You are on page 1of 31

Accepted Manuscript

Pristine and amino functionalized carbon nanotubes reinforced glass fiber epoxy
composites

Mohit Garg, Shruti Sharma, Rajeev Mehta

PII: S1359-835X(15)00166-9
DOI: http://dx.doi.org/10.1016/j.compositesa.2015.05.012
Reference: JCOMA 3933

To appear in: Composites: Part A

Received Date: 11 December 2014


Revised Date: 13 May 2015
Accepted Date: 15 May 2015

Please cite this article as: Garg, M., Sharma, S., Mehta, R., Pristine and amino functionalized carbon nanotubes
reinforced glass fiber epoxy composites, Composites: Part A (2015), doi: http://dx.doi.org/10.1016/j.compositesa.
2015.05.012

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Pristine and amino functionalized carbon nanotubes

reinforced glass fiber epoxy composites

Mohit Garg1a*, Shruti Sharmab and Rajeev Mehtac


a
Department of Mechanical Engineering, Thapar University, Patiala, India, 147004 (mohitgarg@thapar.edu)
b
Department of Civil Engineering, Thapar University, Patiala, India, 147004 (shruti.sharma@thapar.edu)
c
Department of Chemical Engineering, Thapar University, Patiala, India, 147004 (rmehta@thapar.edu)

Abstract. Multi-phase composites have been studied by incorporating carbon nanotubes (CNTs) as a

secondary reinforcement in an epoxy matrix which was then reinforced with glass fiber mat. Different types

of CNTs e.g. amino functionalized carbon nanotubes (ACNT) and pristine carbon nanotubes (PCNT) were

homogeneously dispersed in the epoxy matrix and two-ply laminates were fabricated using vacuum-assisted

resin infusion molding technique. The issues related to CNT dispersion and interfacial bonding and its affect

on the mechanical properties have been studied. An important finding of this study is that PCNT scores over

ACNT in composites prepared under certain conditions. This is a very significant finding since PCNT is

available at a much lower cost than ACNT.

Keywords: Polymer-matrix composites (A); Mechanical properties (B); CNT (C); Functionalization (D).

1
Corresponding author: Mr. Mohit Garg

E-mail: mohitgarg@thapar.edu

Department of Mechanical Engineering, Thapar University, Patiala, INDIA

Phone number: +919646527547

1
1 INTRODUCTION
Glass and carbon fibers have been used as reinforced materials in preparing light-weight composites for

structural applications such as aerospace, marine and automotive. These fibers are embedded in a relatively

compatible polymeric resin, such as epoxy and unsaturated polyester [1, 2] for preparing fiber reinforced polymer

composites (FRP). The FRP composites exhibit high strength to weight ratio, excellent corrosion and fatigue

resistance in relation to conventional metals and alloys used. The commonly used reinforcement in FRP composites

is glass fiber due to its easier availability, processing, and low cost. The most common form in which FRP

composites are used in industry is a laminate (made by stacking a number of thin layers of fibers and matrix

integrated into a desired thickness). These laminates involve thermoset matrix such as epoxy for preparing FRP

composites. The epoxy polymer is preferred due to its superior thermal, mechanical and electrical properties,

dimensional stability and chemical resistance [3]. In-order to improve the mechanical performance of the

conventional FRP composites much effort has been done in the past by modifying the polymer matrix with

nanofillers such as graphene, carbon nanotubes (CNTs) and nanoclay [4-10].

The CNT which are long and thin cylinders of carbon were discovered by S. Iijima in 1991[11]. After this

discovery there has been a phenomenal growth in fabricating functional polymer composites using CNTs as

nanofillers. CNT has excellent intrinsic properties due to its diameter in the nano range, low density (1.3 – 2 g/cm3)

and high aspect ratio. The physical properties of CNT include: high electrical (106 – 107 S/m) and thermal

conductivities (> 3000 W/m.k), and exceptionally high mechanical properties such as high strength (10 – 60 GPa)

and stiffness (1 TPa), which make them an ideal candidate for preparing multi-functional polymer composites

[5,12-17]. For efficient utilization of CNT as reinforcement in the polymer matrix, homogeneous dispersion and

stable interfacial bonding between the nanotubes and polymer matrix needs to be achieved [18-20]. The high aspect

ratio of CNTs and strong forces of attraction between the carbon particles results in the formation of agglomerates

which consequently influence the homogenous dispersion and interfacial bonding and hence affects the material

properties. The main cause of formation of agglomerates is that the nanotubes tend to attract each other due to the

presence of weak van der Waals forces and Coulomb attractions even in an uncured polymer matrix [21, 22]. On

the other hand, the presence of strong interfacial adhesion between the matrix and CNT would allow the stress to be

transferred from the matrix to the nanotubes owing to which the mechanical properties would be enhanced. This is

because the interface acts as an intermittent bridge which allows the stress to be immediately transferred from the

2
weak matrix to the stronger nanotubes embedded in the matrix when subjected to different loading conditions.

Therefore, a design incorporating interface as a critical parameter needs to be considered to control the strength and

toughness of the resulting polymer composites [23, 24].

To improve the covalent bonding and uniform dispersion of nanotubes in polymer matrix (for example epoxy,

vinyl-ester etc.) different processing techniques such as calendaring [25-28], ultrasonication [25, 29] and high-shear

mixing [12, 30] have been used. However, no standard procedure or protocol has been established for achieving

homogeneous dispersion of CNT in the polymer matrix and good interfacial bonding between the CNTs and

polymer matrix. From literature it has been observed that, the simple and effective methods such as sonication and

shear mixing results in considerable improvement in the mechanical properties [12, 25, 29, 30].

Recently, epoxy nanocomposites based on pristine and amino-functionalized CNTs were explored [12, 31-36]

and it was reported that the improvement in the composites with pristine nanotubes (PCNT) is not significant due to

poor dispersion and wetting of CNTs with the resin. Also, the reinforcement effect of CNTs in stiff epoxy matrix is

hindered by weak interfacial bonding which results in pull-out of the nanotubes under loading conditions. However,

chemical functionalization of the nanotubes such as amino functionalization CNTs leads to better dispersion than

pristine CNTs in the polymer matrix. Chemical functionalization of CNTs aids in forming strong covalent and ionic

bonds with the epoxy matrix and enable a more effective transfer of stress between the epoxy matrix and CNTs

with an improvement in interfacial adhesion [30]. As a result, better dispersion and superior mechanical properties

are achieved. Generally, the chemical functionalization of CNTs introduces carboxylic groups (-COOH) on the

surface, which can further be chemically treated to attach functional groups such as chlorine, amine and silane on

the surface of CNT depending upon its applicability. Recently, it was reported [41- 44] that chemical

functionalization of CNTs results in better dispersion in the epoxy resin and improves mechanical and thermal

properties of the polymer composites. Out of the various functionalization schemes, amino functionalization of

CNTs has gained a lot of interest in recent years due to improved dispersion and interfacial adhesion with polymer

resins [14, 25, 29 – 31, 37 – 40, 45]. Few studies [29, 46] have also investigated the effect of reactivity of amino

groups with the epoxide groups during the curing process and the resulting strong interfacial bonding which

facilitated the stress-transfer mechanism. Interestingly, the microscopic images of the fractured surface of amino

functionalized nanotubes (ACNT) in an epoxy system showed pull-out of the outermost layer that was covalently

3
bonded to the matrix while the inner layers of ACNT acted as a bridge between the matrix and CNTs thus

confirming the existence of good interfacial adhesion with the epoxy resin.

