Sunteți pe pagina 1din 24

The Journal of Adhesion

ISSN: 0021-8464 (Print) 1545-5823 (Online) Journal homepage: https://www.tandfonline.com/loi/gadh20

Effect of nanoalumina in epoxy adhesive on


lap shear strength and fracture toughness of
aluminium joints

Sunil Kumar Gupta, Dharmendra Kumar Shukla & Dhake Kaustubh Ravindra

To cite this article: Sunil Kumar Gupta, Dharmendra Kumar Shukla & Dhake Kaustubh Ravindra
(2019): Effect of nanoalumina in epoxy adhesive on lap shear strength and fracture toughness of
aluminium joints, The Journal of Adhesion, DOI: 10.1080/00218464.2019.1641088

To link to this article: https://doi.org/10.1080/00218464.2019.1641088

Published online: 14 Jul 2019.

Submit your article to this journal

View Crossmark data

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


https://www.tandfonline.com/action/journalInformation?journalCode=gadh20
THE JOURNAL OF ADHESION
https://doi.org/10.1080/00218464.2019.1641088

Effect of nanoalumina in epoxy adhesive on lap shear


strength and fracture toughness of aluminium joints
Sunil Kumar Gupta , Dharmendra Kumar Shukla, and Dhake Kaustubh Ravindra
Department of Mechanical Engineering, Motilal Nehru National Institute of Technology Allahabad, India

ABSTRACT ARTICLE HISTORY


Lap shear strength and opening mode fracture toughness of Received 9 March 2019
Accepted 4 July 2019
epoxy/alumina nanocomposite adhesives in bonded alumi-
nium alloy joints were determined using a single lap joint, KEYWORDS
and double cantilever beam and contoured double cantilever Adhesives with
beam joints, respectively. Epoxy/alumina nanocomposite adhe- nanoparticles; epoxy/
sives were prepared with 0.5, 1.0, 1.5 and 2.0 wt.% of alumina epoxides; fracture
nanospheres (diameter 23–47 nm) and alumina nanorods (dia- mechanics; lap-shear;
meter in the range of 10 nm and length less than 50 nm). alumina nanoparticles
A significant improvement in the lap shear strength and frac-
ture toughness of nanocomposite adhesives was observed
over that of neat epoxy adhesive. Maximum lap shear strength
of joints bonded with nanocomposite adhesives was observed
at 1.5 wt.% of nanospheres and at 1.0 wt.% of nanorods.
Maximum fracture toughness of both types of joint was
observed for adhesive having 1.5 wt.% of nanospheres and
for 1.0 wt.% of nanorods. However, the fracture toughness of
joints having 1.5 wt.% of nanospheres was higher in compar-
ison to the fracture toughness of joints having 1.0 wt.% of
nanorods. Lap shear strength and fracture toughness
decreased on further increment in the wt.% of nanoalumina i.
e. at 2.0 wt.% of nanospheres and, at 1.5 and 2.0 wt.% of
nanorods.

1. Introduction
In many applications, mechanical properties of adhesives are required to be
enhanced for the replacement of conventional joining methods with adhesive
bonding. Epoxy, in its pure state, often undergoes catastrophic failure due to
noticeable brittleness and low fracture toughness. In recent years, extensive
study has been conducted on epoxy adhesive to modify its mechanical and
other properties. Reinforcement of epoxy with nanoparticles is one of the
processes to enhance the properties of adhesives.[1–16]
Patel et al.[1] prepared the nanocomposite adhesive from co-polymeric and
terpolymeric acrylic rubbers and silica or nanoclay. Lap shear tests of alumi-
nium – aluminium single lap joints indicated improvement (126–219%) in

CONTACT Sunil Kumar Gupta sunilg@mnnit.ac.in Department of Mechanical Engineering, Motilal Nehru
National Institute of Technology Allahabad, Prayagraj 211004, India
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/GADH.
© 2019 Taylor & Francis
2 S. K. GUPTA ET AL.

