deposition of CNTs,
fiber has to initially coat with the catalyst and later the growth of CNTs takes place at
a higher temperature. While, in the case of electrophoresis deposition, the process is
simple but the distribution of CNTs on the surface is random. The sizing method is the
simplest one, where the sizing agent is mixed and CNTs in a solution followed by
dipping of fabric and drying the solution. The sizing method is relatively simple to
implement as compared to electrophoresis and CVD method of coating of CNTs.
Among the single, double and multiwalled carbon nanotubes (SWNTs, DWNTs, and
MWNTs), MWNTs are commercially cheaper and available at a large scale globally.
A brief review of CNTs, graphene oxide and graphene-coated glass fiber reinforced
epoxy nanocomposites are given in Table 1 [29–35]. Zhang et al. [32] coated MWNT
over single glass fiber using EPD and dip coating. They found that EPD coated single
glass fiber in epoxy matrix shows higher interfacial strength as compared to dip coated
single glass fiber in an epoxy matrix. Zeng et al. [35] grafted MWNT over GF and its GF
nanocomposites show the improves tensile, flexural and ILSS properties. An et al. [33]
observed that interfacial sliding occurs by coating MWNT over GF and reinforced in an
epoxy matrix. Eskizeybek et al. [34] observed that by incorporation of MWNT in epoxy
and coating over GF show increase in flexural toughness of nanocomposites. Recently,
Swaminathan et al. [6] shows the growth of ZnO nanowires over glass fiber shows an
increase in tensile and interfacial properties by incorporation with epoxy matrix. In
addition, several attempts were made to coat graphene oxide and graphene over GF to
improve the mechanical properties of nanocomposites.
The literature review discussed so far reveals the incorporation of nanoparticle in the
epoxy matrix improves the mechanical properties of GFRP nanocomposites by restricting
matrix phase cracking [1–38]. The graphene addition in an epoxy matrix and its GFRP
nanocomposites are not probed extensively. Graphene has attracted the attention of the
aerospace industry in specific due to the unmatched superior mechanical properties as
compared to other nanomaterials [39]. In specific, the mechanical properties of graphene
can be exploited as one of the filler materials for the epoxy-based aerospace structural
materials in enhancing its mechanical properties. There are many studies available on the
incorporation of graphene in the epoxy matrix to enhance the mechanical properties of
epoxy nanocomposites [4,5]. These studies show the modification of graphene and its
interaction with polymer matrix contributes towards the enhancement in mechanical
properties of epoxy nanocomposites [4,5]. In addition to this, the pristine mechanical
properties of graphene depend on the method of synthesis of graphene [40–42]. The
COMPOSITE INTERFACES 5

superior mechanical properties like Young’s modulus of graphene measured as 1 TPa


earlier were synthesized using the scotch tape method [39]. Recently, Zhang et al. [43]
reported the chemical vapor deposited (CVD) graphene has Young’s modulus of 1 TPa and
that of chemically reduced graphene of 0.25 TPa. The reason for lower Young’s modulus of
0.25 TPa of chemically reduced graphene is the creation of defects during the reduction of
graphene oxide. In addition, the functional group on graphene oxide was not completely
reduced [4]. The mass production of graphene using chemical method is relatively simple
and can be scaled up [44]. Even, Young’s modulus of 0.25 TPa of chemically reduced
graphene is lower, but the functional group on the edges helps in the modification of
graphene to tailor the interface between epoxy and graphene. Thus, a chemical method of
synthesis of graphene can be of commercial interest to the aerospace industry.
The literature review discussed so far reveal the importance of graphene and CNTs
as a nano-fillers in the aerospace industry [12–38]. The graphene incorporation in
epoxy resin and its GFRP is less explored. Thus, we have decided to use graphene as
nanofiller to be added in an epoxy matrix. The coating of nanoparticle over GF
improves the interfacial properties GFRP nanocomposites [29–35]. Deposition of
MWNTs over glass fabric is extensively studied as depicted from the literature [1].
Thus, we have decided to use MWNTs as nanofiller to coat over glass fabric. MWNT is
coated with polyvinyl pyrrolidone (PVP); since PVP has an interaction with epoxy
matrix [3,4]. In addition, PVP as surfactant deagglomerate the MWNTs in aqueous
solution using sonication followed by deposition over glass fabric by dip coating. The
other important papers in this area of research are cited in the references [45–79].
The graphene used is synthesized using the chemical method and CNTs used are
purchased commercially. The epoxy used is diglycidyl ether of bisphenol-A (DEGBA)
and the hardener is an amine-based hardener. The glass fabric used is the bidirectional
type and tightly weaved. This work was carried out in three stages. The three stages are;
(a) Mixing of graphene in an epoxy matrix and optimize the concentration of graphene,
(b) Coating of MWNTs over the glass fiber and (c) Processing of epoxy/graphene/glass
fabric multiscale nanocomposites with and without a coating of MWNTs. The various
characterization techniques were used to characterize graphene, MWNTs coated glass
fabric, epoxy/graphene nanocomposites and epoxy/graphene/glass fabric multiscale
nanocomposites with and without a coating of MWNTs.

2. Experimental
2.1. Materials
The various materials used in this study are listed according to type, manufactures, and
characteristics in Table 2.

2.2. Synthesis of graphene oxide and graphene


The graphene oxide was synthesized according to the improved Hummers method. In
a round bottom flask, 3 grams of graphite powder was mixed with 360 ml of concen-
trated sulphuric acid (H2SO4) using a magnetic stirrer followed by addition of 40 ml of
phosphoric acid (H3PO4). Thereafter, 18 grams of potassium permanganate (KmNO4)
6 A. MANI ET AL.

Table 2. Materials used in the work.


Sr. No Materials Company Characteristics
1 Graphite Flakes Alfa Aesar, India ~325 mesh, 99.8% purity
2 Concentrated Sulphuric Acid S. D. Fine Chemicals, India Analytical Grade
3 Phosphoric Acid S. D. Fine Chemicals, India Analytical Grade
4 Hydrochloric Acid S. D. Fine Chemicals, India Analytical Grade
5 Potassium Permanganate S. D. Fine Chemicals, India 99.8% purity
6 Hydrogen Peroxide S. D. Fine Chemicals, India Analytical Grade
7 Diglycidyl ether of bisphenol-A (DGEBA) Atul Ltd. (India) 9000−12,000 mPa.s at 25°C
8 N,N-Bis(2-aminoethyl) ethane-1,2-diamine Atul Ltd. (India) 10 mPa.s at 25°C
9 Multiwalled Carbon Nanotubes Nanocyl, Belgium NC7000
10 Polyvinylpyrrolidone Sigma Aldrich Average mol. Wt. 40,000
11 Sodium Hydrosulphite Sigma Aldrich 99.7% purity
12 E-Glass fabric Suntech Fibre Pvt. Ltd India 200 gsm

was slowly added to the flask and the reaction was maintained in an ice bath. The entire
reaction was maintained at 55°C for 12 hours. Further, distilled water was added slowly
(400 ml), followed by addition of hydrogen peroxide (3 ml) dropwise. The graphene
oxide was filtered and washed with distilled water, hydrochloric acid (HCl) and ethyl
alcohol. The reduction of graphene oxide was carried out using sodium hydrosulfite.
The yield of graphene oxide after filtration was mixed with 400 ml of water using bath
sonication followed by addition of 8 grams of sodium hydrosulfite. The graphene
obtained was centrifuged and washed with ethyl alcohol followed by drying in
a vacuum oven at 80°C.

2.3. Processing of epoxy/graphene nanocomposites


The schematic of epoxy/graphene nanocomposites is shown in Figure 1. The graphene
was mixed with ethyl alcohol and sonicated for 30 minutes using a probe sonicator.
Then the mixture of ethyl alcohol and graphene was mixed with epoxy resin using
a magnetic stirrer. The magnetic stirrer is maintained at 80°C for 2 hours to evaporate
the ethyl alcohol. The obtained mixture of epoxy/graphene was ball milled at 200 rpm
for 2 hours. Thereafter, the epoxy/graphene mixture is degassed in a vacuum oven for
30 minutes. The degassed epoxy/graphene mixture was added with the amine-based
hardener with the ratio of epoxy: hardener as 10:1 (Vol/Vol) and stir slowly to mix
followed by degassing for 10 minutes. Then, the solution was cast in an open mold for
24 hours. The cured epoxy/graphene nanocomposites were further post-cured at 70°C
for 7 hours in a hot air oven. The graphene concentration in the epoxy matrix was
maintained at 0.05, 0.1 and 0.2 wt.% respectively.