The findings from the above studies highlighted the use of different processing techniques such as calendaring,

sonication, high-shear mixing and mechanical stirring to disperse CNTs in matrix. However, the experimental

results obtained are quite below the theoretical predictions either due to poor dispersion or in-efficient load transfer

from the matrix to CNTs. Therefore, it is essential to select the optimal processing techniques to homogenously

disperse CNTs at nanolevel and to achieve even better mechanical properties.

The nanotube modified matrix can be combined with traditional high-performance reinforcement (e.g. carbon or

glass fibers) for preparing multi-phase composites. The application of these multi-phase composites as structural

materials is found in aerospace industry, marine structures, armor, automobiles and sporting goods etc.. However,

only a few studies have incorporated different processing routes [47 - 49]. A study by Gojny et al. [47] used resin

transfer molding (RTM) to illustrate 20% increase in interlaminar shear strength (ILSS) due to the presence of

amine groups on the surface of CNTs. Zhihang et al. [48] reported a further increase in ILSS properties by 30%

when CNT were aligned perpendicular to the direction of fiber surface. The results indicated that an alignment of

CNT showed improvement in both ILSS and Mode-1 fracture toughness properties of the multi-phase composites.

Consequently, the aim for adding nanotubes to the conventional fiber composites is to minimize the matrix

limitations and prepare composites with improved mechanical performance. Additionally, the multi-phase

composites are usually fabricated by resin transfer molding (RTM), vacuum assisted resin transfer molding

(VARTM) or vacuum assisted resin infusion molding (VARIM) techniques. In the present study VARIM has been

used to fabricate multi-phase composites due to its applicability in medium to large-scale industries [50].

In addition to the parameters discussed above for enhancing the properties of multi-phase composites, effect of

different epoxy/amine ratio is also an important parameter for consideration. Recently, a study by Gude et al. [51]

brought an innovative concept that, changing the stoichiometric ratio of the resin in the matrix had a greater effect

on the properties of the resin as compared to modifying the matrix with the addition of CNTs. It was also shown

that the effect of CNT on the properties is negligible in stoichiometric and amine-rich resin systems. However, the

effect of addition of ACNT in epoxy resin was not explored completely, since there were several other parameters

that could also affect the final properties.

4
Thus, there are significant gaps in literature pertaining to effect of PCNT/ACNT dispersion and interfacial

bonding on the mechanical properties. The present study aims at investigating various factors related to processing

and fabrication of composites. The first one involves the use of low viscosity resin and hardener to assist better

dispersion of CNTs, applicable for marine structures. Secondly, a combination of optimal dispersion techniques i.e.

sonication and high-shear mixing have been used to achieve homogenous dispersion of CNTs in epoxy resin.

Furthermore, a comparative study has been carried out by incorporating both pristine and amino-functionalized

CNTs in the epoxy resin to study their effect on the composite properties. In the literature, generally either tensile or

flexural properties have been studied for CNT composites but the present study has focused on both of them along

with fracture-surface analysis. Moreover, a cost-effective manufacturing technique, VARIM, has been used for

preparing the laminate sheets with minimal porosity and uniform thickness. Taking a clue from some reported

studies [51, 52] the effect of CNTs in both stoichiometric and amine-rich resins has been investigated since the

presence of excess amine group in the composites would result in faster degradation of the polymer nanocomposites

when subjected to hygrothermal environment. This study will have a significant effect on enhancing the mechanical

properties of the CNT reinforced glass fiber polymer composites, while opening new horizons of preparing FRP

modified polymer composites applicable for marine environment.

2 EXPERIMENTAL
2.1 Materials

Advantex® E-glass woven roving fiber mat (GF), WR310 was purchased from Owens Corning Inc., (India).

Commercially available, multi-wall carbon nanotubes (MWCNTs) with carbon purity of 99% and outer diameter of

10-20 nm and length of 3-8 µm were purchased from Nanoshel (U.S.A.). Catalytic carbon vapor deposited

MWCNTs functionalized with amino groups (<0.5%) (Nanocyl, NC3152) with a carbon purity of >95%, average

length < 1 micron and diameters around 9.5 nm were purchased from Nanocyl S.A. (Sambreville, Belgium).

Commercially available, DGEBA-based epoxy resin (VORAFORCE TF304) and poly(oxypropylene) diamine-

based hardener (VORAFORCE TF366) were supplied by DOW Chemical Company and used as the polymer

matrix with a stoichiometric weight ratio of 100:31. All the chemical agents were of analytical grade and were used

as-received without further purification.

5
2.2 Preparation of epoxy resin mixture for infusion

2.2.1 Neat epoxy mixture

Pre-calculated amounts of resin and hardener were carefully weighed and placed in a beaker. VORAFORCE

TF366, hardener was added to the resin and mixed using a mechanical stirrer at 500 RPM for about 5 min. Curing

being an exothermic reaction lead to gradual increase of the temperature of the mixture.

2.2.2 Dispersion of CNT into epoxy resin

0.1 wt% of CNT was added to the epoxy resin and the mixture was manually stirred for 5 min. The procedure

followed for preparing CNT/Epoxy homogeneous dispersion was same for both pristine nanotubes (PCNT) and

amino functionalized nanotubes (ACNT). It was observed that the viscosity of the resin increased on addition of

CNT during the stirring process. The CNT/Epoxy mixture was then subjected to high shear homogenization (T25),

IKA Inc., at 20,000 rpm for 5 min. It was followed by ultrasonication using a sonicator probe (QSonica 700 with a

power of 700W) at 40% amplitude for 10 min with a shutdown temperature set to 40 oC. The beaker was submerged

in ice-water to prevent the rise in temperature of the solution during the sonication process. The resulting

CNT/Epoxy mixture was again subjected to high shear homogenization for 5 min and sonication for 10 min. The

final CNT/Epoxy mixture was kept in vacuum desiccator at 400 mBar for 1 h in-order to remove the entrapped air

bubbles in the mixture that might have been absorbed from the atmosphere during its processing phase. Finally, the

stoichiometric ratio of hardener was added to the CNT/Epoxy mixture using mechanical stirrer at 500 rpm for 5

min. An exothermic reaction occurred as soon as the hardener was added and to prevent the mixture from becoming

too viscous, it was immediately infused into VARIM.

2.3 Fabrication of laminates

The fabrication of two-ply glass fiber laminates was carried out in VARIM. Firstly, the mold was setup which

was then followed by infusion of the mixture and lastly the recommended curing schedule was carried out. To begin

with the fabrication process, woven roving glass fibers with a nominal weight of 300 g/m 2 were placed on the top of

the composite mold. The glass fiber mat was followed by a layer of peel ply, perforated foil, a runner mesh, and

breather cloth around the corners of the composite mold that covered the vacuum outlets. The sealant tape was

6
pasted on the sides of the composite mold and the mold was covered with vacuum bagging film to enable achieving

a huge vacuum in the mold. Additionally, a pipe was connected to a simple manifold at the center of the glass fiber

mat so that the resin flow was equally distributed in all the directions in the composite mold. Subsequently, an

outlet of the pipe was manually blocked using a stopper to prevent air from entering the mold. After the composite

mold was ready and the vacuum of 1mPa was attained, the final solution i.e. either neat epoxy (described in section

2.2.1) or modified epoxy with CNTs (described in section 2.2.2) was infused into the mold by placing the outlet of

the pipe fully immersed in the beaker containing the mixture. The vacuum was maintained at 1 – 2 mPa throughout

the experiment. After the matrix was successfully infused into the mold, the panel was cured at 70oC for 7 h, giving

the composite an average thickness of 0.7 mm. Figure 1 shows the schematic representation of the processing steps

used to fabricate multi-phase CNT/Epoxy/GF composites. For the experimental study the composites were either

prepared with ACNT or PCNT nanotubes at 0.1wt% concentration of the matrix.

