You are on page 1of 28

Accepted Manuscript

Experimental fracture study of MWCNT/epoxy nanocomposites under the combined


out-of-plane shear and tensile loading

Behnam Saboori, Majid R. Ayatollahi

PII: S0142-9418(16)31103-5
DOI: 10.1016/j.polymertesting.2017.01.028
Reference: POTE 4914

To appear in: Polymer Testing

Received Date: 18 October 2016


Revised Date: 30 January 2017
Accepted Date: 31 January 2017

Please cite this article as: B. Saboori, M.R. Ayatollahi, Experimental fracture study of MWCNT/epoxy
nanocomposites under the combined out-of-plane shear and tensile loading, Polymer Testing (2017),
doi: 10.1016/j.polymertesting.2017.01.028.

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.
ACCEPTED MANUSCRIPT
Experimental fracture study of MWCNT/epoxy nanocomposites under the

combined out-of-plane shear and tensile loading

Behnam Saboori and Majid R. Ayatollahi1

Fatigue and Fracture Research Laboratory, Center of Excellence in Experimental Solid Mechanics and

PT
Dynamics, School of Mechanical Engineering, Iran University of Science and Technology, Narmak, 16846

Tehran, Iran

RI
SC
Abstract

U
In order to explore the role of multi-walled carbon nanotubes (MWCNTs) on the fracture
AN
behavior of epoxy-based nanocomposites, fracture tests were conducted under the combined

out-of-plane shear and tensile loading. Epoxy resin LY-5052 together with MWCNT contents
M

of 0.1, 0.5 and 1.0 wt.% were used to produce nanocomposite specimens. The results showed
D

that increasing the contribution of out-of-plane shear from pure mode I towards pure mode III
TE

enhanced fracture toughness for both pure epoxy and nanocomposites. Additionally, it was

found that in both loading conditions of pure mode III and mixed mode I/III, increasing
EP

MWCNT content up to 1.0 wt.% enhanced fracture toughness with an ascending trend. The

mechanisms involved in the fracture behavior of polymer-based nanocomposites were also


C

studied in detail using the photographs taken from the fracture surfaces by scanning electron
AC

microscopy.

Keywords: Polymer-based Nanocomposite; Multi-walled carbon nanotube; Mixed mode I/III

loading; Fracture toughness; Fracture mechanisms

1
Corresponding author. Tel.: +98 21 77240201; fax: +98 21 77240488.

E-mail address: m.ayat@iust.ac.ir (M.R. Ayatollahi).

1
ACCEPTED MANUSCRIPT

1. Introduction

Epoxy-based composites are widely used in various industrial applications. Due to their

excellent stiffness and strength as well as their extraordinary electrical and thermal properties,

PT
carbon nanotubes (CNTs) have been suggested as an ideal reinforcement material for the

epoxy-based composites in recent years. Nonetheless, the epoxy/CNT composites (often

RI
called nanocomposites) are susceptible to brittle fracture specially in the presence of pre-

SC
existing cracks.

A large number of studies have been dedicated to investigate the fracture behavior of

U
nanocomposite materials. Therefore, most of the previous experimental researches dealing
AN
with fracture in epoxy-based nanocomposites are limited to pure mode I loading cases. The
M

methods of dispersing carbon nanotubes in resin and their effects on the fracture resistance of

epoxy/CNT nanocomposites have been the subject of interest in several studies [1-4]. While
D

the calendaring technique has been found an effective homogenization method in these
TE

papers, some other researchers were successful in dispersing CNTs within the resin properly

using the ultrasonic wave technique [5-7]. In addition, the influence of CNT functionalization
EP

on the mode I fracture toughness (KIc) of nanocomposite materials has been assessed in some
C

papers such as [1, 8, 9]. The effect of CNT types (single-, double- and multi-walled) on KIc
AC

has also been evaluated experimentally (see for instance [10]). Numerous other researchers

have investigated the variation of KIc with CNT content percentage [11-14]. In comparison

with the mode I fracture, there are very few experimental studies conducted in the case of

mixed-mode I/II and mode II fracture of epoxy/CNT nanocomposites. Carried out in recent

years, these studies have investigated the influences of CNT content [15], CNT aspect ratio

[16], and loading conditions [15, 16] on the mixed mode I/II fracture resistance and electrical

2
ACCEPTED MANUSCRIPT
properties of epoxy-based nanocomposites. Also reported in these researches are the micro-

mechanisms involved in the material fracture energy, based on the microscopic examination

of the fracture surfaces. Additionally, some other researchers have studied the mixed mode

I/II fracture resistance of the epoxy-based nanocomposites reinforced with three different

types of carbon nanoparticles [17-19]. The nanofillers added to epoxy resin were the

PT
nanodiamond (ND) of spherical shape, the graphene oxide (GO) nanoplatelets and the carbon

RI
nanofiber (CNF) of cylindrical shape.