joint strength with nanocomposite adhesives compared to neat acrylic adhe-


sive. Change in lap shear strength of epoxy adhesive as a function of nano-
Al2O3 (average size of about 80 nm) inclusion was determined by Tabaei
et al.[2] Lap shear strength of single lap joints was observed to be maximum
at 1.5 wt.% (weight percentage) of nano-Al2O3 and the strength decreased on
further inclusion of nano-Al2O3 in adhesive. The pull off strength of adhesive
bonded steel joints was investigated as a function of wt.% of alumina
nanoparticles (average diameter of 80 nm) in epoxy adhesive by Zhai et al.[3]
Results showed that the adhesion strength of epoxy adhesive reinforced with
2 wt.% of nano-Al2O3 was about four times higher than that of neat epoxy
adhesive. Adhesion strength increased due to change of failure mode from
interfacial to mixed (cohesive and interfacial). Bhowmik et al.[4] observed
significant increase (300%) in the joint strength when the plasma–nitrided
titanium adherends were bonded with epoxy adhesive reinforced with silicate
nanopowder of 50 nm of size (10 wt.%) over that of epoxy adhesive. Lap
shear and tensile strength of joints with similar adherends of aluminium alloy
(Al2024-T3) and mild steel bonded with epoxy/alumina (10–20 nm) nano-
composite adhesive were investigated by May et al.[5] It was observed that the
lap shear strength of adhesive having alumina nanofillers improved
(140–150%) in comparison to the unmodified epoxy adhesive. Lap shear
and tensile strength of joints with mild steel adherends were comparatively
higher than that of joints with aluminium adherends. Tutunchi et al.[6]
studied the shear and tensile (butt) strength of steel-glass/epoxy composite
joints bonded with two-part structural acrylic adhesive reinforced with Al2O3
nanoparticles (diameter of 20 nm) at different wt.%. Shear and tensile
strength of joints increased with addition of filler content up to 1.5 wt.%
and decreased when more nanoparticles were added. Failure behaviour of
single lap joints prepared with adhesive reinforced with waste composite
particles (size 80–300 µm) and unidirectional fibre glass/epoxy composite
plate substrates was analysed by Turan and Pekbey.[7] The failure load of
reinforced adhesive bonded single lap joint was 1.3–22.8% higher than that of
the unreinforced adhesive. Nassar et al.[8] investigated the effect of reinfor-
cing materials (alumina versus silica), particle size (20 nm versus 80 nm) and
concentration (2.5 wt.% versus 5.0 wt.%) of nanoparticles on the load transfer
capacity of magnesium-steel single lap joint. Results showed that load trans-
fer capacity of joint was higher for adhesive with silica nanoparticles com-
pared to that of adhesives having alumina nanoparticles. But load transfer
capacity decreased with the increase in the concentration of nanoparticles.
Shear strength of adhesives having smaller size nanoparticles (20 nm) was
higher compared to that of adhesives having nanoparticles of size 80 nm.
From these works, it could be concluded that the reinforcement of nanopar-
ticles in the adhesive is advantageous. However, improvement in properties
THE JOURNAL OF ADHESION 3

of adhesive is optimum at a given value of weight percentage of


reinforcement.
Determination of fracture toughness is important to predict the behaviour
of joints for better design, increased performance, reduced cost and enhanced
safety. To improve the fracture toughness or crack resistance of neat epoxy,
inorganic fillers have been used. Imanaka et al.[9] investigated the effect of
size of silica particles (range, 6–30 μm) and content on fracture toughness of
adhesive using double cantilever beam specimen. Results showed that frac-
ture toughness increased with inclusion of bigger size of particles. The effect
of particle size on the fracture toughness of adhesive significantly strength-
ened with increasing content of particles. Buchman et al.[10] studied the effect
of nanoparticles, nano-IF-WS2 (Inorganic Fullerene-like tungsten disul-
phide) and functionalized nano-POSS (Polyhedral-Oligomeric-Sil-
Sesquioxane), on the toughening of low and high temperature curing epoxy
systems. Fracture toughness of both the epoxy system having very low
concentrations (0.3–0.5 wt.%) of nano IF-WS2 was increased by more than
ten fold compared to neat epoxy. At higher concentrations (3.0 and 4.0 wt.
%), epoxy functionalized POSS imparted higher improvement (485% and
380%) in the toughness of epoxy adhesive. Improvement in fracture tough-
ness of adhesive using micro (3.2 ± 0.6 μm in diameter) and nano
(22.3 ± 3.1 nm in diameter) silica particles was reported by Meng et al.[11]
A significant enhancement of 738% in fracture toughness of adhesive was
achieved at 4.4 vol.% of nano-silica. Whereas, the fracture toughness of
adhesive having microparticles (4.4 wt.%) was improved by 60% compared
to neat adhesive. Abenojar et al.[12] fabricated composites of epoxy having 6
and 12% of SiC micro and nanoparticles and studied various mechanical
properties. Addition of SiC nanoparticles increased the bending strength and
hardness compared to that of neat epoxy and epoxy reinforced with micro-
particles. Khoramishad and Khakzad[13] studied fracture energy of epoxy
adhesive reinforced with 0.1, 0.3 and 0.5 wt.% of multi-walled carbon nano-
tubes (MWCNTs) using double cantilever beam of aluminium 6061-T6. The
fracture energy of adhesive having 0.3 wt% of MWCNTs was improved by
58.4% compared to that of neat adhesive. Nemati Giv et al.[14] reviewed the
effect of reinforcements size, content and interface reinforcement/matrix on
mechanical properties of epoxy adhesives. Mechanical properties of epoxy
adhesive and adhesive joints could be enhanced by reinforcing the suitable
nanoparticles.
In the present article, lap shear strength and opening mode fracture
toughness of epoxy/alumina nanocomposite adhesive bonded joints were
analysed experimentally. Epoxy was modified by reinforcing with 0.5, 1.0,
1.5 and 2.0 wt.% of nanoalumina. Two different shapes (spherical and rod) of
alumina nanoparticles were used to study the effect of shape of nanoparticles
on shear strength and fracture toughness of adhesive. The epoxy/alumina
4 S. K. GUPTA ET AL.

nanocomposite was prepared using in situ polymerisation technique. Single


lap joint was used for analysis of shear strength of adhesives. Whereas, for
analysis of fracture toughness of nanocomposite adhesive, two different
configurations of joint (double cantilever beam (DCB) and Contoured dou-
ble cantilever beam (CDCB) were used.