2.4 Characterization and testing

The matrix characterization was performed before and after the curing process. The existence of bonding

between the matrix constituents was verified by Fourier-transform infrared (FT-IR) (Perkin-Elmer FTIR

spectrometer) spectra. The samples of the neat epoxy prior to its curing with hardener, Epoxy/hardener/GF

composites and ACNT/Epoxy/GF composites were analyzed by FTIR. In-order to prepare the samples, the liquid or

the powdered mixture was pressed into a pellet with potassium bromide (KBr) and scanned from 4000 to 400 cm-1

with a resolution of 4 cm-1.

The fracture surfaces of the specimens tested under tensile and flexural loading conditions were examined by

scanning electron microscopy (SEM), JEOL (JSM 6510LV) at 20kV. Prior to the SEM analysis; all the specimens

were sputter-coated with a thin layer of gold to avoid charging effects.

The morphology at the molecular level was investigated using transmission electron microscope (TEM), Hitachi

(H-7500) operating at 110 kV. The images showed the dispersion of CNTs in epoxy resin. One drop of CNT/Epoxy

mixture was dispersed in ethanol and sonicated for 2 h. After sufficient sonication, a drop was placed onto the

surface of a copper grid and excess solvent was removed by using a filter paper. The sample was then observed

under TEM to identify the bonding of CNTs with the epoxy resin.

7
Tensile strength of the composites was determined using Zwick/Roell Z010 Universal testing machine at a

crosshead speed of 2 mm/min. The specimens were prepared according to ASTM D3039 standard. System control

and analysis were performed using TestXpert software supplied by Zwick/Roell. The FRP specimen with length of

250 mm and width of 25 mm were prepared using a marble cutter. Average reading was taken for four samples.

Flexural tests were carried out according to ASTM D790 standard. The strength and modulus were evaluated

using 3-point bending test at a crosshead speed of 1.33 mm/min. For flexural test, the span to width ratio was kept

at 25.4 mm as per ASTM standards since the thickness of specimens was less than 1.6 mm. The test specimen had

length of 60 mm, width of 12.7 mm and prepared using a marble cutter. Four samples were tested for each of the

composite prepared.

3 RESULTS AND DISCUSSION


3.1 Polymerization

Figure 2a illustrates the FT-IR of DGEBA resin supplied by DOW Chemical’s Co. Ltd. The DGEBA resin

contains bisphenol-A-(epichlorhydrin) epoxy with 1, 4-Bis (2, 3 epoxypropoxy) butane and butanedioldiglycidyl

ether groups. The peak of vital importance is seen at 915 cm-1 representing the C-O stretch vibrations in an oxirane

ring. Another peak confirming the presence of epoxy resin is the symmetric stretch of oxirane ring located at 3056

cm-1. These two peaks are clearly visible in Figure 2a. The presence of ether group can be identified by the intense

peak at 1035 cm-1. The stable aromatic rings stretching were confirmed by the peaks at 831, 1184 and 1509 cm -1.

However, after the curing reaction between the epoxy and amine hardener, shifts and elimination of peaks are seen

in Figure 2b which verify that the cross linking has taken place.

Figure 2b shows the FT-IR spectrum of cross-linked VORAFORCE TF304 with amine hardener in GF

composite. The peak representing the oxirane ring (915 cm-1) disappeared after the curing process. The

phenomenon behind the disappearance of the peak from the spectrum is due to the ring-opening polymerization

reaction. However, several new peaks are observed after the curing process is completed. A broad peak appeared at

3431 cm-1 which is likely due to the secondary amine (N-H) and O-H stretching. Interestingly, strong peak observed

at 2925 cm-1 is due to the –CH2– group stretching vibrations. Additionally, the strong band observed at 1609 cm-1

represents the stretching of – OH group. Furthermore, the presence of several IR bands at 1602, 1580, 1509 cm-1

illustrate the stretching of aromatic rings. In addition, the appearance of bands at 1459 and 1381 cm -1 are possibly

due to the bending of –CH2– and –CH3– groups. The stretching of –C-C-O-C- can be seen by analyzing the peaks

8
at 1251 and 1182 cm-1. Furthermore, the peak at 1106 cm-1 characterizes the stretching of –C-N group while the

peak at 1038cm-1 is due to the stretching of aliphatic ring (-C-O-C-). Therefore, the absence of transmittance of

epoxide ring and presence of OH group and C-N group, as shown in Figure 2b validate the crosslinking

polymerization process.

Figure 2c shows the IR spectra of the ACNT/Epoxy/GF composite and the epoxide group characteristic peak at

915 cm-1 is not present in the spectra and several news peaks have emerged due to the addition of ACNT in the

matrix. The single band at 3431 cm-1 shows the stretch of primary amine group from the ACNT. The symmetric and

asymmetric vibrations of ethylenediamine molecules have been identified by the peaks at 2964, 2928 and 2870 cm -
1
. In addition, the presence of bands at 1581 and 1247 cm -1, which corresponds to the N-H in plane and C-N

stretching, respectively, is possibly due to the available amino groups on the surface of CNT. The peak at 1182 cm-1

represents the stretching of amide group. Lastly, the small peak at 1085 cm -1 represents the stretching of C-N of

amine molecules. Therefore, the FT-IR spectrum (Figure 2c) validates the coupling reaction between DGEBA and

amino functionalized nanotubes. Furthermore, the verification of the stronger interfacial bonding between ACNT

and the epoxy as borne out from the TEM and SEM micrographs are discussed in the following section.

3.2 Microstructures of epoxy resin containing CNT

The microstructure of amino-functionalized nanotubes in epoxy resin was investigated by TEM. Figure 3 shows

the morphology of 0.1 wt% of amino functionalized nanotubes dispersed in epoxy resin and the cross-linking prior

to its curing stage. The liquid sample (ACNT/epoxy) was placed on the grid and analyzed under the microscope. It

is interesting to see that epoxy molecules have build covalent bonds with the surface of CNTs at multiple sites in

the samples being examined. Moreover, ACNTs (with approximately 9.5 nm diameter and length of <1 micron) are

clearly visible in the TEM image (Figure 3) and in-fact individual strands of ACNT dimensions can be seen clearly

i.e. there is no agglomeration of ACNTs in epoxy resin. Also, the images did not show the presence of any physical

damage on the walls of ACNT which possibly could have occurred during the processing stage. FT-IR peaks and

TEM images are in correlation to support the stronger chemical interaction between the amino group of CNTs and

the epoxide of resin prior to its curing with the amine hardener. Similar observations of improved interfacial

adhesion of ACNT with the epoxy matrix have been examined by other researchers as well [14, 25, 52].