Crack extension usually initiates from the defects that are directed at arbitrary angles relative

SC
to the loading direction. Therefore, cracked nanocomposites are sometimes subjected to the

U
out-of-plane shear (mode III) loading. The fracture behavior of CNT-filled nanocomposites in
AN
the loading modes that include out-of-plane shear has not been assessed yet. In view of the

gap observed in the literature, this paper deals with examining the influence of carbon
M

nanotubes on the fracture behavior of the epoxy-based nanocomposites subjected to combined

“out-of-plane shear” and “tensile” loading. The experimental program was performed via a
D

series of fracture tests in pure mode III, pure mode I and mixed mode I/III loading conditions
TE

on neat epoxy and MWCNT/epoxy nanocomposites. For exploring the effect of nanoparticle
EP

content on the fracture resistance, different test samples were manufactured using three

MWCNT weight percentages of 0.1, 0.5 and 1.0. Finally, the scanning electron microscopy
C

(SEM) was used to survey the fracture surfaces with the goal of studying the fracture
AC

mechanisms participating in mixed mode I/III fracture of the nanocomposites.

2. Mixed mode I/III loading apparatus

In this research, mixed mode I/III fracture tests were performed using a new loading

configuration recently proposed by the authors [20]. The geometry of loading fixture and its

test specimen designed for mixed mode I/III fracture tests can be seen in Fig. 1. A brief

3
ACCEPTED MANUSCRIPT
description of the test set up is presented in this section. As shown in Fig. 2, the test specimen

is a plate of rectangular shape with length L = 88 mm, width W = 20 mm, thickness about t =

8.5 mm and an edge crack of length about a = 10 mm.

PT
RI
U SC
AN
(a) (b)
M

Fig. 1. Mixed mode I/III loading configuration.


D
TE
C EP
AC

Fig. 2. The test specimen with an edge crack (dimensions in mm).

The loading fixture comprises of two identical portions. Each half of the loading fixture

contains five loading holes as shown in Fig. 1. In order to provide the needed rigidity to

4
ACCEPTED MANUSCRIPT
preclude unwanted deformation around the loading holes, the fixture was designed relatively

thick and manufactured from a high strength steel. The angle between the loading direction

and the crack plane can be changed by applying the loads at suitable loading holes on the

fixture. The positions of loading holes on each half of the fixture were determined in a way to

achieve an appropriate distribution of mixed modes between mode I and mode III. Depending

PT
on the magnitude of the loading angle α defined in Fig. 1, pure mode I ( = 0°), pure mode

III ( = 90°) and three mixed mode loading cases ( = 40°, 65° and 78°) are provided. As

RI
shown in Fig. 1, the test specimen is fixed to the loading fixture by bolt and nut through two

SC
holes of 10 mm diameter.

U
In order to assess the new test configuration and to study its mixed mode deformation, 3D
AN
finite element (FE) model of the test configuration (including the test sample and the loading

fixture) were analyzed for different loading conditions [20]. The FE analyses confirmed the
M

efficiency of test configuration by demonstrating that the loading fixture could suitably

provide pure mode I, pure mode III and some mixed mode I/III loading cases with a favorite
D

distribution in between.
TE

3. Materials and specimen preparation


EP

The epoxy resin Araldite LY 5052 was selected as matrix because of its low viscosity and
C

extensive industrial applications. The low viscosity of the matrix facilitates the dispersion of
AC

additives. The curing agent was Aradur 5052 and nanoparticles employed in these

experiments were the functionalized multi-walled carbon nanotubes (MWCNTs). According

to the supplier (Nanostructured and Amorphous Materials Inc. [21]), the MWCNTs had

diameters between 10 and 20 nm, lengths between 10 and 30 µm, and a carbon purity of 95%.

Apart from pure epoxy, three different nanocomposites, containing 0.1, 0.5 and 1.0 wt.% of

CNT were prepared. The nanocomposites were produced by dispersing particles in the epoxy

5
ACCEPTED MANUSCRIPT
using the ultrasonic wave technique. First, the desired amounts of CNTs were dispersed in the

epoxy by mechanical stirring and the mixtures containing 0.1, 0.5 and 1.0 wt.% MWCNTs

were sonicated at 70 W for 36, 43 and 57 min, respectively. During the sonication, the

mixture container was held in cold water to prevent the temperature of the mixture becoming

very high. In addition, during the sonication the mixture was stirred using a small spoon every

PT
10 min to make sure that the sonication energy was applied uniformly to the entire mixture.

RI
Afterwards, the hardener was added and the mixture was degassed in the vacuum

environment and molded. Eventually, on the basis of the resin supplier recommendation [22],

SC
the specimens were cured for 24 hours at the ambient conditions followed by 4 hours at 100

°C. The cured specimens were then drilled to create two holes required for connecting them in

U
the loading fixture. To generate the crack in the fabricated specimens for fracture tests, first, a
AN
primary crack was created using a very thin saw blade. Then, the crack tip was sharpened
M

utilizing a razor blade. Based on the measurement by an optical microscope, the width of

created notches didn’t exceed 0.4 mm in all the specimens. The depth of each pre-crack was
D

also measured by a microscope for determining the total crack length in each specimen (as
TE

given in Table 1).


EP

Fig. 3 illustrates samples of pure epoxy and nanocomposite specimens and Fig. 4 is the

sharpened notch tip of a nanocomposite specimen magnified under the microscope.