2. Experiments
2.1. Materials
An unmodified liquid resin, Diglycidyl ether of bisphenol A (DGEBA),
Araldite LY556® having density, 1.15–1.20 g/cm3 at 25°C, and an aliphatic
primary amine hardener, HY951® having density, 0.97– 0.99 g/cm3 at 25°C,
were used to prepare the neat adhesive, which were supplied by Huntsman
International (India) Pvt. Ltd., Mumbai, India. Aluminium alloy IS 24345
(Young’s modulus, 71 GPa, Poisson ratio, 0.35) was used as adherend
material provided by Hindalco Industries Limited, Noida, India. Spherical
and rod shaped alumina nanoparticles, called as nanospheres and nanorods,
respectively, were used as reinforcement. Alumina nanospheres were sup-
plied by Nanostructured & Amorphous Materials, Inc., Katy, USA. The
diameter, specific surface area and density of alumina nanospheres are 23–-
47 nm, 35 m2/g and 3.5–3.9 g/cm3, respectively. Alumina nanorods were
procured from Sigma-Aldrich Co, Bangalore, India. Diameter and length of
the alumina nanorods are in the range of 10 nm and less than 50 nm,
respectively. Specific surface area and density of alumina nanorods are
40 m2/g and 4 g/cm3, respectively.

2.2. Specimen geometry


The single lap joints were prepared by joining the two flat aluminium alloy
adherends with epoxy/alumina nanocomposite adhesives. The flat adherends
for lap shear test were fabricated following ASTM Standard D1002[17] as
shown in Figure 1. Length, thickness and width of specimen were 101.6, 1.62

Figure 1. Geometry of flat tensile (FT) joint (dimensions are in mm).


THE JOURNAL OF ADHESION 5

and 25.4 mm, respectively. Smooth surface and uniform thickness of adher-
ends were maintained throughout the area. Overlap length of adhesive
bonded surface of adherends was 12.7 mm.
To evaluate opening mode fracture toughness (GIC), DCB (Figure 2) and
CDCB specimens (Figure 3) were used. DCB and CDCB specimens were
fabricated following ASTM standard D3433.[18] DCB joints with flat adher-
ends are most commonly used specimen to determine fracture toughness as

Figure 2. Geometry of DCB specimen (dimensions are in mm).

Figure 3. Geometry of CDCB specimen (dimensions are in mm).


6 S. K. GUPTA ET AL.

it is easier to fabricate and test these joints. The main drawback of the DCB
test is the requirement of measurement of crack length during the test. Due
to this there is high amount of scatter in data in fracture toughness obtained
from DCB test. However, a large variety of data reduction methods are also
available to analyse the results obtained from DCB test. Whereas, in case of
CDCB test, it is not required to measure the crack length even for classical
formulation. This results in consistent value of fracture toughness of adhesive
along the crack length. DCB specimen consists of two flat adherends with
dimensions 356 × 25.4 × 12.7 mm3 (see Figure 2). A hole of diameter
6.35 mm was drilled at a distance of 25.4 mm from one end to apply the
tensile load through pin and a holding device. The dimensions of CDCB joint
are shown in Figure 3. Contoured (tapered) adherends provide variation in
the thickness of specimen along the length to develop a linear compliance
specimen with the growth of crack. Its height is varied so that the slope, m
3a2
 −1 −1
h3 þ h remains constant (m = 90 in = 3.54 mm ) throughout the overlap
1

length of joints while performing the test.[18]

2.3. Preparation of joints


Epoxy/alumina nanocomposite adhesives were synthesised and prepared
through In-situ polymerization technique. According to wt.% of alumina
nanoparticles, required amount of dried nanoparticles were manually
mixed with acetone. This mixture was sonicated (at 50 Hz frequency and
230 W power rating) for an hour. Then, required amount of epoxy was
added in the sonicated mixture. This mixture was sonicated for an hour for
proper dispersion of nanoparticles in the epoxy matrix. After sonication,
partial distillation process was used to remove the acetone from mixture.
Nanocomposite adhesive was synthesized with 0.5, 1.0, 1.5 and 2.0 wt.% of
the alumina nanoparticles for both types of nanoparticles.
Intensive care was taken for the preparation of bonding surfaces and
alignment of bonding adherends of joints. Steps followed for preparation of
joints are forth mentioned. Adherends were properly cleaned and degreased
using acetone to remove traces of previous applied adhesive if any. After
proper cleaning and drying of adherends, the bonded length was measured
and marked on adherends. The surfaces of bonded area of adherends were
abraded with abrasive paper (grit no. #60). Then adherends were again
cleaned by acetone to remove particles from bonded surface produced during
surface abrasion. The adherends were numbered with letter punch and were
paired together.
The epoxy adhesive was prepared by blending the resin and hardener at
the recommended ratio (10:1 by weight). Then required amount (weight) of
adhesive for an adhesive thickness of 100 μm was applied on surfaces to be
THE JOURNAL OF ADHESION 7

bonded. At the time of application of adhesive on adherends, special care was


taken to avoid entrapping of air on the rough surface and the shortage of
adhesive on the bonding area. FT joints were kept in a fixture for maintain-
ing the alignment of adherends. It was ascertained that loading surfaces were
parallel to each other. In the case of fracture toughness test specimens, non-
adhering shims having a thickness of 100 µm was used at the left end of joint
to create an initial crack of length of 51 and 48.62 mm for DCB and CDCB
joints, respectively.
The applied adhesive in the bonded joint was cured at room temperature
at least for 24 h. After curing, the spilled out adhesive at the edges of the joint
was removed with the help of a blade. Net weight of adhesive in the joint was
calculated by subtracting the weight of adherends from the weight of the
joint. Thickness of adhesive layer was determined by dividing the net weight
of adhesive with product of density of adhesive and bonded area.