9
Figure 4 is a representation of the mechanism of nucleophilic polymerization reaction between the epoxy groups

and the primary amine groups of carbon nanotubes. As described in Reaction 1 (Figure 4), after dispersing amino

functionalized CNT in epoxy resin, a chemical reaction takes place between amine group of CNTs and epoxide

monomer group that result in a nucleophilic ring-opening reaction and formation of a polymer. The primary amine

undergoes a reaction process to form secondary amines and continue to build longer polymer chains. The product

from reaction 1 further undergoes a chemical reaction with another epoxide molecule to form a polymer chain as

shown in Reaction 2 mechanism. Lastly, the resulting polymer chain further interacts with other epoxide groups

present and starts building cross-linking polymer chain as shown in Figure 4. The presence of primary and

secondary amine groups has previously been observed in FT-IR spectrum (Figure 2c). This cross-linking reaction

builds an interlocking mechanism in the resin which hinders the mobility of polymer chains in the system. Hence,

the chemical interaction between ACNT and epoxy resin is effectively improved due to the presence of strong

covalent bonds between them. This strong interfacial interaction is expected to facilitate efficient load transfer from

the matrix to CNTs and thus result in improved mechanical behavior [29, 44, 45, 53].

3.3 Mechanical properties of composites

3.3.1 Effect of resin: hardener ratio

Composites were fabricated with amine-rich resin composition (referred to as 69%) and were mechanically

tested to compare with the stoichiometric composition (referred to as 76%). Comparison was carried out for both

tensile and flexural properties of the epoxy-glass fiber 2-ply laminates. The results obtained for the experimental

study match with the reported results in literature [51, 52, 54 - 58]. The mechanical behavior of the composites is

dependent upon the epoxy: hardener ratio. In the matrix with stoichiometric ratio (i.e. 76%), all the amine hydrogen

react with all epoxide groups resulting in the formation of an interconnected network. Whereas, in the matrix with

excess of epoxy resin, all the primary and secondary amines react with 80% of the epoxy groups in DGEBA resin.

The left 20% epoxy molecules are still freely available in the matrix that do not react and thus affect the mechanical

properties. While in the case of matrix with excess amine, all the epoxy groups react with primary amines while

leaving a few secondary amines unreacted. According to Morgan et al. [59], epoxy tends to give branched

structures, and possibly results in rings with partially reacted triamine molecules and epoxide chain. In this study,

10
the effect of excess epoxy groups in the resin on the mechanical behavior as compared to the availability of more

amine groups in the matrix has also been investigated.

The maximum tensile modulus (6.4 GPa) is obtained for the stoichiometric epoxy/amine ratio (76%

composition). Samples with an excess of amine (69% composition) result in a significant increase in some

mechanical properties, such as flexural strength (+ 77%), flexural modulus (+ 79%), and tensile strength (+ 14%).

These results are not in agreement with some of the results reported in literature [51, 55, 56, 58] wherein a decrease

in flexural modulus is reported on increasing the amine content in the matrix. However, the experimental results

obtained showed enhanced ductile behavior (for excess amine specimens) as compared to a brittle behavior (for

specimens with stoichiometric ratio). It is possibly due to the fact that, free uncured molecules are available in the

composite which tend to further increase the ductility of the laminates. Thus, the effect on flexural modulus,

flexural and tensile strength is more remarkable in the amine-rich resin while tensile modulus shows better

properties in epoxy-rich resin.

3.3.2 CNT/Epoxy/GF composites

The mechanical behavior of CNT/Epoxy/GF composite depends on the dispersion of CNTs and interfacial

interaction among the composite constituents. For this study, low viscosity resin was chosen to allow easy infusion

and improved wettability during VARIM manufacturing of the polymer nanocomposites (PNC). The viscosity of

the epoxy resin was 1400 mPa.s at 25oC and it increased as CNT content increased in the resin. Interestingly, the

rise in temperature of the resin to 50oC during the processing stages lowered the viscosity of the resin and assisted

in better dispersion of CNT.

Figure 5 shows the tensile and flexural strength of the laminates Vs. type of CNT in the matrix. In-order to

compare the mechanical behavior of the composites with the addition of CNTs in the matrix, Epoxy/GF composite

has been taken as the reference, as shown in the figure. The tensile strength increased by 42% on addition of PCNT

and decreased by 15% for ACNT composites. The flexural strength of composites showed an improvement of 74%

and 46% in PCNT/Epoxy/GF and ACNT/Epoxy/GF composites respectively, as compared to the control sample.

The significant increase in flexural properties of the composites shows a clear evidence of strong interfacial

bonding between ACNT and the matrix as previously confirmed by TEM analysis (Figure 3). Similar results for

improved tensile properties in epoxy matrix incorporated with CNTs were also reported by other researchers [26, 60

11
- 62]. However, a slight decrease in the tensile strength of ACNT/Epoxy/GF composite is possibly due to the

stiffening effect caused by the strong covalent bond between the amino groups on the surface of CNT and the

matrix. Another study by Ci and Bai [63] reported that there is a negligible effect on the mechanical properties by

the addition of CNT in the base epoxy matrix if the matrix is too rigid. The epoxy matrix used for the present

experimental study has a tensile modulus of approximately 3.3 GPa which is significantly greater than 2.4 GPa

reported in the study by Ci [63]. Since the matrix chosen is more rigid, decrease in tensile properties were expected.

The soft and ductile matrix allows stronger interfacial interaction whereas the reverse would be true for rigid

matrix.

It can be observed from histograms in Figure 6 that there is significant improvement in the flexural modulus by

incorporating CNT in the FRP composites. The improved interfacial adhesion between CNT and epoxy results in an

increase in flexural modulus under compression. Iwahori et al. [64] and Sanchez et al. [65] have also documented

improvement in flexural properties over tensile properties in CNT/Epoxy/Fiber composites. On the contrary, slight

drop in tensile modulus was observed in ACNT while an increase of 26% in PCNT composites since the properties

in the longitudinal direction is dependent on the continuous fiber. Other researchers have also faced a challenging

task to improve the elastic modulus by incorporating CNT in the composites [17, 47, 64, 66, 67]. The experimental

results demonstrate significant improvement in the flexural modulus and strength by incorporating nanotubes in the

polymer composites. The improvement in flexural properties is due to the fiber breakage followed by small

increments in strength possibly due to the matrix cracking and local delamination occurring inside the sample.

Moreover, the flexural properties are matrix dominated while the tensile ones are fiber dominated. Therefore, the

stronger bonding between the matrix and CNT is highlighted by the significant improvement observed in the

flexural response of the samples tested.

The reinforcement of Epoxy/GF composites with CNTs in stoichiometric ratio have for the first time shown that

pristine CNTs are better nanofiller than amine-functionalized CNTs. Nevertheless, the presence of amino groups

had a positive effect on the degree of dispersion. It has also been found that this type of CNTs tend to form re-

agglomerates, due to attraction between the amine groups on the surface of CNT. Moreover, at stoichiometric ratio,

it is believed to have a fully cross-linked structure and the presence of additional amino group of CNTs may have

not found enough un-reacted epoxide groups for covalent bonding. In addition, during the functionalization of

nanotubes they are oxidized in high concentration acids, which often introduce damages on the surface of CNT,

12
reducing its aspect ratio, thus resulting in diminished mechanical properties of CNTs and the composites [68, 69].