2.4. Modification of glass fabric with MWNT


The E-glass fabric of 200 gsm was coated with MWNT. The aqueous solution of PVP
was prepared to have PVP concentration in solution as 1 mg/ml. Then, MWNT was
added into the aqueous PVP solution and probe sonicated for 30 minutes. Then, the
glass fabric was dipped into the solution of PVP modified MWNT. The glass fabric was
dipped in the solution of PVP modified MWNT for 30 minutes followed by drying the
coated PVP-MWNT in a hot air oven.
COMPOSITE INTERFACES 7

Figure 1. Schematic of processing of epoxy/graphene nanocomposites.

2.5. Processing of glass fabric filled epoxy/graphene multiscale nanocomposites


The glass fabric filled epoxy composites were prepared using the Hand-Layup method
followed by cold pressing and post-curing. In the Hand-Layup method, a layer of glass
fabric was placed on the workbench and a layer of epoxy and hardener mixture was
poured. Then the epoxy and hardener mixture was rolled with the help of roller over glass
fabric. The process of layering of glass fabric and epoxy were repeated until the 12th layer.
Further, vacuum bagging was done by applying a pressure of 1 bar with a vacuum pump.
The vacuum bagged glass fabric filled epoxy composites were pressed in a hydraulic press
at 5 MPa and pressure was maintained till final curing to occur. The glass fabric filled
epoxy composites were post-cured in hot air at 70°C for 7 hours in a hot air oven. The
same procedure was used to prepare MWNT coated glass fabric epoxy graphene filled
multiscale nanocomposites; epoxy/glass fabric multiscale nanocomposites; MWNT coated
glass fabric and it’s graphene/epoxy multiscale nanocomposites.

2.6. Characterization
The graphene and graphene oxide was characterized using X-ray diffraction (XRD)
setup of BRUKER, Germany (D8 Advance). The XRD patterns were recorded in the
range of diffraction angle of 5° and 80° (2θ). The wavelength (λ) of an X-rays is
1.54 nm. The step scan was recorded with a step size of 0.02°. Further, the
transmission electron microscope (TEM) studies were carried out over graphene
using TEM of Tecnai, G2 20 Twin. For the preparation of TEM sample of graphene,
8 A. MANI ET AL.

2 mg of graphene was mixed with isopropyl alcohol using bath sonicator for
10 minutes and a drop is poured over the copper grid. The copper grid is heat at
an elevated temperature to evaporate the isopropyl alcohol. The TEM images of
graphene were taken on the dried copper grid. In addition, the microtome epoxy/
graphene nanocomposites sample was used to study the dispersion of graphene in
TEM. The flexural testing was carried over the epoxy/graphene nanocomposites and
multiscale nanocomposites. The tests were conducted as per ASTM D790 on
a universal testing machine manufactured by Instron (8801). Thermogravimetric
analysis (TGA) was carried over the epoxy/graphene nanocomposites samples on
TA instruments Q500. The operating temperature of TGA is 50–900°C @ 10°C/min
under nitrogen atmospheric condition. The flexural fractured epoxy/graphene nano-
composites and MWNT coated glass fabric were characterized using a scanning
electron microscope (SEM) of ZEISS (EVO18). Fourier transform infrared (FTIR)
spectroscopy was carried out over graphene oxide, graphene and epoxy/graphene
nanocomposites using IRAffinity-1, SHIMAZU. The spectra were recorded in the
scanning range of 400 to 4000 cm−1 and at a resolution 0.5 cm−1. The dynamic
mechanical analysis (DMA) of epoxy/graphene nanocomposites and multiscale nano-
composites was carried out using Anton Paar MCR 102 rheometer in a torsional
mode at a frequency of 1 Hz and in the temperature range of 25 to 200°C.

3. Results and discussion


3.1. Characterization of graphene
In this work, graphene was synthesized using chemical oxidation of graphite followed
by chemical reduction method of graphene oxide. The synthesis of graphene oxide was
carried out using an improved method developed by Tour group [44]. Further, the
reduction of graphene oxide was carried out using sodium hydrosulfite as a reducing
agent. The obtained graphene oxide and graphene are characterized using XRD. XRD is
the simplest characterization tool to confirm the formation of graphene oxide and
graphene; as the different d-spacing of graphene oxide and graphene are different.
Figure 2 shows the XRD spectra of graphene oxide and graphene respectively. The XRD
spectra of graphene oxide show a peak centered at 10° depicting the (001) plane [2,44].
While the reduction of graphene oxide with sodium hydrosulfite shows an XRD peak
centered on 23° depicting the (002) plane of graphene. The d-spacing of graphene oxide
and graphene can be calculated using Bragg’s law: λ = 2dsin(θ), where, λ is the
wavelength of X-ray, d is the d-spacing, and θ is the diffraction angle. The calculated
value of d-spacing using Bragg’s law for graphene oxide and graphene are 0.84 nm and
0.38 nm respectively. After confirmation of graphene oxide and graphene using XRD,
the shape and size of graphene nanosheets prepared can be probed using TEM. Figure 3
shows the TEM micrographs of graphene taken by dropping a drop of sonicated
graphene on the copper grid followed by evaporation of the solvent. The TEM images
of graphene exhibit the wrinkled structure. The wrinkling of graphene is due to
buckling of graphene nanosheets on a metal substrate like copper [80]. In addition,
there is folding/curling of graphene nanosheets at the edges. The folding/curling of
COMPOSITE INTERFACES 9

GO
Graphene

Intensity (a.u)

10 15 20 25 30 35
2θ (degree)
Figure 2. XRD of graphene oxide and graphene.

(a) (b)

Figure 3. TEM of graphene.

graphene is due evaporation of solvent; and on the basis of energy criteria, there is
a surface energy difference between graphene and solvent [2].
In order to probe the presence of functional groups on graphene, FTIR spectroscopy
was performed over graphene oxide and graphene respectively. Figure 4(a) shows the
FTIR spectra of graphene oxide and Figure 4(b) shows the FTIR spectra of graphene.
The Figure 4(a) of graphene oxide shows the three prominent peaks marked as
carbonyl, hydroxyl, and epoxy group respectively. The presence of a peak at the
wavelength of 860 cm−1 is due to asymmetric stretching of the epoxy group [81–85].
In addition, the peak centered around 1170 cm−1 is the peak of the hydroxyl group and
1710 cm−1 is the peak of the carbonyl group respectively [81–85]. These peaks of
graphene oxide undergo reduction reaction while reacting with hydrazine. The reduced
10 A. MANI ET AL.

(a) (b)

% Transmission
% Transmission

-C=O
C-O
C=O

Epoxy
-OH
O-H

2000 1800 1600 1400 1200 1000 800 3500 3000 2500 2000 1500

Wavenumber (cm-1) Wavenumber (cm-1)

Figure 4. FTIR of (a) graphene oxide (b) graphene.

graphene oxide FTIR spectra in Figure 4(b) shows the peaks at 1635 cm−1 and
1132 cm−1 assigned to carbonyl and C-O stretching respectively [40]. In addition,
there is a peak at 3450 cm−1 depicting the presence of O-H group [83]. Thus, from
FTIR spectra of graphene, it is clear that graphene is not completely reduced. Due to the
incomplete reduction of graphene oxide, the d-spacing achieved of graphene is 0.38 nm.
The presence of a functional group on graphene can have an influence on the mechan-
ical properties of epoxy nanocomposites and is discussed subsequently.