2.4. Characterisation
2.4.1. Surface roughness test
The Surtronic 25 surface checker (Taylor Hobson Precision, India), having
diamond stylus tip of radius 5 μm and a resolution of 0.01 μm, was used to
measure the surface roughness of adherends after making the abraded pat-
tern. The traverse speed of stylus was set at 1 mm/s. Surface roughness was
defined in term of arithmetic mean deviation (Ra). Average roughness of the
abraded surface of adherends was calculated along with standard deviation.

2.4.2. Lap shear strength test


Lap shear strength test was performed following ASTM standard D1002 on
servohydraulic universal testing machine (Nano Plug n Play, supplied by
Bangaluru Integrated System Solutions (BISS) Bangaluru, India). Tests were
performed with a 25 kN load cell and cross head speed of 1.27 mm/min.
Failure load and corresponding displacement were recorded for each sample.
Lap shear strength was calculated by dividing the maximum load taken by
the joint with bonded overlap area. At least five tests were conducted for each
type of adhesive joint. The average shear strength of adhesive at 95% con-
fidence level was calculated.

2.4.3. Fracture toughness test


Fracture toughness tests were performed following ASTM standard D3433
on an universal testing machine (Tinius Olsen, India) with a load cell of 2
kN. Cross head speed for DCB and CDCB test were 0.5 and 0.2 mm/min,
respectively. Bonded specimens (DCB and CDCB) were tested such that the
crack was made to extend by tensile force acting in the normal direction to
the crack surface. Load was applied through a holding device, which was
8 S. K. GUPTA ET AL.

attached with the specimen by the help of two pins inserted in drilled hole at
the ends of upper and lower adherend. Drop in the load was observed at the
start of crack growth. Loading was stopped to arrest the crack growth and
crack length was marked. After marking the crack length, the joint was
unloaded. The cycle of loading and unloading was repeated till the crack
extended to the full length of the joint. Loads at the different crack length of
DCB and CDCB joints were recorded. Linear elastic behaviour of adhesive
was assumed for calculation of fracture toughness. For DCB specimen, the
maximum load for each crack length was recorded to calculate the crack
initiation fracture toughness (GIC) from each arrested points as the crack
started to re-grow using equation (1).[18]
4F2 ð3a2 þ h2 Þ
GIC ¼ (1)
Eb2 h3
For CDCB joints, the fracture toughness was calculated using the following
equation.[18]
4F 2 m
GIC ¼ (2)
Eb2

3a2 1
where m ¼ þ
h3 h
where F is the load (N) to start crack at a crack length of a, h and b are the
thickness and width of adherend in mm. E is the Young’s modulus of
adherend. At least five specimens were tested for each type of joints. The
average fracture toughness and 95% confidence level were calculated for the
joints.

3. Results and discussion


The mechanical behaviour of epoxy/alumina nanocomposite adhesives in lap
shear and fracture in opening mode was analysed. Statistical analysis of
experimental data was performed to study the results.

3.1. Surface roughness of adherends


Before smearing the adhesive on the overlap area, surfaces were abraded
manually using abrasive papers to make it rough for better bonding.
Generally, the rough surface of bonding area gives the better bonding
strength of joint compared to smooth surface. The surface roughness of
adherends used to fabricate FT, DCB and CDCB joints were determined.
Surface roughness of adherends is reported in terms of arithmetic mean
THE JOURNAL OF ADHESION 9

Figure 4. Surface roughness of adherends.

deviation (Ra) in Figure 4. Error bars show the standard deviation of surface
roughness data of adherends. Surface roughness of aluminium alloy adher-
ends in FT joints, DCB and CDCB joints were 2.55, 1.67 and 1.61 μm,
respectively.

3.2. Lap shear strength


Representative load-displacement curves of lap shear test of FT joints bonded
with epoxy and nanocomposite adhesives having alumina nanospheres and
nanorods are shown in Figure 5. Load-displacement curves are linear for neat
epoxy adhesive and nanocomposite adhesives. Further, it is clear from the
Figure 5 that the slope of load-displacement curve of nanocomposite adhe-
sives is comparatively higher than that of neat adhesive. Rigid nanoparticles
impose constraint on the local deformation of adhesive which causes incre-
ment in the slope.

(a) (b)
Figure 5. Representative load–displacement curves for FT joints having (a) nanospheres
(b) nanorods.
10 S. K. GUPTA ET AL.