C

4. Fracture experiments
AC

To investigate the effect of CNT content on the fracture behavior of epoxy-based

nanocomposites under mixed mode I/III loading, a series of fracture tests were conducted in

pure mode I, pure mode III and also one mode III-dominated mixed mode (corresponding to

= 78°). The fracture tests were carried out under a constant loading rate of 0.2 mm/min

using a uniaxial loading machine. Fig. 5 shows the loading set up for a sample nanocomposite

specimen tested under mode III loading.

6
ACCEPTED MANUSCRIPT
Table 1. Crack lengths, specimen thicknesses and fracture loads of the specimens tested under

different loading modes.

Pure mode I Mixed mode: KIII > KI (α=78°) Pure mode III

Material a (mm) t (mm) Pcr (N) a (mm) t (mm) Pcr (N) a (mm) t (mm) Pcr (N)
Pure

PT
10.0 7.7 437.3 10.4 8.5 949.8 10.5 8.5 1188.2
Epoxy
10.0 7.7 492.9 10.9 8.5 896.7 10.4 7.6 892.7
10.0 7.6 580.0 10.5 8.5 1131.9 10.5 7.8 1285.9

RI
0.1 wt% 10.1 8.5 542.7 10.2 8.4 1036.2 10.2 8.4 1271.6
10.2 8.5 618.1 10.2 8.4 1211.5 9.4 7.8 1139.2

SC
10.4 8.5 653.6 9.8 8.4 1153.4 9.4 8.0 1549.9
0.5 wt% 10.2 8.3 574.8 10.3 8.3 1052.1 10.6 8.5 1480.1
10.0 8.2 560.4 10.3 8.5 1027.7 10.2 8.4 1207.8

U
11.0 8.4 620.7 10.3 8.5 1314.6 9.4 8.0 1357.5
1 wt% 9.8 8.4 584.9 10.3 8.4 1298.1 10.2 8.3 1594.2
AN
10.3 8.3 654.5 10.3 8.5 1309.6 10.4 8.5 1219.2
10.0 8.3 657.2 10.3 8.4 1103.7 10.4 8.5 1612.7
M
D
TE
C EP
AC

Fig. 3. Sample test specimens made from pure Fig. 4. The sharpened notch tip of a nanocomposite
epoxy and MWCNT/epoxy nanocomposite. sample shown under an optical microscope.

7
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 5. The loading set up of a typical nanocomposite specimen within the test machine.

According to ASTM-D5045 standard [23], each fracture test was replicated at least three

U
AN
times. Consequently, a total number of 36 experimental data were obtained from the tests

conducted on both pure epoxy and different nanocomposites containing 0.1, 0.5 and 1.0 wt.%
M

of CNT under three loading modes. The load and displacement data were recorded during the

tests.
D
TE

Fig. 6 and Fig. 7 exhibit typical load–displacement curves obtained for pure epoxy and

nanocomposite specimens fractured under mode I and mode III loading conditions,
EP

respectively. The load–displacement curves were nearly linear up to the fracture load at which

the specimens broke instantly showing the brittle fracture behavior of the tested specimens.
C

Details of all the performed tests including the thicknesses, crack lengths and fracture loads of
AC

the test samples have been presented in Table 1.

8
ACCEPTED MANUSCRIPT

PT
RI
Fig. 6. Load–displacement curves obtained for the Fig. 7. Load–displacement curves obtained for the

SC
pure epoxy and nanocomposite test samples under pure epoxy and nanocomposite test samples under
pure mode I loading. pure mode III loading.

U
AN
5. Results and discussion
M

The fracture load obtained for each specimen was utilized to calculate its critical stress
D

intensity factors (KIf and KIIIf). Since there are no analytical equations to calculate KIf and KIIIf
TE

for the tested samples, fracture parameters were obtained from the FE analyses. As stated

before, details of the created FE model can be found in [20]. The calculated critical stress
EP

intensity factors (SIFs) obtained for all the specimens are listed in Table 2, and the average

values calculated for each specific experiment are depicted in Fig. 8. This figure shows that
C

the fracture resistance under all mode mixities, apart from pure mode I, grows with increasing
AC

MWCNTs content up to 1.0 wt.%. On the other hand, for pure mode I loading, the

nanocomposite containing 0.5 wt.% of fillers has the maximum fracture toughness.

9
ACCEPTED MANUSCRIPT
Table 2. The critical stress intensity factors obtained from the mixed mode I/III fracture tests
conducted on pure epoxy and nanocomposite specimens.

Mixed mode: KIII > KI


Pure mode I Pure mode III
(α=78°)
KIf KIIIf KIf KIIIf KIf KIIIf
Material
( √ ) ( √ ) ( √ ) ( √ ) ( √ ) ( √ )
Pure Epoxy 1.56 0.00 1.35 2.16 0.00 2.91

PT
1.75 0.00 1.33 2.12 0.00 2.52
2.09 0.00 1.80 2.55 0.00 3.53

RI
0.1 wt% 1.77 0.00 1.55 2.32 0.00 3.09
2.08 0.00 2.00 2.68 0.00 2.88
2.29 0.00 1.74 2.48 0.00 3.79

SC
0.5 wt% 1.98 0.00 1.72 2.42 0.00 3.65
1.87 0.00 1.48 2.29 0.00 2.94
2.61 0.00 2.26 2.90 0.00 3.29

U
1 wt% 1.82 0.00 2.30 2.91 0.00 3.91
AN
2.29 0.00 2.20 2.87 0.00 2.96
2.15 0.00 1.74 2.48 0.00 3.89
M

In order to compare the values of fracture resistance for different nanocomposites in various

loading modes, the effective fracture toughness, Keff, with the following definition can be
D

employed.
TE

(1)
= +
C EP
AC

Fig. 8. The average values of critical SIFs obtained for the conducted experiments.