As a result, an effective stress was not transferred to ACNT during the tensile loading conditions and lower

properties were obtained as compared to those of neat epoxy and PCNT composites. Under flexural loading

conditions, PCNT has again achieved the maximum modulus and strength while ACNT has got intermediate results

between PCNT and neat epoxy composites. During the testing it was seen that neat epoxy resin composite failed

immediately (brittle fracture) after reaching maximum stress level. However, for PCNT composites, a gradual

decrease in load was observed. This ductile response is possibly due to the slower crack propagation through the

fibers, which was followed by an increase in strength due to matrix cracking and local delamination occurring in the

specimen.

Therefore, the experimental results highlight that PCNT are better reinforcing fillers than ACNT in epoxy resin

at stoichiometric ratio. While the results obtained for ACNT composites with stoichiometric ratio are in agreement

with those stated by Gude et al.[51] i.e. in a highly cross-linked network, the addition of amino-CNTs might have a

detrimental effect during the cure, delaying the peak of reaction or preventing the resin to obtain their optimum

properties [70 - 72].

The next parameter to investigate is the effect of CNTs in amine-rich resin composition. Table 1 summarizes the

experimental results obtained for both the tensile and flexural tests conducted along with the results obtained with

stoichiometric ratio. Interestingly, the maximum enhancement in tensile and flexural properties among all the

samples were observed for ACNT/Epoxy/GF (69% resin) composites (row 3b in Table 1) i.e. an increase in tensile

modulus by 39%, tensile strength by ~ 38%, and flexural modulus by ~ 43% and flexural strength by 95% as

compared to the neat epoxy sample prepared with stoichiometric ratio. Whereas, PCNT/Epoxy/GF (69% resin)

composite showed an improvement of 5% in tensile modulus, 52% in flexural modulus and 71% in flexural strength

(row 3a in Table 1). These results are in contradiction to those reported by Gude et al.[51] which mentioned that

reinforcement effect of CNT is almost negligible in amine-rich resin composition. Their results showed an

improvement in flexural modulus in amine-rich composition over stoichiometric but not comparable to epoxy-rich

resin system. However, our results are different from their study possibly due to the use of different curing agent in

our case. It is believed that the curing mechanism varies for the type of resin and hardener used. It is also possible

that the excess amine groups in the matrix underwent nucleophilic ring-opening mechanism with the free available

13
epoxide groups and formed strong interconnected network with the capability of transferring the load from the

matrix to CNTs under stress.

In a recent study by Sanchez et al. [65] it was found that maximum increase in flexural strength was measured

at lower concentration of pristine nanotubes and at high percentages of amino-functionalized nanotubes. Hence, if

the concentration of ACNT is increased in the matrix it may further enhance the mechanical performance (more

than PCNT/Epoxy/GF composite) of ACNT/Epoxy/GF composites in both stoichiometric and amine-rich resin

systems. Tensile modulus and strength seems to be positively affected by using ACNT over PCNT in the amine-

rich matrix composition due to the good interfacial bonding between the matrix and amino-CNT. On comparing 3a

and 3b (Table 1), it can be seen that there is significant enhancement in mechanical properties obtained for

ACNT/Epoxy/GF (69% resin) composition. The reason(s) for this is not clear and is further being investigated.

3.4 Morphology

Results from the scanning electron micrograph (SEM) study support the quantitative results obtained through

mechanical testing of the various composites. Figures 7 (a) – (d) shows the SEM images of the fractured control

samples and samples with 0.1 wt% ACNT and PCNT embedded in epoxy matrix when reinforced by glass fibers.

The fractured surfaces under SEM showed rougher surface in all the samples containing CNTs as compared to the

neat epoxy sample (Figure 7a). For the composites prepared with PCNT large CNT agglomerates were clearly seen

(Figure 7c) whereas, homogenously dispersed particles were seen for the samples prepared with amino-CNTs

(Figure 7b). The images have confirmed the successful dispersion of ACNT in the epoxy matrix and highlight the

stability of the dispersed amino-CNT in the matrix after the curing process. Similar results with respect to ACNTs

illustrating better bonding for longer time durations have also been reported by other researchers [14].The presence

of agglomerates in SEM images is due to the van der Waals attraction between the nanotubes and act as stress

concentrating sites during the tensile and flexural loading conditions. The number of agglomerate locations

observed in Figure 7b and 7c are less, but it increases with the increase in concentration of CNT in the epoxy

matrix. The incorporation of CNT in the epoxy matrix assists in postponement of crack initiation or hinders the

crack propagation either by deflection or crack bridging. The 3-D network observed in ACNT (amine-rich resin)

sample (Figure 7d) provides a visual of the strong bonding between the modified matrix and the glass fiber that

results in slower crack propagation and minimal pull-out of nanotubes that is characterized by increased mechanical

14
properties as mentioned in Table 1. In addition, careful examination of Figure 7d shows the nanotubes are coated

with the epoxy resin due to the improved interfacial adhesion between the CNT and epoxide group by the amino

groups on the surface of CNT.

The multi-phase composites with addition of CNTs have shown maximum improvement in the mechanical

properties that is a result of enhanced interfacial bonding between the nanotubes and the matrix. The matrix is

modified by using nanotubes as secondary reinforcement that participates in reducing the crack initiation and

propagation sites. The phenomenon of CNT bridging in CNT/Epoxy/GF composites are well in agreement with the

previously reported crack mechanism studies related with CNT bridging effect [53, 67] that increases the fracture

toughness of the composites.

From the results studied, it can be inferred that there are various parameters such as processing techniques used

to disperse CNT in epoxy matrix, amount of resin content, concentration of CNT in the matrix etc. which affect the

mechanical properties of the multi-phase composites as the mechanism behind the scene is very complex. The

proper understanding of such complex mechanisms would require further in depth investigation. The chemical

bonding varies for the type of CNT used in the matrix and its consequent bonding with GF mat.

4 CONCLUSION
In the present study, pristine and amino-functionalized CNTs have been incorporated in E-glass/Epoxy

composites to enhance the mechanical properties. The observations made in the present study illustrate that it is a

complex mechanism to transfer the intrinsic properties of CNTs in a multi-phase composite. The addition of pristine

CNT in the stoichiometric matrix composition has shown improvement by 26% in tensile modulus, 42% in tensile

strength, 50% in flexural modulus and 74% in flexural strength, while with amino functionalized CNTs the flexural

modulus and strength increased by 36% and 46% respectively, as compared to neat epoxy composite properties.

Remarkable results were obtained when amino-functionalized nanotubes were dispersed in amine-rich resin

composition i.e. an increase in tensile modulus by 39%, tensile strength by 38% and flexural modulus by 43% and

flexural strength by 95% as compared to the control sample. On the contrary, pristine nanotubes did not show

improvement in tensile properties but showed better performance in flexural modulus and strength in amine-rich

resin system as compared to the control sample. The enhanced properties of ACNT in amine-rich resin system is

possibly due to the highly cross-linked network built between the amino-functionalized CNTs and the epoxide

group of resin during the early curing stage which was minimal in the case of pristine CNT. The stronger interfacial

15
bonding with the matrix and efficient stress transfer from the matrix to CNTs results in improved mechanical

properties.

TEM and SEM images have also confirmed the formation of chemical bonds between the epoxy resin, CNTs

and glass fiber mat. Especially, the CNT-bridging mechanism in the crack propagation regions as observed in SEM

images has verified the strong bond between amino-CNT, epoxy matrix and the glass fiber. FT-IR spectrums for the

samples are in good agreement with the chemical reactions occurring between the nanotubes and the matrix which

have further been correlated with improved mechanical properties.