Average failure load and displacement of FT joints having alumina nano-


spheres and nanorods are mentioned in the Table 1. Failure load of FT joints
was increased with increase in wt.% of nanospheres up to 1.5 wt.%. Whereas,
further inclusion of nanospheres (2.0 wt.%) in adhesive caused reduction in
the failure load. The average failure displacement of FT joints having nano-
spheres was lower compared to single lap joints bonded with neat adhesive.
Nanocomposite adhesives became stiffer compared to neat adhesive. Similar
to FT joints with nanocomposite adhesives having nanospheres, the failure
load increased with the inclusion of nanorods in adhesive. But increment in
the failure load of joints with nanocomposite adhesives was observed only up
to 1.0 wt.% of nanorods. On further increase in the wt.% of nanorods (i.e. at
1.5 and 2.0 wt.%), the failure load of joints decreased. The failure displace-
ment of joints bonded with neat adhesive was higher than that of joints
bonded with nanocomposite adhesives having nanorods. However, the drop
in failure displacement of nanocomposite adhesives having alumina nanor-
ods (1.0 wt.%) was insignificant (7% for FT joints) in comparison to that of
adhesive having 1.5 wt.% of nanospheres (25% for FT joints) over that of the
neat epoxy adhesive.
Variation in shear strength of FT joints as a function of wt.% of alumina
nanospheres and nanorods is shown in Figure 6. Shear strength increased with
reinforcement of nanospheres in the range of 0.5 wt.% to 1.5 wt.%. Shear
strength of FT joints at 1.5 wt.% of nanospheres was 6.68 MPa, which is 52%
higher than that of joints bonded with neat adhesive. Shear strength decreased
on further reinforcement of alumina nanospheres (i.e. at 2.0 wt.%) in epoxy
adhesive. However, the shear strength of FT joints having 2.0 wt.% of nano-
spheres was higher than that of joints with neat adhesive.
Shear strength of FT joints with 1.0 wt.% of nanorods was higher than those of
neat adhesive and adhesive having 0.5, 1.5 and 2.0 wt.% nanorods. Shear
strength of FT joints having 1.0 wt.% of nanorods was 6.83 MPa, which are
55% higher than those of neat adhesive joints. At higher wt.% of nanoalumina,
the failure crack passes through the agglomerate of nanoparticles in adhesive. As
the nanoparticles had contact with each other but did not have bonding in
agglomerate. A comparison of shear strength of all the types of joint is given in
Table 1. It is clear from Table 1 that there is less difference in maximum shear
strength of FT joints having 1.5 wt.% of nanospheres and 1.0 wt.% of nanorods.
Shear strength of FT joints having 1.0 wt.% of nanorods was higher (3%)
compared to that of FT joints having 1.5 wt.% of nanospheres. Thus, it is
interesting to observe that alumina nanorods provides higher improvement in
shear strength of joints even at lower wt.% of nanoparticles. And hence nano-
composite adhesive having alumina nanorods will have higher specific shear
strength under tension loading. The use of nanocomposite adhesive having
alumina nanorods is suggested where weight of the joint is one of the critical
parameters for design.
Table 1. Average failure load, shear strength and failure displacement of FT joints.
S. No. Wt.% of nanoalumina Nanospheres Nanorods
Force (kN) Shear Strength % increase over neat Displacement Force Shear Strength % increase over neat Displacement
(MPa) adhesive (mm) (kN) (MPa) adhesive (mm)
1 0.0 1.42 4.39 – 0.28 1.42 4.39 – 0.28
2 0.5 1.67 5.18 18 0.25 1.63 5.05 15 0.27
3 1.0 1.73 5.36 22 0.15 2.20 6.83 56 0.26
4 1.5 2.06 6.68 52 0.21 1.43 4.59 5 0.15
5 2.0 1.59 4.95 13 0.13 1.90 5.90 35 0.27
THE JOURNAL OF ADHESION
11
12 S. K. GUPTA ET AL.

(a) (b)
Figure 6. Effect of wt.% of (a) nanospheres and (b) nanorods on shear strength of FT joints.

The improvement in shear strength of nanocomposite adhesives due to


reinforcement of nanoalumina could be attributed to the efficient stress
transfer between nanoparticles and epoxy matrix. Due to the transfer of
local stress easily onto the joined particles, the local plastic deformation of
matrix could be larger by the crack pinning and crack bridging mechanism
between the alumina nanoparticles and epoxy. Thus, strength of the joint can
be assumed to be higher when the particles are in intimate contact with the
polymer matrix. Further, the dependence of the shear strength of the nano-
composite adhesives to the wt.% of nanoalumina could be attributed to
following reasons. The cohesive strength of nanocomposite adhesives
increases as the number of points of interaction between epoxy and nanoa-
lumina in adhesive joint increases. Nanoalumina fills any microscopic void
or gap present in the adhesive during synthesis of nanocomposite adhesives.
Extensive contact regions between the nanoalumina and adhesive enhance
the mechanical interlocking. This led to the higher mechanical strength of
the interface.[6] Diameter of alumina nanorods is smaller than that of nano-
spheres. So nanorods will be able to fill voids or gaps (even the smaller ones)
in better manner than the nanospheres. Further, the increase of shear
strength may be attributed to the cohesive fracture resistance provided by
the nanoparticles with the help of crack blunting mechanisms in the
matrix.[15]
Shear strength decreased at higher nanoalumina content (2.0 wt.% of
nanospheres and, 1.5 and 2.0 wt.% of nanorods). It seems that once nano-
particles fully filled the porosities and gaps, further inclusion of nanoparticles
could not interact effectively within the epoxy adhesive and consequently led
to poor matrix infiltration. TEM images of bulk nanocomposite adhesive
having 0.5, 1.0, 1.5 and 2.0 wt.% of nanospheres and nanorods are shown in
Figure 7.[19] It could be observed from Figure 7 that the nanorods at higher
wt.% have higher tendency of agglomeration compared to nanospheres. Due
to agglomeration the nanoalumina did not interact properly with epoxy. At
THE JOURNAL OF ADHESION 13

(a) (e)

(b) (f)

(c) (g)

(d) (h)

Figure 7. Dispersion of (a) 0.5, (b) 1.0 (c) 1.5 (d) 2.0 wt.% of alumina nanospheres and (e) 0.5, (f)
1.0, (g) 1.5 (h) 2.0 wt.% of alumina nanorods in bulk nanocomposite sample.[19].
14 S. K. GUPTA ET AL.