10
ACCEPTED MANUSCRIPT
Using Eq. (1), the results of fracture experiments have been illustrated in Fig. 9 in terms of

effective fracture toughness.

PT
RI
U SC
AN
Fig. 9. The values of effective fracture toughness for pure epoxy and for nanocomposites with
M

different contents of CNT.


D

This figure shows that by increasing the contribution of out-of-plane shear (mode III loading),
TE

the effective fracture toughness of the fractured specimen (Keff,f) grows for both pure epoxy

and nanocomposites. A similar trend has also been observed in the results of mixed mode I/III
EP

fracture tests recently conducted on PMMA [20] where enhancement in fracture toughness

was suggested to be due to larger size of plastic zone around the crack tip. A larger plastic
C

zone dissipates more energy and leads to an increase in fracture toughness.


AC

Some other studies have also demonstrated that plastic deformation near the crack tip plays a

vital role in brittle fracture of polymeric materials [24, 25]. Two main micromechanical

processes related to plastic deformation in the region near the crack tip are shear yielding and

crazing [26, 27]. Shear yielding commonly takes place when the shear stress around the crack

tip attains a critical value and consequently, shear bands are extended from the crack tip in the

11
ACCEPTED MANUSCRIPT
direction that the shear stress is maximum. While crazing denotes the formation of micro

voids in a plane perpendicular to the maximum principle stress, one can find little evidence

for crazing in thermosetting polymers such as epoxy [28]. Thus, shear yielding and shear band

formation can be the major energy dissipative mechanisms in the epoxies. This mechanism

might be an acceptable rationale for the fracture resistance boost both in nanocomposites and

PT
in pure epoxy under mixed mode loading cases. Indeed, increasing the contribution of mode

RI
III in the applied loading increases the shear stresses around the crack tip and as a result,

enhances the energy dissipation due to the shear band formation. Larger energy dissipation

SC
makes lower energy available for crack extension and therefore the material fracture

resistance in the loading modes with higher contributions of mode III is greater. Fig. 10a-c

U
show the SEM (scanning electron microscopy) micrographs of fracture surfaces for the pure
AN
epoxy in various cases of mode I, mode III and mixed mode (KIII > KI). These Figures
M

illustrate well the effect of out-of-plane shear load on the formation of shear bands, such that

by increasing the out-of-plane shear load (i.e. approaching pure mode III loading), the
D

roughness of fracture surfaces becomes denser and deeper representing more energy
TE

dissipation during the fracture process. A comparison between the fracture surfaces of the

nanocomposite containing 0.5 wt.% of MWCNT tested under different loading conditions is
EP

also displayed in Fig. 10d-f. It is observed that the shear bands of the fracture surface for the

specimens filled with CNTs have increased with increasing the out-of-plane shear load.
C
AC

Another point found in Fig. 10 is that, generally, the fracture surface characteristics of

nanocomposites in a specific loading mode are also denser compared to the neat epoxy

specimen in the same loading mode. This difference can be due to the presence of CNTs, as

discussed later.

12
ACCEPTED MANUSCRIPT

PT
RI
(a) (b) (c)

U SC
AN
M

(d) (e) (f)


D

Fig. 10. SEM micrographs of the fracture surface of pure epoxy under: (a) mode I (b) mixed
TE

mode (KIII > KI) (c) mode III, and nanocomposites containing 0.5 wt.% CNT under: (d) mode

I (e) mixed mode (KIII > KI) (f) mode III. All the pictures are taken from a region close to the
EP

initial crack front.


C

It is seen from Fig. 9 that the mode I fracture toughness of nanocomposites enhances as the
AC

content of CNTs is increased. The enhancement percentage of effective fracture toughness of

the nanocomposite specimens with respect to the pure epoxy ones has been depicted in Fig.

11 for different mode mixities. The maximum reinforcement effect of nanoparticles on the

mode I fracture resistance is about 19.5% which belongs to the MWCNT content of 0.5 wt.%.

This is while the addition of 1.0 wt.% of MWCNTs to epoxy resulted in a lower mode I

fracture toughness (KIc) than that of the 0.5 wt.% addition. A similar behavior has been

13
ACCEPTED MANUSCRIPT
reported earlier for the variations of KIc with CNT content [15, 16]. Moreover, Fig. 9 and Fig.

11 indicate that in the mode III and mixed mode I/III loading conditions, the fracture

resistance has an ascending trend with increasing nanotube content and this trend is held even

up to 1.0 wt.%. For instance, the addition of 1.0 wt.% of MWCNTs to epoxy resulted in the

enhancement of mode III fracture toughness (KIIIc) as much as about 20%.