This is the first study in which the composites with PCNT reinforcement in epoxy resin have shown higher

mechanical properties as compared to those with ACNT under certain conditions. This study has been able to

demonstrate the relation between the functionalized and non-functionalized CNTs and the mechanical behavior of

CNT/Epoxy/GF composites. However, the reason for significant enhancement of tensile properties of ACNT

composites prepared in excess amine systems needs to be further investigated by considering few other parameters

such as infusion temperature of the epoxy mixture. A few other critical issues which need to be further explored in

detail are the effect of nano-level dispersion of CNT in epoxy resin, degree of functionalization of amino-modified

CNT, viscosity control at different stages of processing and minimal void formation and other defects during

processing.

5 ACKNOWLEDGMENTS
This project was supported by the Naval Research Board of India (Project No. 268). The authors would also like

to thank Dr. Manoranjan Patri, Dr. Debdatta Ratna, and Dr. Sangram K. Rath, NMRL for their guidance with the

preparation of composites. The authors would like to thank SAIF Labs, Punjab University, Chandigarh, India for

carrying out TEM analysis of the samples. Also, the help of SAI Labs, Thapar University, Patiala and Dr. Sunil

Chaudhary and his team working at NIPER, Mohali, for the SEM images of fractured composite surfaces is duly

acknowledged.

6 REFERENCES
[1]. Lata D. Fabrication of Decorative Fluoroscent Composite Material. Department of Physics, National Institute

of Technology Rourkela, 2000: 6-7.

[2]. Jin Z., Khunlavit C., Shuai H., Chun H., Wang. Hybrid composite laminates reinforced with glass/carbon

woven fabrics for lightweight load bearing structures. Materials and Design 2011; 36: 75 - 80.

16
[3]. Dalina W.A.D.W, Mariatti M., Ramlee R., Ishak Z.A.M., Mohamed A.R. Comparison on the properties of

Glass Fiber/MWCNT/EPoxy and Carbon Fiber/MWCNT/Epoxy Composites. Advanced Materials Research

2014; 858: 32-39.

[4]. Wong E.W., Sheehan P.E.& Lieber C.M. Nanobeam mechanics: elasticity, strength and toughness of nanorods

and nanotubes. Science 1971; 277: 1971 - 1975.

[5]. Ajayan P.M., Stephan O., Colliex C., Trauth D. Aligned carbon nanotube arrays formed by cutting a polymer

resin nanotube composite. Science 1994; 265 (5176): 1212 - 1214.

[6]. Ebbesen T. W., Lezec H. J., Hiura H., Bennett J. W., Ghaemi H. F., Thio T. Electrical conductivity of

individual carbon nanotubes. Nature 1996; 382 (6586): 54-56.

[7]. Treacy M., M.J., Ebbesen T.W. & Gibson J.M. Exceptionally high Young's modulus observed for individual

carbon nanotubes. Nature 1996; 381: 678-680.

[8]. Lu, J.P. Elastic properties of nanotubes and nanoropes. Phys. Rev. Lett 1997; 79: 1297 - 1300

[9]. Thostenson E.T., Ren Z., Chou TW. Advances in the science and technology of carbon nanotubes and their

composites: a review. Composites science and technology 2001; 61 (13): 1899-1912.

[10]. Morais WA, de D'almedia JRM, Godefroid LB. Effect of the fiber reinforcement on the low energy impact

behavior of fabric reinforced resin matrix composite materials. J. Braz. Soc. Mech. Sci. & Eng 2003; 25.

[11]. Iijima S. Helical microtubules of graphitic carbon. Nature 1991; 354: 56-58.

[12]. Sandler J., Shaffer M.S.P.,Prasse T., Bauhofer W., Schulte K., Windle A.H. Development of a dispersion

process for carbon nanotubes in an epoxy matrix and the resulting electrical properties. Polymer 1999; 40 (21):

5967 - 5971.

[13]. Li C., Chou T-W. Electrical conductivities of composites with aligned carbon nanotubes. J Nanosci

Nanotechnol 2009; 9(4): 2518 - 2524.

[14]. Ma P.C., Siddiquia N.A., Maromb G., Kima J.K. Dispersion and Functionalization of Carbon Nanotubes for

Polymer-Based Nanocomposites: A Review. Composites Part A- Appl. Science and Manufturing 2010; 41(10):

1345-1367.

[15]. Min-Feng Yu, Oleg Lourie, Mark J. Dyer, Katerina Moloni, Thomas F. Kelly, Rodney S. Ruoff. Strength and

Breaking Mechanism Of Multiwalled Carbon Nanotubes Under Tensile Load. Science 2000; 287(5453): 637-

640.

17
[16]. Bekyarova E., Thostenson E.T., Yu A., Itkis M.E.., Fakhrutdinov D., Chou T.W. Functionalized single-walled

nanotubes for carbon fiber-epoxy composites. J Phys Chem C 2007; 111(48): 17865-17871.

[17]. Bekyarova E., Thostenson E.T., Yu A., Kim H., Gao J., Tang J., Hahn H. T., Chou T.W. , Itkis M. E., Haddon

R. C. Multiscale Carbon Nanotube-Carbon Fiber Reinforcement for Advanced Epoxy Composites. Journal of

Langmuir (ACS Publications) 2007; 23(7): 3970–3974.

[18]. Fiedler B., Gojny F.H., Wichmann M.H.G., Nolte M.C.M., Schulte K. Fundamental aspects of nano-reinforced

composites. Composites Science and Technology 2006; 66(16): 3115-3125.

[19]. Schadler L.S., Kumar S.K., Benicewicz B.C., Lewis S.L., Harton S.E. Designed interfaces in polymer

composites: A fundamental viewpoint. MRS Bull 2007; 32: 335 - 340.

[20]. Yeh M.K., Hsieh T.H., Tai N.H. Fabrication and mechanical properties of multi-walled carbon

nanotubes/epoxy nanocomposites. Mater Sci Eng A 2008; 483-484: 289 - 292.

[21]. Bryning M.B., Islam M.F., Kikkawa J.M., Yodh A.G. Very low conductivity threshold in bulk isotropic single-

walled carbon nanotube-epoxy composites. Adv. Matter 2005; 17: 1186 - 1191.

[22]. Li J., Ma P.C., Chow W.S., To C.K., Tang B.Z., Kim J.K. Correlations between percolation threshold,

dispersion state and aspect ratio of carbon nanotube. Adv Funct Matter 2007; 17: 3207 - 3215.

[23]. Kim J.K., Mai Y-W. High strength, high fracture toughness fibre composites with interface control - a review.

Compos Sci Technol 1991; 41(4): 333 - 378.

[24]. Kim J.K., Mai Y.W. Interfaces in composites, In: Chou TW, editor. Materials science and technology: structure

and properties of composites 1993; 13: 229 - 289.

[25]. Gojny F.H., Schulte K. Functionalisation effect on the thermo-mechanical behavior of multi-wall carbon

nanotube/epoxy composites. Compos Sci Technol 2004; 64: 2303 - 2308.

[26]. Thostenson E.T., Chou T.W. Processing-structure-multi-functional property relationship in carbon

nanotube/epoxy composites. Carbon 2006; 44(14): 3022-3029.

[27]. Seyhan A.T., Gojny F.H., Tanoglu M., Schulte K. Critical aspects related to processing of carbon

nanotube/unsaturated thermoset polymer polyester nanocomposites. Eur. Polym. J. 2007; 43: 374 - 379.