1.5 wt.% of nanospheres, the size of agglomerate is less than 100 nm whereas
for nanorods it is more than 100 nm. At higher wt.% of nanoalumina, a non-
uniform distribution of nanoalumina within the adhesive matrix takes place
due to the increase of viscosity.[19] Reduction in inter-particulate distance at
higher wt.% of nanoparticles could be another reason for the decrease in
shear strength at higher nanoalumina content.[15] TEM images of nanocom-
posite adhesive having nanospheres and nanorods at 2.0 wt.% are given in
Figure 7 (g) and Figure 7 (h), respectively. It is clear from Figure 7 (g) and
7 (h) that the agglomeration is higher in the case of 2.0 wt.% of nanoparticles
than that of nanocomposite adhesive having 1.5 wt.% of nanoparticles.
Fractured surfaces of joints were analysed to find out modes of failure.
Fractured surfaces of FT joints with nanocomposite adhesives having 1.5 wt.%
nanospheres and 1.0 wt.% of nanorods are shown in Figure 8. The mode of
failure is not very predictable as it is not possible to monitor the growth of
crack fully due to their dependence on large number of factors[20] such as
number of micro-cracks, flaws, defects etc. presents at the interface or within
the adhesive layer, uniformity of roughness on bonded surfaces, uniformity of
thickness of adhesive layer (ideally, the thickness of adhesive layer shall be
uniform), reaction of bonding surface with the environment etc. The max-
imum stress intensity would be observed at free edges. So that the failure in the
single lap joints was initiated at the free edges. Mixed mode of failure (cohesive
and interfacial failure) was observed.

3.3. Fracture toughness


Opening mode fracture toughness of epoxy/alumina nanocomposite adhe-
sives using DCB and CDCB joints was determined.

3.3.1. Double cantilever beam (DCB) joints


Representative fracture toughness of DCB joints bonded with neat epoxy and
nanocomposite adhesives having 0.5, 1.0, 1.5 and 2.0 wt.% of alumina nano-
spheres and nanorods as a function of crack length is shown in Figure 9. It

(a) (b)
Figure 8. Fractured surfaces of FT joints having (a) 1.5 wt.% of nanospheres (b) 1.0 wt.% of
nanorods (Bonded area is shown in the box).
THE JOURNAL OF ADHESION 15

(a) (b)
Figure 9. Representative fracture toughness of DCB joints as a function of crack length for neat
and nanocomposite adhesives (a) nanospheres and (b) nanorods.

was not possible to arrest the crack growth exactly at an equal crack length in
each specimen bonded with the nanocomposite adhesives for comparison.
So, that the initiation fracture toughness (from load to start the crack) of
DCB joints were averaged over the initial crack length as shown in Table 2.
The value of fracture toughness increases for the new crack length even
though the new crack length is larger than the initial crack length.[12] The
value of fracture toughness was observed to be minimum at initial crack
length.
Fracture toughness, in general, increased at higher crack length because the
plastic zone size ahead of the crack tip increased with the crack length which in
turn would require higher energy. It is clear from Table 2 that fracture
toughness of DCB joints increased with increase in the wt.% of alumina
nanospheres (up to 1.5 wt.%). Further, inclusion of nanospheres (i.e. at
2.0 wt.%) in composite adhesive led to decrease in fracture toughness of joints
for both types of the adherend compared to adhesive having 1.5 wt.% of
nanospheres. In the case of alumina nanorods, fracture toughness of DCB
joints with nanocomposite adhesive also increased with increase in wt.% of
nanorods in adhesive but only up to 1.0 wt.%. Fracture toughness of DCB
joints decreased on further inclusion of nanorods (i.e. at 1.5 and 2.0 wt.%).

Table 2. Fracture toughness at initial crack length of DCB joints bonded with
nanocomposite adhesives.
S. No. Wt.% of Initiation fracture toughness (J/m2)
nano-alumina (± standard deviation)
Nano-spheres Nano-rods
1 0.0 8.67 ± 2.33 8.67 ± 2.33
2 0.5 11.75 ± 2.57 13.83 ± 2.44
3 1.0 15.33 ± 6.44 16.04 ± 6.44
4 1.5 17.55 ± 3.60 13.47 ± 4.82
5 2.0 11.70 ± 4.96 11.83 ± 3.02
16 S. K. GUPTA ET AL.

Variation in the average fracture toughness of DCB joints bonded with


nanocomposite adhesives as a function of wt.% of nanoparticles is shown in
Figure 10. Average fracture toughness of joints was calculated by taking the
average of fracture toughness value of all the five specimen of same type of
joint. Whereas average fracture toughness of a specimen was calculated by
taking the average of all values of fracture toughness obtained at different
crack lengths.
To investigate the effect of shape of alumina nanoparticles on fracture
toughness of DCB joints, Figure 10 is analysed. Average fracture toughness of
DCB joints with nanocomposite adhesives having 0.5, 1.0, 1.5 and 2.0 wt.% of
nanospheres was 144, 184, 407 and 97% higher than that of DCB joints with
neat adhesive, respectively. Similar to DCB joints having nanospheres, the
average fracture toughness of DCB joints with nanocomposite adhesives
having 0.5, 1.0, 1.5 and 2.0 wt.% of nanorods was 98, 156, 119 and 55%
higher than that of DCB joints bonded with neat adhesive.
Average fracture toughness of DCB joints was maximal for adhesive
having 1.5 wt.% of nanospheres and 1.0 wt.% of nanorods. It is clear from
Figure 10 that the reinforcement of nanospheres in adhesive imparts higher
improvement to fracture toughness of adhesive than the nanorods. Average
fracture toughness of DCB joints having 1.5 wt.% of nanospheres was 98%
higher than the average fracture toughness of joints having 1.0 wt.% of
nanorods.
Mechanisms responsible for the improvement in fracture toughness of
nanocomposite adhesives having alumina nanoparticles could be crack
deflection and twisting, crack tip blunting and crack bridging or pinning.
In the case of composites with nanospheres, crack deflection and crack tip
blunting would be the dominant mechanisms. Crack bridging might take
a secondary role in the nanocomposite adhesives having nanospheres due to
their spherical shape and untreated surface. As the epoxy is reinforced with