PT
RI
U SC
AN
M
D

Fig. 11. Enhancement in effective fracture toughness of MWCNT/epoxy nanocomposites


TE

relative to pure epoxy, in different combinations of mode I and III.

To identify the reasons for the mentioned behaviors, the fracture mechanisms in the broken
EP

samples were explored by using SEM technique. Fig. 12 illustrates the SEM pictures taken
C

from the pure epoxy and nanocomposites samples broken under mode I loading. As it is seen
AC

in Fig. 12a, the fracture surface of the neat epoxy is smooth and mirror-like, which represents

its brittle fracture behavior. Moreover, the images of Fig. 12 show that the addition of

nanofiller resulted in generating regular shear steps on the fracture surfaces. The increased

surface roughness in higher nanoparticle contents indicates that the crack propagation in the

nanocomposites was opposed by rigid and stiff MWCNTs enforcing local deviation of the

crack from its main plane. These local deviations enhance the fracture surface, which in turn

14
ACCEPTED MANUSCRIPT
requires more energy for the crack extension process and results in an increase in fracture

toughness. Therefore, the crack deviation mechanism can be one of the major factors for

improved fracture toughness of MWCNT/epoxy nanocomposites, as shown in Fig. 9.

However, adding 1.0 wt.% of MWCNTs to epoxy leads to a lower value of KIc relative to the

nanocomposite with 0.5 wt.% MWCNT. Another point which can be observed in Fig. 12d is

PT
that the regularity of surface roughness in the 1.0 wt.%-MWCNT nanocomposite is lost in

RI
some extent which can be due to the agglomeration of nanoparticles. For a more detailed

study, high magnification images were also taken from the fracture surfaces (Fig. 13). This

SC
images reveal that in lower contents of MWCNTs (i.e. 0.1 and 0.5 wt.%), nanofillers were

distributed well (Fig. 13a and Fig. 13b), but as Fig. 13c shows, in specimens with 1.0 wt.%

U
enhancers, MWCNTs tended to be agglomerated more. The presence of these agglomerates
AN
can be a major reason for reduction in KIc in the 1.0 wt.%-MWCNT nanocomposite compared
M

to 0.5 wt.% one. This is because in the agglomerates, the resin does not penetrate into and wet

the MWCNTs accumulation and hence, a weak CNT/matrix interface is formed. The
D

regularity of the fracture surface characteristics in the low magnification images may also
TE

indicate appropriate nanofillers dispersion, while a fracture surface with non-regular

roughness can represent a non-proper CNT distribution.


C EP
AC

15
ACCEPTED MANUSCRIPT

PT
RI
SC
(a) (b)

U
AN
M
D
TE
C EP

(c) (d)
AC

Fig. 12. SEM images of fracture surface for mode I specimens: (a) neat epoxy (b) 0.1 wt.%

MWCNT/epoxy (c) 0.5 wt.% MWCNT/epoxy (d) 1.0 wt.% MWCNT/epoxy. All the pictures

are taken from a region close to the initial crack front.

16
ACCEPTED MANUSCRIPT

PT
RI
(a) (b)

U SC
AN
M

(c)
D

Fig. 13. SEM images of mode I fracture surfaces for MWCNT/epoxy nanocomposites: (a) 0.1
TE

wt.% (b) 0.5 wt.% (c) 1.0 wt.%. All the pictures are taken from a region close to the initial

crack front.
EP

For further exploration of the crack deviation mechanism, the images with higher

magnifications were also captured from the fracture surface of one of the nanocomposite
C

specimens (see Fig. 14). Fig. 14a and Fig. 14b exhibit the presence of many fine cracks within
AC

the matrix polymer which are deviated from their propagation paths when facing with very

stiff nanotubes. In order to prove that the observed nanocracks are nucleated due to the

applied mechanical load, a picture with a similar magnification factor from the fracture

surface of a pure epoxy specimen with no mechanical load applied to has also been provided

17
ACCEPTED MANUSCRIPT
in Fig. 14c. According to this Figure, in the absence of a mechanical load no fine crack could

be observed within the matrix.

PT
RI
SC
(a) (b)

U
AN
M
D
TE

(c)

Fig. 14. SEM micrographs of fracture surfaces for: (a) nanocomposite containing 0.5 wt.%
EP

CNT (b) higher magnification of nanocomposite containing 0.5 wt.% CNT (c) pure epoxy
C

not-subjected to any mechanical load. All the pictures are taken from a region close to the
AC

initial crack front.

Other effective factor in enhancing the fracture resistance of nanocomposites with respect to

pure epoxy is the crack bridging which occurs for the fillers like nanotubes having a high

aspect ratio. When a CNT bridges between two crack planes, it might be pulled out or broken

depending on the interface strength and embedment length [29]. The recognition of filler

fracture from filler pullout is principally based on the lengths of nanotubes left on the fracture

18
ACCEPTED MANUSCRIPT
surface. Fig. 15 displays some high magnification SEM images of fracture surface in the

tested nanocomposites.