[28]. Thostenson E.T., Ziaee S., Chou T.W. Processing and electrical properties of carbon nanotube/vinyl ester

nanocomposites. Compos Sci. Technol. 2009; 69(6): 801 - 804.

18
[29]. Shen J., Huang W., Wu L., Hu Y., Ye M. The reinforcement role of different amino-functionalized multi-

walled carbon nanotubes in epoxy nanocomposites. Composites Science and Technology 2007; 67: 3041-3050.

[30]. Gojny F.H., Nastalczyk J., Roslaniec Z., Schulte K. Surface modified carbon nanotubes in CNT/epoxy-

composites. Chem Phys Lett. 2003; 370: 820 - 824.

[31]. Chandrasekaran V.C.S., Santare M.H., Krishnan P., Advani S.G. Amino functionalization of MWCNTs and

their effect on ILSS of hybrid nanocomposites. Compos Interface 2011; 18(4): 339 - 355.

[32]. Song Y.S., Yuon J.R. Influence of dispersion states of carbon nanotubes on physical properties of epoxy

nanocomposites. Carbon 2005; 43.

[33]. Kovacs J.Z., Andresen K., Pauls J.R., Garcia C.P., Schossig M., Schulte K. Analyzing the quality of carbon

nanotube dispersions in polymers using scanning electron microscopy. Carbon 2007; 45: 1279 - 1288.

[34]. Kinloch I.A., Roberts S.A., Windle A.H. A rheological study of concentrated aqueous nanotube dispersions.

Polymer 2002; 43: 7483 - 7491.

[35]. Huang Y.Y., Ahir S.V., Terentjev E.M. Dispersion rheology of carbon nanotubes in a polymer matrix. Phys.

Rev. B 2006; 73: 125422-1-9.

[36]. Rahatekar S.S., Koziol K.K.K., Butler S.A., Elliott J.A., Shaffer M.S.P., Mackley M.R. Optical microstructure

and viscosity enhancement for an epoxy resin matrix containing multiwall carbon nanotubes. J. Rheol. 2006;

50(5): 599 - 610.

[37]. Eitan A., Kuiyang J., Dukes D., Andrews R., Scadler L.S. Surface modification of multiwalled carbon

nanotubes: toward the tailoring of the interface in polymer composites. Chem. Matter 2003;15: 3198 - 3201.

[38]. Ramanathan T., Fisher F.T., Ruoff R.S., Brinson L.C. Amino-functionalized carbon nanotubes for binding to

polymers and biological systems. Chem. Matter 2005;17: 1290 - 1295.

[39]. Wang S., Liang Z., Liu T., Wang B., Zhang C. Effective amino functionalization of carbon nanotubes for

epoxy polymer composites. Nanotechnology 2006; 17: 1551 - 1557.

[40]. Li J., Fang Z., Tong L., Gu A., Liu F. Improving dispersion of multiwalled carbon nanotubes in polyamide six

composites through amino-functionalization. J. Appl. Poly. Sci. 2007; 106: 2898 - 2906.

[41]. Spitalsky Z., Matejka L., Slouf M., Konyushenko E.N., Kovarova J., Zemek J., Kotek J. Modification of

carbon nanotubes and its effect on properties of carbon nanotube/epoxy nanocomposites. Polymer composites

2009; 30(10): 1378-1387.

19
[42]. Guadagno L., Vivo B. De, Bartolomero A. Di, Lamberti P., Sorrentino A., Tucci V., Vertuccio L., Vittoria V.

Effect of functionalization on the thermo-mechanical and electrical behavior of multi-walled carbon

nanotube/epoxy composites. Carbon 2011; 49 (6): 1919-1930.

[43]. Theodore M., Hosur M., Thomas J., Jeelani S. Influence of functionalization on properties of MWCNT/Epoxy

nanocomposites. Materials Science and Engineering:A 2011; 528(3): 1192-1200.

[44]. Salam M.B.A., Hosur M.V., Zainuddin S., Jeelani S. Improvement in mechanical and thermo-mechanical

properties of epoxy composite using two different functionalized multi-walled carbon nanotubes. Open Journal

of Composite Materials 2013; 3: 1-9.

[45]. Rahman M.M., Zainuddin S., Hosur M.V., Malone J.E., Salam M.B.A., Kumar A., Jeelani S. Improvements in

mechanical and thermo-mechanical properties of e-glass/epoxy composites using amino functionalized

MWCNTs. Composite Structures 2012; 94: 2397 - 2406.

[46]. Wang J.G., Fang Z.P., Gu A.J., Xu L.H., Liu F. Effect of amino-functionalization of of multi-walled carbon

nanotubes on the dispersion with epoxy resin matrix. J. Appl. Polym. Sci. 2006;100: 97 - 104.

[47]. Gojny F.H., Wichmann M.H.G. , Fiedler B. , Bauhofer W., Schulte K. Influence of nano-modification on the

mechanical and electrical properties of conventional fibre-reinforced composites. Composites Part A 2005; 36:

1525-1535.

[48]. Zhihang F., Santare M.H., Advani S.G. Interlaminar shear strength of glass fiber reinforced epoxy composites

enhanced with multi-walled carbon nanotubes. Composites Part A 2008; 39: 540 - 554.

[49]. Yokozeki T., Iwahori Y., Ishiwata S. Matrix cracking behaviors in carbon fiber/epoxy laminate with cup-

stacked carbon nanotubes (CSCNTs). Composites Part A 2007; 38: 917 - 924.

[50]. Brouwer W.C., Van Herpt E., Labordus M. Vacuum injection moulding for large structural applications.

Compos Part A: Appl Sci Manuf 2003; 34(6): 551 - 8.

[51]. Gude M.R., Prolongo S.G., Urena A. Effect of the epoxy/amine stoichiometry on the properties of carbon

nanotube.epoxy composites. J. Therm. Anal. Calorim. 2012; 108: 717 - 723.

[52]. Prolongo S.G., Gude M.R., Urena A.The curing process of epoxy/amino-functionalized

MWCNTs:Calorimetry, Molecular Modelling, and Electron Microscopy. Journal of Nanotechnology 2010; 11.

20
[53]. Gojny F.H., Wichmann M.H.G. , K€opke U. , Fiedler B. , Schulte K. Carbon nanotube-reinforced epoxy-

composites: enhanced stiffness and fracture toughness at low nanotube content. Journal of Composites Science

and Technology 2004; 64: 2363–2371.

[54]. Calventus Y., Montserrat S., Hutchinson JM. Enthalpy relaxation of non-stoichiometric epoxy-amine resins.

Polymer 2001; 42: 7081 - 7093.

[55]. Meyer F., Sanz G., Eceiza A., Mondragon I., Mijovie J. The effect of stoichiometrix and thermal history during

cure on structure and properties of epoxy networks. Polymer 1995; 36: 1407 - 1414.

[56]. Fernandez-Nograo F., Valea A., Llano-Ponte R., Mondragon I. Dynamic and mechanical properties of

DGEBA/poly(propylene oxide) amine based epoxy resins as a function of stoichiometry. Eur. Polym. J. 1996;

2: 257 - 66.

[57]. Vallo CI., Frontini PM., Williams RJJ. The glass transition temperature on nonstoichiometric epoxy-amine

networks. J. Polym. Sci. B. 1991; 29: 1503 - 11.

[58]. Palmese GR., McCullough RL. Effect of epoxy-amine stoichiometry on cured resin material properties. J.