(a) (b)
Figure 10. Average fracture toughness of DCB joints as function of wt.% of (a) nanospheres and
(b) nanorods in adhesive.
THE JOURNAL OF ADHESION 17

untreated alumina nanoparticles, chances of formation of a strong chemical


bond (covalent bond) between the alumina particles and epoxy matrix are
very little. Whereas, the crack bridging will play a primary role and crack
deviation and crack tip blunting might have played secondary role in the
enhancement of fracture toughness of nanocomposite adhesive having
nanorods. Alumina nanorods have lesser diameter (average value
7.56 nm)[21] than that of nanospheres (average diameter, 35 nm) due to
which energy requirements for crack deflection will be little and crack tip
blunting will be less occurred. Whereas the long length of nanorods (average
length 40.56 nm)[21] causes entanglement of epoxy chain which in turn
increases the energy requirement for pull out of rods and provides better
crack bridging than the nanospheres.
The decrease in the fracture toughness of nanocomposite adhesives having
2.0 wt.% of nanospheres and, 1.5 and 2.0 wt.% of nanorods could be attributed
to the agglomeration of nanoparticles (see Figure 7). TEM images of bulk
epoxy/alumina nanocomposite adhesives having nanospheres and nanorods
are shown in Figure 7. It was observed that the nanorods had higher tendency
of agglomeration compared to nanospheres. A line contact between nanorods
and point contact between nanospheres may be assumed when two nanopar-
ticles form an agglomerate. Due to line contact between nanorods, higher
restriction will be offered by nanorods in complete wetting of alumina nano-
particles with epoxy compared to nanospheres. This lead to poor interaction
between nanorods and epoxy at higher wt.% causing detrimental effect on
enhancement in fracture toughness in comparison to nanospheres.
To look into the mode of failure present on the fractured surface, analysis
of fractured surfaces was conducted for each sample. Representative failure
surfaces of DCB joints having 1.5 wt.% of nanospheres and 1.0 wt.% of
nanorods are shown in Figure 11. Mixed mode of failure, cohesive and
interfacial failure, was observed on fractured surfaces of joints having nano-
spheres and nanorods. It is clear from the Figure 11 that the crack propa-
gated from one adhesive/adherend interface to the opposite interface by
propagating through the adhesive layer. It indicates that nanoalumina
improved the interface between adhesive and adherend, preventing the
crack propagation along the interface and causing ‘stick-slip’ crack growth
behaviour.[16]

3.3.2. Contoured double cantilever beam (CDCB) joints


Variation in fracture toughness as function of crack length of CDCB joints
having alumina nanospheres and nanorods is shown in Figure 12. The
fracture toughness (from load to start the crack) of CDCB joints were
averaged over the initial crack length and is listed in Table 3. The variation
in fracture toughness of CDCB joint is not significant with increase in the
crack length when compared with that of DCB specimen.
18 S. K. GUPTA ET AL.

(a)

(b)
Figure 11. Fractured surfaces of DCB joints having (a) 1.5 wt.% of nanospheres, (b) 1.0 wt.% of
nanorods.

(a) (b)
Figure 12. Representative fracture toughness of CDCB joints as a function of crack length for
neat and nanocomposite adhesives having (a) nanospheres and (b) nanorods.

Table 3. Fracture toughness at initial crack length of CDCB joints bonded with
nanocomposite adhesives.
S. No. Wt.% of Initiation fracture toughness (J/m2)
nano-alumina (± standard deviation)
Nano-spheres Nano-rods
1 0.0 33.55 ± 4.63 33.55 ± 4.63
2 0.5 59.21 ± 8.88 40.50 ± 2.87
3 1.0 64.66 ± 5.19 67.51 ± 6.03
4 1.5 84.56 ± 15.50 40.19 ± 7.82
5 2.0 53.34 ± 4.48 37.86 ± 4.36

Fracture toughness of CDCB joints increased with the increase in wt.% of


nanospheres (up to 1.5 wt.%) in adhesive. Fracture toughness of CDCB joints
decreased on further inclusion of alumina nanospheres (2.0 wt.%). Similarly,
the fracture toughness of CDCB joints increased due to the reinforcement of
THE JOURNAL OF ADHESION 19