3.2. Characterization of epoxy/graphene nanocomposites


The graphene nanosheets synthesized using the improved method of obtaining gra-
phene oxide followed by chemical reduction method of graphene oxide was used as
a reinforcing filler to process epoxy nanocomposites. In this work, the processing
protocol of epoxy nanocomposites processing involves the use of ball milling operation
to mix the epoxy and graphene respectively. The ball mill was rotated at a rotational
speed of 200 rpm. It was observed that at the rotational speed of 200 rpm; the shear
forces acting over the graphene nanosheets are highest when the balls strike the
graphene nanosheets [86]. Since, during ball milling operation, a huge amount of
shear stresses are applied over the graphene nanosheets; it is expected that the shear
stress is not only responsible for the exfoliation of graphene nanosheets but also breaks
the graphene nanosheets. To probe the breakage if graphene nanosheets in epoxy
matrix qualitatively, we have carried out TEM analysis of epoxy/graphene nanocompo-
sites samples. Figure 5 shows the TEM image of microtome epoxy/graphene nanocom-
posites samples. The encircled graphene is shown in Figure 5 shows the graphene
nanosheets size is in tens of nanometres. It is noted that the graphene obtained using
the improved method followed by chemical reduction shows the graphene size in
hundreds of nanometers (Figure 3). Thus, the ball milling of epoxy/graphene mixture
results in the breakage of graphene nanosheets.
It was established that the graphene size reduction takes place during the processing
of epoxy/graphene nanocomposites samples. And, from the FTIR spectroscopic char-
acterization of graphene, it was clear that there exist the functional group attached with
COMPOSITE INTERFACES 11

(a) (b)

Figure 5. TEM of epoxy/graphene nanocomposites.

the graphene. These functional groups may interact with the epoxy matrix. To probe
the interaction of a functional group of graphene with the epoxy matrix, FTIR spectro-
scopy was carried out. Figure 6 shows the FTIR spectra of graphene, epoxy and epoxy/
graphene nanocomposites. Pure epoxy matrix sample shows a peak at 1611 cm−1 which
arises due to C-N stretching of the epoxy matrix [87]. While graphene shows a peak at
1635 cm−1 which arise due to C=O stretching. The addition of graphene in epoxy
matrix shows a peak shift in the peak of a C-N stretch to 1622 cm−1 revealing an
interaction between graphene and epoxy matrix.
The dispersion of graphene and the interaction of graphene and epoxy can restrict
the movement of epoxy under an applied load. It is characterized with the help of DMA
characterization on epoxy and epoxy/graphene nanocomposites samples. Before

GNP
E
E0.05GNP
% Transmission

1611 1635

1200 1300 1400 1500 1600 1700 1800


Wavenumber (cm-1)
Figure 6. FTIR of graphene, epoxy and epoxy/graphene nanocomposite.
12 A. MANI ET AL.

discussing the DMA results, it is important to briefly review the state of literature in the
epoxy/graphene system. Several researchers observed that there is an interface created
between the graphene and epoxy matrix due to adsorption of epoxy chains over the
graphene [88]. In addition, the epoxy chains are interaction with the functional group
of graphene. These conditions influence the glass transition temperature (Tg) of epoxy.
Figure 7(a) shows the graphene substrate over which how epoxy chains can possibly
arrange. It shows the creation of an interface layer at the graphene-epoxy interface. The
interfacial layer creation is due to the strong interfacial interaction between graphene
and epoxy [89]. The interaction is of repulsive in nature with a π-π stacking of π
electron cloud of graphene and benzene ring of epoxy [89,90]. In addition, there are
ends of epoxy chains interact with the graphene restricting the curing reaction [4].
Further, there may be the presence of unreacted chain away from the interface of epoxy
and graphene [89]. Figure 7(b) shows the tanδ curves of epoxy and epoxy/graphene
nanocomposites samples. The peak value of tanδ curves is revealing the Tg of the
samples. The Tg of epoxy/graphene nanocomposites samples increases with increase in
the concentration of graphene in epoxy as compared to pure epoxy sample. At low
concentration of 0.1 wt.% of graphene in epoxy, the Tg of epoxy/graphene nanocom-
posites increase is not significant, while at 0.2 wt.% of graphene in epoxy, the Tg of
epoxy/graphene nanocomposites increase is significant. The DMA test epoxy/graphene
nanocomposites were carried out in torsion mode. Let us try to interpret the observa-
tion with the help of schematic 4.7c. In Figure 7(b), it shows the random distribution of
graphene nanosheets in an epoxy matrix. The random distribution of graphene in

(a)

0.8
(b) E
(c)
E0.1GNP
E0.2GNP
0.6
tan δ

0.4

0.2

0.0
60 75 90 105 120 135
Temperature (OC)

Figure 7. (a) Schematic of interface of epoxy/graphene nanocomposites (b) tanδ as a function of


temperature of epoxy and epoxy/graphene nanocomposites (c) Schematic of distribution of gra-
phene in epoxy/graphene nanocomposites.
COMPOSITE INTERFACES 13

epoxy is well reported in the literature [88]. The creation of an interface between
graphene and epoxy makes the Tg value to be decreased [4]. But the random distribu-
tion of graphene restricts the interface movement under the torsional loading, thereby
increasing the Tg of epoxy/graphene nanocomposites. At 0.2 wt.% of graphene in the
epoxy matrix has more graphene nanosheets, which restricts the epoxy chain movement
more and increase the Tg significantly as compared to the 0.2 wt.% of graphene in an
epoxy matrix [91]. In addition, the interaction between the functional group of gra-
phene and epoxy also contributes towards the increase in Tg.
So far, we have observed that there exist an interaction between the functional group
of graphene with epoxy matrix; and the presence of strong interfacial inter-facial
interaction between graphene and epoxy due to interface formation. These interactions
can significantly enhance the mechanical properties of epoxy/graphene nanocompo-
sites. In this regard, the flexural testing was carried over the pure epoxy and epoxy/
graphene nanocomposites. Figure 8 shows the flexural modulus and flexural strength of
the pure epoxy and epoxy/graphene nanocomposites samples. The flexural modulus of
pure epoxy is 2.95 GPa. With incorporation of 0.05 wt.% graphene in the epoxy matrix
shows an increase in flexural modulus. Further, the addition of 0.1 wt.% graphene in the
epoxy matrix again enhances flexural modulus as compared to 0.05 wt.% graphene in
an epoxy matrix. At higher concentration of 0.2 wt.% of graphene in an epoxy matrix,
there is a decrease in flexural modulus as compared to the 0.1 wt.% graphene reinforced
epoxy nanocomposites. It is observed that the highest enhancement in flexural modulus
is obtained at the 0.1 wt.% of graphene in an epoxy matrix. The increase in flexural
strength of 0.05 and 0.1 wt.% graphene in the epoxy matrix are similar as compared to
pure epoxy, while at higher concentration of 0.2 wt.% of graphene in epoxy matrix
show a decrease in flexural strength as compared to the low concentration of graphene
in an epoxy matrix. The quantified values of increase in flexural modulus as; pure epoxy
has the flexural modulus value of 2.9 ± 0.05 GPa which is increased to 3.7 ± 0.05 GPa
(increased by ~27% for 0.05 wt.% of graphene), 3.9 ± 0.07 GPa (increased by ~34% for
0.1 wt.% of graphene), and to 3.8 ± 0.05 GPa (increased by ~31% for 0.2 wt.% of
graphene). The quantified values of increase in flexural strength as; pure epoxy has the
flexural strength value of 105 ± 1.3 MPa which is increased to 110 ± 1.8 MPa (increased

(a) (b)
4.5
Flexural Strength (MPa)

E
Flexural Modulus (GPa)

E
E0.05GNP 120 E0.05GNP
E0.1GNP E0.1GNP
4.0 E0.2GNP
E0.2GNP
105

3.5 90

3.0 75

60
2.5

Figure 8. (a) Flexural modulus and (b) Flexural strength of epoxy and epoxy/graphene
nanocomposites.
14 A. MANI ET AL.

by ~27% for 0.05 wt.% of graphene), 108 ± 1.5 MPa (increased by ~34% for 0.1 wt.% of
graphene), and to 81 ± 1.6 MPa (increased by ~31% for 0.2 wt.% of graphene).
The experimentally obtained flexural modulus of epoxy/graphene nanocomposites
needs to be compared with theoretically predicted values. The modified Halpin-Tsai
model has been used to predict the flexural modulus for two-dimensional nanosheets in
epoxy nanocomposites [89]. The modified Halpin-Tsai model equations are:
 