PT
RI
U SC
AN
(a) (b)
M
D
TE
C EP
AC

(c) (d)

Fig. 15. SEM micrographs of fracture surfaces of nanocomposite containing: (a) 1.0 wt.%
CNT (b) 1.0 wt.% CNT with higher magnification (c) 0.5 wt.% CNT (d) 0.5 wt.% CNT with
higher magnification. All the pictures are taken from a region close to the initial crack front.

19
ACCEPTED MANUSCRIPT
Both broken and pulled out fillers can be seen on the fracture surface, representing the

bridging mechanism. The likely broken nanotubes appear with a very short length left on the

surface, because their breakage usually takes place near the fracture surface [29].

Additionally, Fig. 15d clearly shows some holes created on a nanocomposite fracture surface

because of the CNTs pulled out from the matrix.

PT
Similar to the case of pure mode I, in the loading cases of mode III and mixed mode I/III, the

RI
mechanisms of local crack deviation and bridging play important roles in the fracture

resistance improvement (depicted in Fig. 9) due to the presence of CNTs. First, it is necessary

SC
to point out the difference between the fracture paths resulted from the mode I loading

U
conditions and from the cases where the out-of-plane shear load is also applied to the
AN
specimens. Fig. 16 shows samples of the specimens broken under pure mode I and pure mode

III loading.
M
D
TE
EP

(a) (b)
C

Fig. 16. Fracture surfaces of nanocomposite specimens broken under: (a) mode I and (b)
AC

mode III loading.

In contrast to the mode I case, the crack subjected to mode III loading does not extend along

its initial plane but its surface deviates in the out-of-plane direction. This is because the

maximum principal stress direction keeps changing as the crack advances. Therefore, at each

step of crack extension, the crack plane rotates towards a new direction of maximum principal

stress. As a result, with increasing the mode III dominancy, the crack propagates in a longer

20
ACCEPTED MANUSCRIPT
path and the fracture surface becomes larger. This can provide more opportunity for the

nanotubes to resist the crack extension and to dissipate more energy which both lead to an

enhancement in fracture toughness.

It is noteworthy that contrary to the early expectation, the addition of 0.1 and 0.5 wt.% of

PT
CNTs into the matrix polymer together with increasing the contribution of out-of-plane shear

load result in a lower augmentation in nanocomposites fracture toughness (see Fig. 11). The

RI
reason can be that in the mixed mode loading conditions with KIII > KI, crack opening is very

small and there is no crack opening under pure mode III loading. Consequently, the crack

SC
bridging mechanism (that plays a significant role in enhancing the fracture resistance under

U
the opening mode) loses its influence with increasing the mode III contribution in the applied
AN
load. Indeed, with increasing the out-of-plane shear load applied to the 0.1 and 0.5 wt.%-

MWCNT nanocomposites, the combined effects from the mechanisms of crack deviation and
M

bridging lead to a lower increase in the fracture resistance with respect to pure epoxy and

compared to the mode I case. According to Fig. 11, in the case of nanocomposite containing
D

1.0 wt.% MWCNTs, the descending trend of fracture toughness enhancement with reference
TE

to pure epoxy still exists by moving from mixed mode I/III towards pure mode III. However,
EP

under mode I loading, the mentioned nanocomposite exhibits minimum fracture toughness

improvement in comparison with two other loading modes. In general, the test results shown
C

in Fig. 11 declare that in the mode I loading case, the nanocomposite reinforced by 0.5 wt.%
AC

of MWCNTs has the maximum fracture resistance in comparison with the other weight

percentages. On the other hand, with dominating the out-of-plane shear load, the increasing

trend of fracture toughness due to increasing CNT content is continuous even up to 1.0 wt.%.

As stated before for 1.0 wt.%-MWCNT nanocomposite, the relatively higher amount of

agglomerations can be responsible for the reduction in the mode I fracture toughness

enhancement. On the other hand, the ascending trend of fracture toughness up to 1.0 wt.% of

21
ACCEPTED MANUSCRIPT
CNTs with the dominancy of mode III loading might suggest that the negative effect of

nanoparticle agglomerations under shear loading is lower than under the tensile loading.

6. Conclusions

In this paper, the fracture resistance of MWCNT/epoxy nanocomposites under mixed mode

PT
I/III loading was investigated experimentally. The experimental program included fracture

tests under pure mode I, pure mode III and mixed mode I/III loading performed using a newly

RI
developed loading configuration. The test specimens were made from both pure epoxy

SC
(Araldite LY 5052) and nanocomposites comprising different contents of functionalized

MWCNT (0.1, 0.5 and 1 wt.%). The MWCNT/epoxy nanocomposites and pure epoxy

U
specimens fractured in a brittle manner for all mode mixities. It was observed that the
AN
effective fracture toughness of both nanocomposites and pure epoxy were improved when

loading condition was shifted from pure mode I towards pure mode III. The obtained results
M

revealed that the increase in the content of carbon nanotubes enhances the mode I fracture
D

toughness of MWCNT/epoxy nanocomposites. The maximum enhancement of KIc was 19.5%


TE

which was obtained for the nanocomposite with 0.5 wt.% of MWCNT. Also in mode III

loading conditions, the nanocomposite fracture resistance was enhanced with increasing
EP

MWCNT content even up to the filler content of 1.0 wt.%. For instance, with addition of 1.0

wt.% of MWCNT to epoxy, mode III fracture toughness grew as much as about 20%. Finally,
C

the mechanisms involved in the observed fracture behavior of nanocomposites were examined
AC

and discussed by using the SEM images of the fracture surfaces.