Appl. Polym. Sci. 1992; 46: 1863 - 73.

[59]. Morgan RJ, Kong FM, Walkup CN. Structure-property relations of polyethertriamine-cured bisphenol-A-

diglycidyl ether epoxies. Polymer 1984; 25: 375 - 86.

[60]. Mallick P.K. Fiber-Reinforced Composites: Materials, Manufacturing and Design. New York: CRC Press,

Taylor & Francis Group, 3rd Edition, 2013.

[61]. Shen J., Huang W., Wu L., Hu Y., Ye M. Thermo-physical properties of epoxy nancomposites reinforced with

amino-functionalized multi-walled carbon nanotubes. Composites Part A 2007; 38.

[62]. Diez-Pascual A., Ashrafi B., Naffakh M., Gonza lez-Dominguez J., Johnston A., Simard B., Martinez M.,

Gomez-Fatou M. Influence of carbon nanotubes on the thermal, electrical, and mechanical properties of

poly(ether ether ketone)/glass fiber laminates. Carbon 2011; 49.

[63]. Ci L., Bai J. The reinforcement role of carbon nanotubes in epoxy composites with different matrix stiffness.

Compos. Sci. Technol. 2006; 66: 599 - 603.

[64]. Iwahori Y., Ishiwata S., Sumizawa T., Ishikawa T. Mechanical properies improvements in two-phase and

three-phase composites using carbon nanofiber dispersed resin. Compos. Part A: Appl. Sci. 2005; 36 (10):

1430 - 1439.

21
[65]. Sanchez M., Campo M., Suarez A.J, Urena A. Effect of the carbon nanotube functionalizaton on flexural

properties of multiscale carbon fiber/epoxy composites manufactured by VARIM. Composites: Part B 2013;

45: 1613 - 1619.

[66]. Quaresimin M., Varley RJ. Understanding the effect of nano-modifier addition upon the properties of fiber

reinforced laminates. Compos. Sci. Technol. 2008; 68 (3-4): 718 - 726.

[67]. Godara A., Mezzo L., Luizi F., Warrier A., Lomov S.V., Vuure A.W. van, Gorbatikh L. , Moldenaers P.,

Verpoest I. Influence of carbon nanotube reinforcement on the processing and the mechanical behaviour of

carbon fiber/epoxy composites. Carbon 2009; 47: 2914-2923.

[68]. Monthioux M., Smith BW., Burteaux B., Claye A., Fischer JE., Luzzi DE. Sensitivity of single-wall carbon

nanotubes to chemical processing: an electron microscopy investigation. Carbon 2001; 39: 1251 - 72.

[69]. Shen K., Xu H., Jiang Y., PietraB T. The role of carbon nanotube structure in purification and hydrogen

adsorption. Carbon 2004; 42: 2315 - 22.

[70]. Bae J., Jang J., Yoon S.H. Cure behavior of the liquid-crystalline epoxy/carbon nanotube system and the effect

of surface treatment of carbon fillers on cure reaction. Macromol. Chem. Phys. 2002; 203: 2196-2204.

[71]. Lopez J., Rico M., Montero B., Diez J., Ramirez C. Polymer blends based on an epoxy-amine thermoset and a

themoplastic. J. Therm. Anal. Calorim. 2009; 95: 369-376.

[72]. Rico M., Lopez J., Bouza R. Thermal behavior of blends based on a thermoplastic-modified epoxy resin with a

crosslinking density variation. J. Therm. Anal. Calorim. 2011; 105(2): 599-606.

22
List of Figures and Tables in the manuscript

1. Figure 1: Schematic of CNT/Epoxy/ GF composites preparation


2. Figure 2: FT-IR spectra of different composites ((a) neat epoxy, (b) epoxy/hardener/GF
composite, (c) ACNT/Epoxy hardener/GF composite)
3. Figure 3: TEM image showing dispersion of 0.1wt% ACNT in epoxy resin
4. Figure 4: Mechanism of reactions between ACNT and epoxide group
5. Figure 5: Effect of CNT composition on the tensile and flexural strength
6. Figure 6: Effect of CNT composition on the tensile and flexural modulus
7. Figure 7: Fracture surface of E-glass fiber/epoxy composites containing: (a) 0.0% CNT;
(b) 0.1wt% ACNT; (c) 0.1wt% PCNT; (d) 0.1 wt% ACNT (69% resin).
8. Table 1: Summary of mechanical properties obtained.

23
Figures:

Figure 1: Schematic of CNT/Epoxy/ GF composites preparation

Figure 2: FT-IR spectra of different composites ((a) neat epoxy,

24
Figure 2: FT-IR spectra of different composites (b) Epoxy/Hardener/GF composite

Figure 2: FT-IR spectra of different composites (c) ACNT/Epoxy/GF composite

25
Figure 3: TEM image showing dispersion of 0.1wt% ACNT in epoxy resin

26
Reaction1: Reaction of ACNT with epoxide group of epoxy resin

H O
H2
MWCNT N H2C C MWCNT N C CH

H H H OH

Amino Epoxide
functionalized molecule of
MWCNT molecule epoxy resin

Reaction 2: Reaction of product from Reaction 1 with another epoxide molecule of epoxy resin

OH

H2C C
O
H2
MWCNT N C CH MWCNT N H
H2C C
H
H OH H
H2C C
MWCNT-NH2 and Epoxide 2nd Epoxide
molecule from (1)
molecule OH

Reaction 3: Cross-linking polymerization reaction between ACNT and epoxy molecules

H2 H
OH O C C

H2C C O H2C C
OH
MWCNT N H H2C C MWCNT N H
H H
H2C C H OH
H2C C
OH H2
O C C
H

MWCNT-NH2 and two epoxide Several epoxide Polymer chain of epoxide


molecules available in molecules linked with NH2 of
molecule from (2) matrix MWCNT in matrix

Figure 4: Mechanism of reactions between ACNT and epoxide group

27
400 TENSILE
FLEXURAL
350

REFERENCE
300
STRENGTH (MPa)

250

200

150

100

50

0
0.0 WT% CNT 0.1 WT% PCNT 0.1 WT% ACNT

CNT COMPOSITION IN LAMINATES

Figure 5: Effect of CNT composition on the tensile and flexural strength

12
TENSILE
11
FLEXURAL
10

9
REFERENCE
8
MODULUS (GPa)

0
0.0 WT% CNT 0.1 WT% PCNT 0.1 WT% ACNT

CNT COMPOSITION IN LAMINATES

Figure 6: Effect of CNT composition on the tensile and flexural modulus

28
(a) (b)

(c) (d)

Figure 7: Fracture surface of E-glass fiber/epoxy composites containing: (a) 0.0% CNT; (b) 0.1wt% ACNT; (c)

0.1wt% PCNT; (d) 0.1 wt% ACNT (69% resin)

29
Tables:

Table 1: Summary of mechanical properties obtained

Tensile Tensile Flexural Flexural

S.No. Specimen Modulus Strength Modulus Strength

(GPa) (MPa) (GPa) (MPa)

1 Neat epoxy (76% resin) 6.38 268 7.27 131

2a PCNT/Epoxy/GF (76% resin) 8.03 380 10.9 229

2b ACNT/Epoxy/GF (76% resin) 5.48 227 9.90 191

3a PCNT/Epoxy/GF (69% resin) 6.69 248 11.03 224

3b ACNT/Epoxy/GF (69% resin) 8.89 369 10.40 255

30

You might also like