3 1 þ ððW þ lG Þ=tG ÞηL VG 5 1 þ 2ηT VG
EC ¼ EP þ (1)
8 1  ηL VG 8 1  ηT VG

ðEG =Ep Þ  1
ηL ¼ (2)
ðEG =Ep Þ þ 2lG =tG

ðEG =Ep Þ  1
ηT ¼ (3)
ðEG =Ep Þ þ 2
Where, Ec is the flexural modulus of epoxy/graphene nanocomposites. EP is the flexural
modulus of pure epoxy and EG are the flexural modulus of pure epoxy graphene. The
reported value of the flexural modulus of graphene is 1 TPa [39]. The graphene shape is
referred to as; length is lG, width is W, the thickness is tG. The VG is the volume fraction
of graphene in an epoxy matrix. From the TEM observation of graphene in an epoxy
matrix, the length and width of graphene were calculated as 100 and 80 nm respectively.
Figure 9 shows the comparison of experimentally obtained and theoretically pre-
dicted flexural modulus using modified Halpin-Tsai model of epoxy/graphene nano-
composites. At 0.05 wt.% concentration of graphene in the epoxy matrix shows the
flexural modulus is agreed with the theoretically predicted flexural modulus using
modified Halpin-Tsai model. This observation depicts that the graphene is well dis-
persed in the epoxy matrix. However, at a relatively high concentration of graphene (0.1
Flexural Modulus (GPa)

7 Experimental
Theoretical
6

0.00 0.08 0.16


Graphite nanoplatelets (wt.%)
Figure 9. Flexural modulus of experiential obtained and theoretically predicted epoxy/graphene
nanocomposites.
COMPOSITE INTERFACES 15

and 0.2 wt.%) of graphene in an epoxy matrix, the flexural modulus do not agree with
the theoretically predicted flexural modulus using modified Halpin-Tsai model. This
observation depicts the aggregation of graphene in an epoxy matrix. It is noted that the
dispensability of graphene depends on the interaction with the surrounding medium. In
this work, graphene was dispersed initially in ethyl alcohol using probe sonication. The
dispersibility of graphene in ethyl alcohol is limited and depends on the amount of
sonication energy provided. Since graphene was dispersed in 50 ml ethyl alcohol, so
lower concentration of graphene in ethyl alcohol yield higher exfoliation of graphene,
while at higher concentration of graphene in ethyl alcohol yield large aggregates of
graphene nanosheets. Thus, it can be concluded that the 0.05 wt.% of graphene is well
dispersed in epoxy matrix due to well exfoliation of graphene nanosheets in ethyl
alcohol and subsequent mixing with epoxy using ball milling.
Fractography study can reveal the details about the failure mechanism of epoxy/
graphene nanocomposites. Figure 10 shows the flexural fractured SEM images of
epoxy and epoxy/graphene nanocomposites. The flexural fractured image of pure
epoxy is shown in Figure 10(a). The failure pattern of pure epoxy shows the parallel
lines on the fracture surface. This shows that the there is no resistance to fracture of
epoxy. This behavior is a typical signature of the brittle fracture. On the other hand,
the flexural fractured SEM image of epoxy/graphene nanocomposites shows the small
facets. The facet formation is a signature of resistance offered by the graphene to the
fracture of the epoxy chain [3,4].
Thus, the discussion so far on the epoxy/graphene nanocomposites reveals that
the graphene incorporation in epoxy matrix increases the flexural properties of the
epoxy/graphene nanocomposites. The increase is due to the interaction of graphene
and epoxy matrix, as confirmed by the FTIR studies. In addition, the dispersion of
graphene resists the deformation of the epoxy matrix, as confirmed by fractography
and DMA studies. Theoretical predictions of flexural modulus show the aggregates
of graphene nanosheets are higher at concentration of 0.1 wt.% of graphene and
more in an epoxy matrix.
At last, the thermogravimetric analysis (TGA) study of epoxy and epoxy/graphene
nanocomposites was carried out. Figure 11 shows the thermogravimetric (TG) and differ-
ential thermogravimetric (DTG) curves of epoxy and epoxy/graphene nanocomposites.

(a) (b)

Figure 10. SEM images of flexural mode fractured surface of (a) epoxy and (b) epoxy/graphene
nanocomposites.
16 A. MANI ET AL.

(a) (b)

Derivative Weight (%/oC)


100 E
E0.05GNP
80 E0.1GNP
1.4
Weight (%)

E0.2GNP

60

0.7
40 E
E0.05GNP
E0.1GNP
20
E0.2GNP
0.0

250 300 350 400 450 500 550 270 360 450
Temperature (oC) Temperature (oC)

Figure 11. (a) TG curves and (b) DTG curves of epoxy and epoxy/graphene nanocomposites.

The pure epoxy shows the maximum degradation occurs in the range of temperature as
310°C to 510°C. It is clearly seen from the TG curves that the TG curve shift to higher
temperature with addition of graphene in an epoxy matrix. The TG curve shows the initial
degradation temperature of an epoxy matrix of epoxy/graphene nanocomposites is higher
as compared to pure epoxy. This observation depicts an increase in thermal stability of
epoxy with addition of graphene. At 0.05 wt.% of graphene in an epoxy matrix, the thermal
stability is maximum. It is due to the finer dispersion of graphene in the epoxy matrix as
established with theoretical and experimentally observed flexural modulus at 0.05 wt.% of
graphene in an epoxy matrix. The degradation rate of epoxy-containing graphene is higher.
It is due to the formation of a char layer due to the presence of graphene in epoxy. The DTG
curves show the maximum degradation temperature of epoxy/graphene nanocomposites is
lower as compared to pure epoxy, confirming the degradation rate of epoxy is higher in
presence of epoxy.

3.3. Characterization of epoxy/graphene multiscale nanocomposites


In this work, MWNT is used to coat the glass fabric. The MWNT used in this work is
commercially available and of industrial grade. The obtained MWNT is characterized using
TEM. Figure 12(a) shows the TEM image of MWNT. From the TEM image of Figure 12(a),
the calculated value of average diameter of MWNT was ~10 nm. The TEM image of
MWNT shows the entangled structure and aggregation of MWNT. These entangle struc-
ture of MWNT needs to be disentangled. The disentanglement of MWNT is crucial in view
of the exploitation of the properties of MWNT. In literature, the disentanglement of
MWNT can be achieved with the helped of surfactant’s [92]. Many studies were reported
about the use of surfactant to disentangle and deagglomeration of the MWNT in aqueous
solution [93]. We are using the aqueous solution as a solvent in view of easy availability and
avoid the use of a toxic solvent. In addition, the surfactant uses was PVP. PVP was selected,
as the oxirane group of epoxy matrix interacts with the amido group of PVP have been
established earlier [3,4]. The function of PVP is multipurpose; it is not only helping in
COMPOSITE INTERFACES 17

(a) (b)

Figure 12. TEM images of (a) MWNT and (b) PVP modified MWNT.

deagglomerating MWNT but also improves the interfacial interaction between the epoxy
and MWNT. In this wor, the desired concentration of PVP in aqueous solution was added
and mixed on a magnetic in such a way that the PVP concentration was 1 mg/ml. Then,
MWNT was added in the solution of the PVP-aqueous solution and sonicated for one hour.
A drop of this solution was taken and pour over the copper grid to taking TEM images of
PVP modified MWNT. The TEM image of PVP modified MWNT is shown in Figure 12(b).
The TEM image of Figure 12(b) shows the individual MWNT. It also shows the low
molecular weight PVP film.
The PVP modified MWNT were coated over the glass fabric. The schematic of the
coating of PVP modified MWNT over glass fabric is shown in Figure 13(a). Initially,
MWNT is mixed with aqueous PVP solution using probe sonication. Then, the glass
fabric was dipped into the solution of PVP modified MWNT. The glass fabric was
dipped in the solution of PVP modified MWNT for 30 minutes followed by drying the
coated PVP-MWNT in a hot air oven. The SEM image of PVP modified MWNT
coated over glass fabric is shown in Figure 13(b). It shows the bidirectional glass
fabric. The average glass fiber size is 12 μm. The magnified inserted SEM image shows
the PVP modified MWNT deposited over the glass fabric surface. The glass fabric
appears rough due to the coating of PVP modified MWNT. The excessive surfactant
deposit also appears in the magnified inserted SEM image. It is important to note here
that the glass fabric was not heat treated to avoid the damage to the glass fabric. It was
reported earlier that the damage to glass fabric decreases the interfacial shear strength
of the composites [1].
The PVP coated MWNT glass fabric was used to prepare the epoxy multiscale
nanocomposites with and without the presence of graphene in epoxy matrix. The
schematic of processing protocol for the processing of glass fabric filled epoxy/
graphene multiscale nanocomposites with and without a coating of MWNT is
shown in Figure 14. The optimized concentration of graphene (0.1 wt.%) in the
epoxy matrix was chosen to prepare glass fabric filled epoxy/graphene multiscale
nanocomposites with and without a coating of MWNT. Three stage protocols, hand
18 A. MANI ET AL.