Funding statement

This research did not receive any specific grant from funding agencies in the public,

commercial, or not-for-profit sectors.

22
ACCEPTED MANUSCRIPT
References

[1] F. Gojny, M. Wichmann, U. Köpke, B. Fiedler, K. Schulte, Carbon nanotube-reinforced


epoxy-composites: enhanced stiffness and fracture toughness at low nanotube content,
Composites Science and Technology, 64 (2004) 2363-2371.
[2] R. Hollertz, S. Chatterjee, H. Gutmann, T. Geiger, F. Nüesch, B. Chu, Improvement of

PT
toughness and electrical properties of epoxy composites with carbon nanotubes prepared
by industrially relevant processes, Nanotechnology, 22 (2011) 125702.
[3] E.T. Thostenson, T.W. Chou, Processing-structure-multi-functional property relationship

RI
in carbon nanotube/epoxy composites, Carbon, 44 (2006) 3022-3029.
[4] H. Hedia, L. Allie, S. Ganguli, H. Aglan, The influence of nanoadhesives on the tensile

SC
properties and Mode-I fracture toughness of bonded joints, Engineering Fracture
Mechanics, 73 (2006) 1826-1832.

U
[5] H. Miyagawa, A.K. Mohanty, L.T. Drzal, M. Misra, Nanocomposites from biobased
AN
epoxy and single-wall carbon nanotubes: synthesis, and mechanical and thermophysical
properties evaluation, Nanotechnology, 16 (2005) 118-124.
[6] N. Yu, Z. Zhang, S. He, Fracture toughness and fatigue life of MWCNT/epoxy
M

composites, Materials Science and Engineering: A, 494 (2008) 380-384.


[7] X.H. Zhang, Z.H. Zhang, W.J. Xu, F.C. Chen, J.R. Deng, X. Deng, Toughening of
D

cycloaliphatic epoxy resin by multiwalled carbon nanotubes, Journal of Applied Polymer


TE

Science, 110 (2008) 1351-1357.


[8] P.C. Ma, J.K. Kim, B.Z. Tang, Effects of silane functionalization on the properties of
EP

carbon nanotube/epoxy nanocomposites, Composites Science and Technology, 67 (2007)


2965-2972.
[9] S.J. Park, H.J. Jeong, C. Nah, A study of oxyfluorination of multi-walled carbon
C

nanotubes on mechanical interfacial properties of epoxy matrix nanocomposites,


AC

Materials Science and Engineering: A, 385 (2004) 13-16.


[10] F.H. Gojny, M.H.G. Wichmann, B. Fiedler, K. Schulte, Influence of different carbon
nanotubes on the mechanical properties of epoxy matrix composites–A comparative
study, Composites Science and Technology, 65 (2005) 2300-2313.
[11] S. Ganguli, M. Bhuyan, L. Allie, H. Aglan, Effect of multi-walled carbon nanotube
reinforcement on the fracture behavior of a tetrafunctional epoxy, J Mater Sci, 40 (2005)
3593-3595.

23
ACCEPTED MANUSCRIPT
[12] O. Gryshchuk, J. Karger-Kocsis, R. Thomann, Z. Kónya, I. Kiricsi, Multiwall carbon
nanotube modified vinylester and vinylester–based hybrid resins, Composites Part A:
Applied Science and Manufacturing, 37 (2006) 1252-1259.
[13] L. Sun, G. Warren, J. O’reilly, W. Everett, S. Lee, D. Davis, D. Lagoudas, H.J. Sue,
Mechanical properties of surface-functionalized SWCNT/epoxy composites, Carbon, 46
(2008) 320-328.

PT
[14] Y. Zhou, F. Pervin, L. Lewis, S. Jeelani, Fabrication and characterization of
carbon/epoxy composites mixed with multi-walled carbon nanotubes, Materials Science

RI
and Engineering: A, 475 (2008) 157-165.
[15] M. Ayatollahi, S. Shadlou, M. Shokrieh, Mixed mode brittle fracture in epoxy/multi-

SC
walled carbon nanotube nanocomposites, Engineering Fracture Mechanics, 78 (2011)
2620-2632.
[16] M.R. Ayatollahi, S. Shadlou, M.M. Shokrieh, Effect of multi-walled carbon nanotube

U
aspect ratio on mechanical and electrical properties of epoxy-based nanocomposites,
AN
Polymer Testing, 30 (2011) 548-556.
[17] M. Ayatollahi, E. Alishahi, S. Doagou-R, S. Shadlou, Tribological and mechanical
M

properties of low content nanodiamond/epoxy nanocomposites, Composites Part B:


Engineering, (2012).
D

[18] M. Ayatollahi, E. Alishahi, S. Shadlou, Mechanical Behavior of Nanodiamond/Epoxy