(a)

(b)

Figure 13. (a) Schematic of coating of PVP modified MWNT over glass fabric (b) SEM image of MWNT
decorated over glass fabric.

lay-up process followed by vacuum bagging and cold pressing was used (Figure 14).
In addition, post-curing was carried out in a hot air oven. There are 12 layers of
200 gsm bidirectional glass fabric in epoxy/graphene multiscale nanocomposites with
and without a coating of MWNT. The flexural test was performed over the glass
fabric filled epoxy/graphene multiscale nanocomposites with and without a coating
COMPOSITE INTERFACES 19

Figure 14. Schematic of processing of glass fabric filled epoxy/graphene nanocomposites.

of MWNT. Figure 15(a) shows the flexural stress-strain curves of the glass fabric
filled epoxy/graphene multiscale nanocomposites with and without a coating of
MWNT. From the flexural stress-strain curves, flexural modulus and flexural
strength were calculated. Figure 15(b) shows the flexural modulus and Figure 15(c)
shows the flexural strength of the glass fabric filled epoxy/graphene multiscale
nanocomposites with and without a coating of MWNT. The flexural modulus and
flexural strength of glass fabric filled epoxy composites are 21.9 GPa and 272 MPa
respectively. With incorporation of glass fabric coated with MWNT in epoxy multi-
scale nanocomposites shows a ~13% increase in flexural strength as compared to the
glass fabric filled epoxy composites. The flexural modulus of glass fabric coated with
MWNT filled epoxy multiscale nanocomposites is similar to the glass fabric filled
epoxy composites. Further, with an addition of 0.1 wt.% of graphene in epoxy and its
multiscale nanocomposites with glass fabric show enhancement in flexural modulus
and flexural strength respectively (30 GPa and 385 MPa; 36% and 41% respectively)
as compared to the glass fabric filled epoxy composites. In addition, 0.1 wt.% of
graphene in epoxy and its multiscale nanocomposites with MWNT coated glass
fabric shows further increase in flexural modulus and flexural strength respectively
as compared to 0.1 wt.% of graphene in epoxy and its multiscale nanocomposites
with glass fabric, and glass fabric filled epoxy composites. The increase in flexural
modulus and flexural strength for 0.1 wt.% of graphene in epoxy and its multiscale
nanocomposites with MWNT coated glass fabric are 71% and 116% respectively as
compared to glass fabric filled epoxy composites.
20 A. MANI ET AL.

(a) (b)
800 50

Flexural Modulus (GPa)


EGF EGF
Flexural Stress (MPa)

EGF-m-MWNT EGF-m-MWNT
EGFGNP 40 EGFGNP
600
EGF-m-MWNTGNP EGF-m-MWNTGNP
30
400
20

200
10

0 0
0.0000 0.0075 0.0150 0.0225 0.0300
Flexural strain

(c)
Flexural Strength (MPa)

600 EGF
EGF-m-MWNT
EGFGNP
450 EGF-m-MWNTGNP

300

150

Figure 15. (a) Flexural stress-strain curves, (b) Flexural modulus and (c) Flexural strength of glass fabric
filled with and without coating of MWNT epoxy and epoxy/graphene multiscale nanocomposites.

To probe the reason for an increase in flexural modulus and flexural strength of
glass fabric filled epoxy/graphene multiscale nanocomposites with and without
a coating of MWNT, DMA analysis was carried out. Figure 16 shows the storage
modulus and loss tangent (tan δ) as a function of the temperature of glass fabric
filled epoxy/graphene multiscale nanocomposites with and without a coating of
MWNT. The storage modulus of glass fabric filled epoxy/0.1 wt.% graphene multi-
scale nanocomposites is significantly higher as compared to glass fabric filled epoxy
composites. Similarly, the glass fabric coated MWNT filled epoxy multiscale nano-
composites, and the glass fabric coated MWNT filled epoxy/0.1 wt.% graphene
multiscale nanocomposites shows the significantly high storage modulus as com-
pared to glass fabric filled epoxy composites. The highest storage modulus is for glass
fabric coated MWNT filled epoxy/0.1 wt.% graphene multiscale nanocomposites.
And, the tanδ value for glass fabric filled epoxy/0.1 wt.% graphene multiscale
nanocomposites is slightly lower as compared to glass fabric filled epoxy composites.
For the glass fabric coated MWNT filled epoxy multiscale nanocomposites, the tanδ
value is slightly higher as compared to glass fabric filled epoxy composites. Further,
the glass fabric coated MWNT filled epoxy/0.1 wt.% graphene multiscale nanocom-
posites shows the slightly lower tanδ values as compared to glass fabric filled epoxy
composites. It is to be noted that the DMA studies were carried out in a torsional
COMPOSITE INTERFACES 21

(a) (b)
0.6 EGF
EGF
Storage Modulus (MPa)

4000 EGF-m-MWNT
EGF-m-MWNT
EGFGNP EGFGNP
EGF-m-MWNTGNP EGF-m-MWNTGNP
3000 0.4

tan
2000
0.2
1000

0 0.0
50 75 100 125 150 175 200 50 100 150
Temperature (oC) Temperature (oC)

Figure 16. (a) Storage Modulus (b) tanδ of glass fabric filled with and without coating of MWNT
epoxy and epoxy/graphene multiscale nanocomposites.

mode so that the in-plane stresses act on the layers of glass fabric and graphene is
maximum. The lower tanδ value for glass fabric coated MWNT filled epoxy/0.1 wt.%
graphene multiscale nanocomposites as compared to glass fabric epoxy composites
depicts lower damping depicting slide at the interface.
The reason for enhancement in flexural modulus and flexural strength of glass
fabric filled epoxy/graphene multiscale nanocomposites with a coating of MWNT are
discussed based on the schematic shown in Figure 17. From the flexural stress-strain
curves shown in Figure 15(a), the toughness of the glass fabric filled epoxy/graphene
multiscale nanocomposites with and without a coating of MWNT are more as
compared to the glass fabric filled epoxy composites. To probe the reason, let us
discuss the available literature. The composite materials fail due to the creation of
cracks at the interface of polymer and matrix, and further crack length increases
followed by failure of samples. It is well known that the tougher material has an
efficient mechanism of internal energy absorption, which restricts the crack propaga-
tion (Figure 17(b)) [94]. A crack cannot propagate, if the energy of the system is not
decreased (Figure 17(b)) [94]. In case of the glass fabric filled epoxy composites, the
surface of the glass fabric is smooth; as it coated with a polymeric coating during
glass fiber processing. But, the surface is rough, when glass fabric is coated with PVP
modified MWNT (Figures 13(b), 17(c)). The rough surface results in mechanical
keying (Figure 17(d)), which offers interfacial resistance to deformation. But, the Tg
value obtained from DMA studies for MWNT coated glass fabric filled epoxy
multiscale nanocomposites is lower as compared to glass fabric filled epoxy compo-
sites. This observation depicts the more mobility of epoxy chains of MWNT coated
glass fabric filled epoxy multiscale nanocomposites. The mobile epoxy chains are
contributing towards the energy absorption, and thus the toughness of the MWNT
coated glass fabric filled epoxy multiscale nanocomposites increases. An et al. [33]
observed that interfacial sliding occurs by coating MWNT over GF and reinforced in
an epoxy matrix. Thus, it confirms from the literature data that there exists the
mobility of epoxy chain at the interface of MWNT coated over GF and epoxy. The
22 A. MANI ET AL.

Figure 17. Schematic of (a) Epoxy/Glass fabric layers (b) Energy absorption and crack length
relationship (c) MWNT coated glass fabric epoxy multiscale nanocomposites (d) Mechanical keying
(e) Mechanical interlocking.

mechanical keying is associated with the increase in flexural strength of MWNT


coated glass fabric filled epoxy multiscale nanocomposites.
In case of glass fabric filled epoxy/graphene multiscale nanocomposites, the flexural
toughness and storage modulus values are higher than that of MWNT coated glass
fabric filled epoxy multiscale nanocomposites and glass fabric filled epoxy composites.
The Tg value obtained from DMA studies for is slightly lower as compared to glass
fabric filled epoxy composites. But, the graphene is randomly dispersed and wrinkled
structure associated with mechanical interlocking during applied loading (Figure 17(e)).
Due to interlocking of the graphene nanosheets, the flexural modulus, flexural strength,
and storage modulus of glass fabric filled epoxy/graphene multiscale nanocomposites
are higher. When the glass fabric is coated with MWNT and its epoxy/graphene
multiscale nanocomposites, the flexural modulus, flexural strength, and storage mod-
ulus obtained are highest as compared to other samples. The Tg value of glass fabric is
coated with MWNT and its epoxy/graphene multiscale nanocomposites are lower as
compared to glass fabric filled epoxy composites. This observation depicts that the
epoxy chains in contact with PVP coated MWNT are mobile. Thus, due to mechanical
interlocking, the mobility of epoxy chains and keying, the flexural modulus, flexural
strength, and storage modulus obtained are highest.

4. Conclusion
In this work, graphene oxide was synthesized and characterized using the improved method.
The graphene oxide was chemically reduced using sodium hydrosulfite. TEM studies reveal
that the obtained graphene nanosheets are multi-layered and have wrinkled structure, as
clear from. FTIR studies confirm the graphene obtained has functional groups attached.
These graphene nanosheets were incorporated in the epoxy matrix using ball milling. There
exist an interaction between C-N group of epoxy and C=O group of graphene, as confirmed
COMPOSITE INTERFACES 23

from FTIR studies. The mixing of graphene and epoxy using ball milling leads to the
breakage of graphene sheets. However, the incorporation of graphene at a lower concentra-
tion of 0.1 wt.% in the epoxy matrix show a significant increase in the flexural properties as
compared to the pure epoxy matrix. But, there is a mismatch in the theoretically predicted
and experimentally obtained flexural modulus; it is due to the presence of multilayer
graphene nanosheets in the epoxy matrix.
The optimized concentration of 0.1 wt.% of graphene in the epoxy matrix is used to
prepare glass fabric reinforced multiscale nanocomposites. In addition, glass fabric was
coated with MWNT; and the MWNT coated glass fabric was used to prepare 0.1 wt.% of
graphene/epoxy multiscale nanocomposites. With addition of 0.1 wt.% of graphene in
epoxy and its multiscale nanocomposites with glass fabric show enhancement in flexural
modulus and flexural strength respectively as compared to the glass fabric filled epoxy
composites. The reason for the increase in flexural properties is mechanical interlocking.
Further, incorporation of the MWNT coated glass fiber in epoxy matrix shows an increase
in flexural strength as compared to glass fiber filled epoxy composites. It is due to
mechanical keying associated with the rough surface obtained with the coating of
MWNT over glass fabric. At last, the addition of 0.1 wt.% of graphene in epoxy and its
multiscale nanocomposites with MWNT coated glass fabric shows the highest enhance-
ment in the flexural modulus and flexural strength. The mechanical keying and mechanical
interlocking are associated with the increase in flexural properties. The toughness of the
MWNT coated glass fabric and graphene filled epoxy multiscale nanocomposites increases
due to the mobile epoxy chains contributing towards the energy absorption. This observa-
tion depicts that the epoxy chains in contact with PVP coated MWNT are mobile. Thus,
due to mechanical interlocking, the mobility of epoxy chains and mechanical keying, the
flexural modulus, flexural strength, and storage modulus obtained are highest for MWNT
coated glass fabric and graphene filled epoxy multiscale nanocomposites.

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

ORCID
Pankaj Tambe http://orcid.org/0000-0003-3536-0551

References
[1] Lubineau G, Rahaman A. A review of strategies for improving the degradation properties
of laminated continuous-fiber/epoxy composites with carbon-based nanoreinforcements.
Carbon. 2012;50:2377–2395.
[2] Rahaman A, Kar K, Chaudhary D. Synthesis of carbon nanotubes on nickel-silica catalyst
coated E-glass fiber/fabric and its nanocomposites. Int J Plast Technol. 2010;14:65–79.
[3] Saha M, Tambe P, Pal S, et al. Effect of non-ionic surfactant assisted modification of
hexagonal boron nitride nanoplatelets on the mechanical and thermal properties of epoxy
nanocomposites. Compos Interfaces. 2015;22:611–627.
24 A. MANI ET AL.

[4] Saha M, Tambe P, Pal S. Thermodynamic approach to enhance the dispersion of graphene
in epoxy matrix and its effect on mechanical and thermal properties of epoxy
nanocomposites. Compos Interfaces. 2016;23:255–272.
[5] Kulkarni H, Pankaj T, Girish J. High concentration exfoliation of graphene in ethyl alcohol
using block copolymer surfactant and its influence on properties of epoxy nanocomposites.
Fuller Nanotub Carb Nanostruct. 2017;25:241–249.
[6] Swaminathan G, Palanisamy C, Chidambaram G, et al. Enhancing the interfacial strength
of glass/epoxy composites using ZnO nanowires. Compos Interfaces. 2017;25:151–168.
[7] Shen S, Duan M, Yang P, et al. Tunable mechanical properties of MWCNT–glass fiber
fabric reinforced epoxy composites by controlling MWCNTs dispersing conditions.
Compos Interfaces. 2018;25:901–918.
[8] Zaheer U, Zulfiqar U, Khurram A, et al. Improving the performance of conventional glass
fiber epoxy matrix composites by incorporating nanodiamonds. Compos Interfaces.
2018;25:1005–1018.
[9] Gojny FH, Wichmann MHG, Fiedler B, et al. Influence of nano-modification on the
mechanical and electrical properties of conventional fibre-reinforced composites.
Compos Part A. 2005;36:1525–1535.
[10] Fan Z, Santare MH, Advani SG. Interlaminar shear strength of glass fiber reinforced epoxy
composites enhanced with multi-walled carbon nanotubes. Compos Part A.
2008;39:540–554.
[11] Manfredi LB, De Santis H, Vazquez A. Influence of the addition of montmorillonite to the
matrix of unidirectional glass fibre/epoxy composites on their mechanical and water
absorption properties. Compos Part A. 2008;39:1726–1731.
[12] Park JM, Wang ZJ, Jang JH, et al. Interfacial and hydrophobic evaluation of glass fiber/
CNT–epoxy nanocomposites using electro-micromechanical technique and wettability
test. Compos Part A. 2009;40:1722–1731.
[13] Warrier A, Godara A, Rochez O, et al. The effect of adding carbon nanotubes to glass/
epoxy composites in the fibre sizing and/or the matrix. Compos Part A.
2010;41:532–538.
[14] Garg M, Sharma S, Mehta R. Pristine and amino functionalized carbon nanotubes rein-
forced glassfiber epoxy composites. Compos Part A. 2015;76:92–101.
[15] Wichmann MHG, Sumfleth J, Gojny FH, et al. Glass-fibre-reinforced composites with
enhanced mechanical and electrical properties – benefits and limitations of a nanoparticle
modified matrix. Eng Fract Mech. 2006;73:2346–2359.
[16] Rahman MM, Zainuddin S, Hosur MV, et al. Improvements in mechanical and
thermo-mechanical properties of e-glass/epoxy composites using amino functionalized
MWCNTs. Comp Struct. 2012;94:2397–2406.
[17] Lin LY, Lee JH, Hong CE, et al. Preparation and characterization of layered silicate/glass
fiber/epoxy hybrid nanocomposites via vacuum-assisted resin transfer molding (VARTM).
Comp Sci Technol. 2006;66:2116–2125.
[18] Bozkurt E, Kaya E, Tanoglu M. Mechanical and thermal behavior of non-crimp glass fiber
reinforced layered clay/epoxy nanocomposites. Comp Sci Technol. 2007;67:3394–3403.
[19] Wicks SS, de Villoria RG, Wardle BL. Interlaminar and intralaminar reinforcement of
composite laminates with aligned carbon nanotubes. Comp Sci Technol. 2010;70:20–28.
[20] Withers GJ, Yu Y, Khabashesku VN, et al. Improved mechanical properties of an epoxy
glass–fiber composite reinforced with surface organomodified nanoclays. Compos Part B.
2015;72:175–182.
[21] Haque A, Shamsuzzoha M, Hussain F, et al. S2-gass/epoxy polymer nanocomposites:
manufacturing, structures, thermal and mechanical properties. J Comp Mat.
2003;37:1821–1837.
[22] Petersen HN, Minty RF, Thomason JL, et al. The amine: epoxideratio at the interface of
a glass fibre/epoxy matrix system and its influence on the interfacial shear strength.
Compos Interfaces. 2018;1–14. DOI:10.1080/19276440.2018.1511107
COMPOSITE INTERFACES 25