Nanocomposites, International Journal of Fracture, 170 (2011) 95-100.
TE

[19] S. Shadlou, E. Alishahi, M.R. Ayatollahi, Fracture behavior of epoxy nanocomposites


reinforced with different carbon nano-reinforcements, Composite Structures, 95 (2013)
EP

577-581.
[20] M.R. Ayatollahi, B. Saboori, A new fixture for fracture tests under mixed mode I/III
loading, European Journal of Mechanics - A/Solids, 51 (2015) 67-76.
C

[21] www.nanoamor.com/carbon_nanotubes_nanofibers_graphene.
AC

[22] www.maxepoxi.com.br/pdf/araldite_ly5052_aradur_5052_english_2007.pdf.
[23] ASTM, ASTM D5045–99, Standard Test Method for Plane-Strain Fracture Toughness
and Strain Energy Release Rate of Plastic Materials, American Society for Tensting and
Materials, Philadelphia, 1999.
[24] W. Araki, K. Nemoto, T. Adachi, A. Yamaji, Fracture toughness for mixed mode I/II of
epoxy resin, Acta Materialia, 53 (2005) 869-875.
[25] E. White, Fracture behaviour of polymers, in: A. Kinloch, R. Young (Eds.), Applied
Science Publishers, London and New York, 1983.

24
ACCEPTED MANUSCRIPT
[26] A.J. Kinloch, Mechanics and mechanisms of fracture of thermosetting epoxy polymers,
Epoxy Resins and Composites I, Springer Berlin Heidelberg1985, pp. 45-67.
[27] C. Liu, Y. Huang, M. Stout, Enhanced mode-II fracture toughness of an epoxy resin due
to shear banding, Acta Materialia, 46 (1998) 5647-5661.
[28] J. Lilley, D. Holloway, Crazing in epoxy resins, Philosophical Magazine, 28 (1973) 215-
220.

PT
[29] D. Qian, E.C. Dickey, R. Andrews, T. Rantell, Load transfer and deformation
mechanisms in carbon nanotube-polystyrene composites, Applied Physics Letters, 76

RI
(2000) 2868-2870.

U SC
AN
M
D
TE
C EP
AC

25
ACCEPTED MANUSCRIPT
Figure captions

Fig. 1. Mixed mode I/III loading configuration.

Fig. 2. The test specimen with an edge crack (dimensions in mm).

Fig. 3. Sample test specimens made from pure epoxy and MWCNT/epoxy nanocomposite.

PT
Fig. 4. The sharpened notch tip of a nanocomposite sample shown under an optical
microscope.

RI
Fig. 5. The loading set up of a typical nanocomposite specimen within the test machine.

SC
Fig. 6. Load–displacement curves obtained for the pure epoxy and nanocomposite test
samples under pure mode I loading.

U
Fig. 7. Load–displacement curves obtained for the pure epoxy and nanocomposite test
AN
samples under pure mode III loading.

Fig. 8. The average values of critical SIFs obtained for the conducted experiments.
M

Fig. 9. The values of effective fracture toughness for pure epoxy and for nanocomposites with
different contents of CNT.
D

Fig. 10. SEM micrographs of the fracture surface of pure epoxy under: (a) mode I (b) mixed
TE

mode (KIII > KI) (c) mode III, and nanocomposites containing 0.5 wt.% CNT under:
(d) mode I (e) mixed mode (KIII > KI) (f) mode III. All the pictures are taken from a
EP

region close to the initial crack front.

Fig. 11. Enhancement in effective fracture toughness of MWCNT/epoxy nanocomposites


C

relative to pure epoxy, in different combinations of mode I and III.


AC

Fig. 12. SEM images of fracture surface for mode I specimens: (a) neat epoxy (b) 0.1 wt.%
MWCNT/epoxy (c) 0.5 wt.% MWCNT/epoxy (d) 1.0 wt.% MWCNT/epoxy. All the
pictures are taken from a region close to the initial crack front.

Fig. 13. SEM images of mode I fracture surfaces for MWCNT/epoxy nanocomposites: (a) 0.1
wt.% (b) 0.5 wt.% (c) 1.0 wt.%. All the pictures are taken from a region close to the
initial crack front.

26
ACCEPTED MANUSCRIPT
Fig. 14. SEM micrographs of fracture surfaces for: (a) nanocomposite containing 0.5 wt.%
CNT (b) higher magnification of nanocomposite containing 0.5 wt.% CNT (c) pure
epoxy not-subjected to any mechanical load. All the pictures are taken from a region
close to the initial crack front.

Fig. 15. SEM micrographs of fracture surfaces of nanocomposite containing: (a) 1.0 wt.%
CNT (b) 1.0 wt.% CNT with higher magnification (c) 0.5 wt.% CNT (d) 0.5 wt.%

PT
CNT with higher magnification. All the pictures are taken from a region close to the
initial crack front.

RI
Fig. 16. Fracture surfaces of nanocomposite specimens broken under: (a) mode I (b) mode III

SC
loading.

U
AN
M
D

Table captions
TE

Table 1. Crack lengths, specimen thicknesses and fracture loads of the specimens tested under
EP

different loading modes.

Table 2. The critical stress intensity factors obtained from the mixed mode I/III fracture tests
C

conducted on pure epoxy and nanocomposite specimens.


AC

27

You might also like