[23] Sadeghi M, Pol MH. Investigation of behaviors of glass/epoxy laminate composites


reinforced with carbon nanotubes under quasi-static punch shear loading. Sand Struct
Mat. 2017;1–22. DOI:10.1177/1099636217719223
[24] Wan Dalina WAD, Mariatti M, Tan SH. Effect of black ink loading on the properties of
multi-walled carbon nanotubes/glass fibre/epoxy laminated hybrid composites. Polym
Bull. 2018;75:3357–3375.
[25] Zeiler R, Kuttner C, Khalid U, et al. The role of multi-walled carbon nanotubes in epoxy
nanocomposites and resin transfer molded glass fiber hybrid composites: dispersion, local
distribution, thermal, and fracture/mechanical properties. Polym Compos. 2015;38:1–15.
[26] Kamar NT, Hossain MM, Khomenko A, et al. Interlaminar reinforcement of glass fiber/
epoxy composites with graphene nanoplatelets. Compos Part A. 2015;70:82–92.
[27] Rathore DK, Prusty RK, Kumar DS, et al. Mechanical performance of CNT-filled glass
fiber/epoxy composite in in-situ elevated temperature environments emphasizing the role
of CNT content. Compos Part A. 2016;84:364–376.
[28] Rahaman A, Imran M. Epoxy-carbon nanotubes as matrix in glass fiber reinforced
laminated composites. Fuller Nanotub Carb Nanostruct. 2017;1–18. DOI:10.1080/
1536383X.2017.1330264
[29] Chen J, Zhao D, Jin X, et al. Modifying glass fibers with graphene oxide: towards
high-performance polymer composites. Comp Sci Technol. 2014;97:41–45.
[30] Mahmood H, Tripathi M, Pugno N, et al. Enhancement of interfacial adhesion in glass
fiber/epoxy composites by electrophoretic deposition of graphene oxide on glass fibers.
Comp Sci Technol. 2016;126:149–157.
[31] Hao B, Ma Q, Yang S, et al. Comparative study on monitoring structural damage in
fiber-reinforced polymers using glass fibers with carbon nanotubes and graphene coating.
Comp Sci Technol. 2016;129:38–45.
[32] Zhang J, Zhuang R, Liu J, et al. Functional interphases with multi-walled carbon nanotubes
in glass fibre/epoxy composites. Carbon. 2010;48:2273–2281.
[33] An Q, Tamrakar S, Gillespie JW Jr., et al. Tailored glass fiber interphases via electro-
phoretic deposition of carbon nanotubes: fiber and interphase characterization. Comp Sci
Technol. 2018;166:131–139.
[34] Eskizeybek V, Avcı A, Gulce A. Preparation and mechanical properties of carbon nano-
tube grafted glass fabric/epoxy multi-scale composites. Adv Comp Mat. 2015;1–12.
DOI:10.1080/09243046.2015.1052188
[35] Zeng S, Duan P, Shen M, et al. Interface enhancement of glass fiber fabric/epoxy compo-
sites by modifying fibers with functionalized MWCNTs. Compos Interfaces. 2018;1–18.
DOI:10.1080/09276440.2018.1499354
[36] Garg M, Sharma S, Mehta R. Role of curing conditions and silanization of glass fibers on
carbon nanotubes (CNTs) reinforced glass fiber epoxy composites. Compos Interfaces.
2017;24:233–253.
[37] Moosburger-Will J, Jäger J, Strauch J, et al. Interphase formation and fiber matrix adhesion
in carbon fiber reinforced epoxy resin: influence of carbon fiber surface chemistry.
Compos Interfaces. 2017;24:691–710.
[38] Rodríguez-González JA, Rubio-González C, Meneses-Nochebuena CA, et al. Enhanced
interlaminar fracture toughness of unidirectional carbon fiber/epoxy composites modified
with sprayed multi-walled carbon nanotubes. Compos Interfaces. 2017;24:883–896.
[39] Lee C, Wei X, Kysar JW, et al. Measurement of the elastic properties and intrinsic strength
of monolayer graphene. Science. 2008;321:385–388.
[40] Thayumanavan N, Tambe P, Joshi G. Effect of surfactant and sodium alginate modifica-
tion of graphene on mechanical and thermal properties of polyvinyl alcohol (PVA)
nanocomposites. Cell Chem Technol. 2015;49:69–80.
[41] Thayumanavan N, Tambe P, Joshi G. Effect of sodium alginate modification of graphene
(by ‘anion-π’type of interaction) on the mechanical and thermal properties of polyvinyl
alcohol (PVA) nanocomposites. Compos Interfaces. 2014;21:487–506.
26 A. MANI ET AL.

[42] Lakshmi NV, Tambe P, Sahu NK. Giant permittivity of three phase polymer nanocompo-
sites obtained by modifying hybrid nanofillers with polyvinylpyrrolidone. Compos
Interfaces. 2018;25:47–67.
[43] Zhang X, Liang C, Zimmermann S, et al. Elastic properties of graphene grown by chemical
vapor deposition. Proceedings of the 17th Symposium on Design Test Integration and
Packaging of MEMS/MOEMS (DTIP); 2015 April 27–30;Montpellier, France.
[44] Marcano DC, Kosynkin DV, Berlin JM, et al. Improved synthesis of graphene oxide. ACS
Nano. 2010;4:4806–4814.
[45] Wu H, Huanmg X, Qian L. Preparation, mechanism and property of metacomposites with
carbon materials as fillers. Eng Sci. 2018;2:17–25.
[46] Kirubasankar B, Murugadoss V, Lin J, et al. In situ grown nickel selenide on graphene
nanohybrid electrodes for high energy density asymmetric supercapacitors. Nanoscale.
2018;10:20414–20425.
[47] Cheng C, Fan R, Ren Y, et al. Radio frequency negative permittivity in random carbon
nanotubes/alumina nanocomposites. Nanoscale. 2017;9:5779–5787.
[48] Su T, Shao Q, Qin Z, et al. Role of interfaces in two-dimensional photocatalyst for water
splitting. ACS Catal. 2018;8:2253–2276.
[49] Du H, Zhao CX, Lin J, et al. Carbon nanomaterials in direct liquid fuel cells. Chem Rec.
2018;18:1365–1372.
[50] Cheng C, Fan R, Wang Z, et al. Tunable and weakly negative permittivity in carbon/silicon
nitride composites with different carbonizing temperatures. Carbon. 2017;125:103–112.
[51] Lin J, Huang Y, Wang S, et al. Microwave-assisted rapid exfoliation of graphite into
graphene by using ammonium bicarbonate as the intercalation agent. Carbon.
2018;140:696–733.
[52] Song B, Wang T, Sun H, et al. Two-step hydrothermally synthesized carbon nanodots/
WO3 photocatalysts with enhanced photocatalytic performance. Dalton Trans.
2017;46:15769–15777.
[53] Lin C, Hu L, Cheng C, et al. Nano-TiNb2O7/carbon nanotubes composite anode for
enhanced lithium-ion storage. Electrochim Acta. 2018;260:65–72.
[54] Li Y, Jing T, Xu G, et al. 3-D magnetic graphene oxide-magnetite poly(vinyl alcohol)
nanocomposite substrates for immobilizing enzyme. Polymer. 2018;149:13–22.
[55] Zhao J, Wu L, Zhan C, et al. Overview of polymer nanocomposites: computer simulation
understanding of physical properties. Polymer. 2017;133:272–287.
[56] Gong K, Hu Q, Yao L, et al. Ultrasonic pretreated sludge derived stable magnetic active
carbon for Cr(VI) removal from wastewater. ACS Sustain Chem Eng. 2018;6:7283–7291.
[57] Huang J, Cao Y, Shao Q, et al. Magnetic nanocarbon adsorbents with enhanced hexavalent
chromium removal: morphology dependence of fibrillar vs particulate structures. Ind Eng
Chem Res. 2017;56:10689–10701.
[58] Li Z, Wang B, Qin X, et al. Superhydrophobic/superoleophilic polycarbonate/carbon
nanotubes porous monolith for selective oil adsorption from water. ACS Sustain Chem
Eng. 2018;6:13747–13755.
[59] Wu N, Liu C, Xu D, et al. Enhanced electromagnetic wave absorption of three-dimensional
porous Fe3O4/C composite flowers. ACS Sustain Chem Eng. 2018;6:12471–12480.
[60] Yang Z, Hao X, Chen S, et al. Long-term antibacterial stable reduced graphene oxide
nanocomposites loaded with cuprous oxide nanoparticles. J Colloid Interf Sci.
2019;533:13–23.
[61] Lu Y, Biswas MC, Guo Z, et al. Recent developments in bio-monitoring via advanced
polymer nanocomposite-based wearable strain sensors. Biosens Bioelectron.
2019;123:167–177.
[62] Dong M, Li Q, Liu H, et al. Thermoplastic polyurethane-carbon black nanocomposite
coating: fabrication and solid particle erosion resistance. Polymer. 2018;158:381–390.
[63] Qian Y, Yuan Y, Wang H, et al. Highly efficient uranium adsorption by salicylaldoxime/
polydopamine graphene oxide nanocomposites. J Mater Chem A. 2018;6:24676–24685.
COMPOSITE INTERFACES 27

[64] Liu H, Li Q, Zhang S, et al. Electrically conductive polymer composites for smart flexible
strain sensors: a critical review. J Mater Chem C. 2018;6:12121–12141.
[65] Guo J, Song H, Liu H, et al. Polypyrrole-interface-functionalized nano-magnetite epoxy
nanocomposites as electromagnetic ave absorbers with enhanced flame retardancy. J Mater
Chem C. 2017;5:5334–5344.
[66] Wu Z, Cui H, Chen L, et al. Interfacially reinforced unsaturated polyester carbon fiber
composites with a vinyl ester-carbon nanotubes sizing agent. Compos Sci Technol.
2018;164:195–203.
[67] Song B, Wang T, Sun H, et al. Graphitic carbon nitride (g-C3N4) interfacially strength-
ened carbon fiber epoxy composites. Compos Sci Technol. 2018;167:515–521.
[68] Deng W, Kang T, Liu H, et al. Guo ZPotassium hydroxide activated and nitrogen doped
graphene with enhanced supercapacitive behavior. Sci Adv Mater. 2018;10:937–949.
[69] Luo Q, Ma H, Huo Q, et al. All-carbon-electrode-based endurable flexible perovskite solar
cells. Adv Funct Mater. 2018;28:1706–1717.
[70] Guo J, Song H, Liu H, et al. Polypyrrole-interface-functionalized nano-magnetite epoxy
nanocomposites as electromagnetic ave absorbers with enhanced flame retardancy. J Mater
Chem C. 2017;5:5334–5344.
[71] Gu H, Zhang H, Ma C, et al. Trace electrosprayed nanopolystyrene facilitated dispersion of
multiwalled carbon nanotubes: simultaneously strengthening and toughening epoxy.
Carbon. 2019;142:131–140.
[72] Zhang Y, Zhao M, Zhang J, et al. Excellent corrosion protection performance of epoxy
composite coatings filled with silane functionalized silicon nitride. J Polym Res.
2018;25:130–142.
[73] Wang C, Zhao M, Li J, et al. Silver nanoparticles/graphene oxide decorated carbon fiber
synergistic reinforcement in epoxy-based composites. Polymer. 2017;131:263–271.
[74] Wang L, Qiu H, Liang C, et al. Electromagnetic interference shielding MWCNT-Fe3O4
@Ag/epoxy nanocomposites with satisfactory thermal conductivity and high thermal
stability. Carbon. 2019;141:506–514.
[75] He Y, Yang S, Liu H, et al. Reinforced carbon fiber laminates with oriented carbon
nanotube epoxy nanocomposites: magnetic field assisted alignment and cryogenic tem-
perature mechanical properties. J Colloid Interf Sci. 2018;517:40–51.
[76] Wu Z, Gao S, Chen L, et al. Electrically insulated epoxy nanocomposites reinforced with
synergistic core–shell SiO2@ MWCNTs and montmorillonite Bifillers, J. Colloid Interf Sci.
2018;517:40–51.
[77] Hu Z, Zhang D, Lu F, et al. Multistimuli-responsive intrinsic self-healing epoxy resin
constructed by host–guest interactions. Macromolecules. 2018;51:5294–5303.
[78] Song B, Wang T, Wang L, et al. Interfacially reinforced carbon fiber/epoxy composite
laminates via in-situ synthesized graphitic carbon nitride (g-C3N4). Compos B.
2019;158:259–268.
[79] Zhao M, Meng L, Ma L, et al. Layer-by-layer grafting CNTs onto carbon fibers surface for
enhancing the interfacial properties of epoxy resin composites. Compos Sci Technol.
2018;154:28–36.
[80] Shen X, Lin X, Yousefi N, et al. Wrinkling in graphene sheets and graphene oxide papers:
effect on young’s modulus. Carbon. 2014;66:84–92.
[81] Lakshmi NV, Pankaj T. EMI shielding effectiveness of graphene decorated with graphene
quantum dots and silver nanoparticles reinforced PVDF nanocomposites. Compos
Interfaces. 2017;24:861–882.
[82] Guo HL, Wang XF, Qian QY, et al. A green approach to the synthesis of graphene
nanosheets. ACS Nano. 2009;9:2653–2659.
[83] Wanga G, Shena X, Wanga B, et al. Synthesis and characterisation of hydrophilic and
organophilic graphene nanosheets. Carbon. 2009;47:1359–1364.
[84] Wang G, Wang B, Park J, et al. Synthesis of enhanced hydrophilic and hydrophobic
graphene oxide nanosheets by a solvothermal method. Carbon. 2009;47:68–72.
28 A. MANI ET AL.

[85] Oh J, Lee JH, Koo JC, et al. Graphene oxide porous paper from amine-functionalized poly
(glycidyl methacrylate)/graphene oxide core-shell microspheres. J Mater Chem.
2010;20:9200–9204.
[86] León V, Quintana M, Herrero MA, et al. Few-layer graphenes from ball-milling of graphite
with melamine. Chem Comm. 2011;47:10936–10938.
[87] Nikolic G, Zlatkovic S, Cakic M, et al. Fast fourier transform IR characterization of epoxy
GY systems crosslinked with aliphatic and cycloaliphatic EH polyamine adducts. Sensors.
2010;10:684–696.
[88] Kulkarni HB, Pankaj T, Girish MJ. Influence of covalent and non-covalent modification of
graphene on the mechanical, thermal and electrical properties of epoxy/graphene nano-
composites: a review. Compos Interfaces. 2018;25:381–414.
[89] Natarajan B, Li Y, Deng H, et al. Effect of interfacial energetics on dispersion and glass
transition temperature in polymer nanocomposites. Macromolecules. 2013;46:2833−2841.
[90] Rittigstein P, Priestley RD, Broadbelt LJ, et al. Model polymer nanocomposites provide an
understanding of confinement effects in real nanocomposites. Nat Mater. 2007;6:278−82.
[91] Yasmin A, Isaac MD. Mechanical and thermal properties of graphite platelet/epoxy
composites. Polymer. 2004;45:8211–8219.
[92] Wang Q, Han Y, Wang Y, et al. Effect of surfactant structure on the stability of carbon
nanotubes in aqueous solution. J Phys Chem B. 2008;112:7227–7233.
[93] Clark MD, Subramanian S, Krishnamoorti R. Understanding surfactant aided aqueous
dispersion of multi-walled carbon nanotubes. J Colloid Interf Sci. 2011;354:144–151.
[94] Hull D, Clyne TW. An introduction to composite materials. 2nd ed. Cambridge, UK:
Cambridge Solid State Science Series; 1996.

You might also like