nanorods (up to 1.0 wt.%) in adhesive and decreased on further addition of


nanorods (i.e. at 1.5 and 2.0 wt.%).
Average fracture toughness of CDCB joints as a function of wt.% of
nanospheres and nanorods is shown in Figure 13. Average fracture toughness
of CDCB joints having 1.0 wt.% of nanorods was 111% higher compared to
that of CDCB joints bonded with neat adhesive. From Figure 13, the effect of
shape of alumina nanoparticles on fracture toughness of CDCB joints could
be analysed. Similar to DCB joints, the maximum average fracture toughness
of CDCB joints was obtained for adhesive having 1.5 wt.% of nanospheres
and 1.0 wt.% of nanorods. Average fracture toughness of CDCB joints having
1.5 wt.% of nanospheres was 191, 42, 27 and 66% higher than that of fracture
toughness of CDCB joints bonded with neat epoxy adhesive, nanocomposite
adhesive with 0.5, 1.0 and 2.0 wt.% of nanospheres, respectively. It is clear
from the Figure 13 that the average fracture toughness of joints with nano-
composite adhesives having nanospheres is generally higher than that of
joints having nanorods. It could be observed that the reinforcement of
alumina nanospheres is advantageous over the alumina nanorods.
The fracture toughness of neat and nanocomposite adhesives of DCB and
CDCB joints could be compared through data mentioned in Table 4.
Fracture toughness of nanocomposite adhesives obtained using DCB joints
was comparatively higher than that of nanocomposite adhesive using CDCB
joints for both types of alumina particles (see Table 4). Higher scatter in data
of fracture toughness was observed in DCB tests. The difference in the value
of fracture toughness of adhesives (Araldite 2015 and Sikaforce 7752) for the
two test geometries (DCB and CDCB) has been reported by Lopes et al.[22]
The change in fracture toughness could be attributed to the different degree
of constraint in the adhesive layer. Further the different degree of curing of
adhesive in the two test geometries would be responsible change in fracture
toughness.

(a) (b)
Figure 13. Average fracture toughness of CDCB joints as function of wt.% of alumina (a)
nanospheres and (b) nanorods in adhesive.
20 S. K. GUPTA ET AL.

Table 4. Comparative analysis of average fracture toughness of joints bonded with nanocompo-
site adhesives.
S. No. Wt.% of Average fracture toughness (J/m2)
nano-alumina (±standard deviation)
Nanospheres Nanorods
CDCB DCB CDCB DCB
Joints Joints Joints Joints
1 0.0 31.59 ± 01 33.87 ± 14 31.59 ± 01 33.87 ± 14
2 0.5 64.66 ± 05 82.7 ± 36 41.76 ± 01 67.08 ± 14
3 1.0 72.49 ± 09 96.18 ± 27 66.74 ± 04 86.78 ± 22
4 1.5 92.07 ± 09 171.61 ± 56 40.01 ± 04 74.43 ± 10
5 2.0 55.31 ± 05 66.73 ± 22 38.41 ± 02 52.57 ± 14

Representative fractured surfaces of CDCB joints with nanocomposite


adhesives having 1.5 wt.% of nanospheres and 1.0 wt.% of nanorods are
shown in Figure 14. Mixed mode of failure (cohesive and interfacial) was
observed on the surfaces for both type of joints having alumina nanospheres
and nanorods.

4. Conclusions
Nanocomposite adhesives of epoxy reinforced with 0.5, 1.0, 1.5 and 2.0 wt.%
of alumina nanospheres and nanorods were synthesised by In-situ polymer-
isation technique. Epoxy adhesive reinforced with 0.5, 1.0, 1.5 and 2.0 wt.%
of nanospheres and nanorods had significant effect on shear strength as well
as opening mode fracture toughness. Following points are concluded from
present analysis.
• Shear strength was maximal for joints prepared with nanocomposite
adhesives having 1.5 wt.% of alumina nanospheres. Shear strength of FT
joints having 1.5 wt.% of nanospheres was 52% higher than that of neat
adhesive.
• In the case of nanocomposite adhesives of alumina nanorods, maximum
shear strength was obtained at 1.0 wt.% of nanorods. Shear strength of FT
joints having 1.0 wt.% of nanorods was 56% higher than that of joints
bonded with neat epoxy adhesive.
• For an equal amount of increment in shear strength of FT joints with
nanocomposite adhesives, lesser amount of alumina nanorods will be
required than the alumina nanospheres.
• Average fracture toughness of joints prepared with nanocomposite
adhesive having 1.5 wt.% of alumina nanospheres was maximal for both
types of joints (DCB and CDCB). Whereas, in the case of alumina nanorods,
the highest value of fracture toughness was obtained for adhesive having
1.0 wt.% of nanorods.
(a)

(b)
Figure 14. Fractured surfaces of CDCB joints having (a) 1.5 wt.% of nanospheres, (b) 1.0 wt.% of nanorods.
THE JOURNAL OF ADHESION
21
22 S. K. GUPTA ET AL.

• Fracture toughness of adhesives having nanospheres was comparatively


higher than that of adhesives having nanorods at each wt.% for both types of
joints.
• Average fracture toughness of adhesives obtained from CDCB specimen
was comparatively lower than that obtained from DCB specimen at all the
wt.% of nanoalumina.

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

Funding
The authors acknowledge the Institute, Motilal Nehru National Institute of Technology
Allahabad, Prayagraj, India, for providing the financial support through plan grant (No.248/
R&C/14-15).

ORCID
Sunil Kumar Gupta http://orcid.org/0000-0003-4387-5440

References
[1] Patel, S.; Bandyopadhyay, A.; Ganguly, A.; Bhowmick, A. K. Synthesis and Properties of
Nanocomposite Adhesives. J. Adhes. Sci. Technol. 2006, 20, 371–385. DOI: 10.1163/
156856106776381794.
[2] Tabaei, M. M.; Jafari, S. H.; Khonakdar, H. A. Lap Shear Strength and Thermal Stability
of Diglycidyl Ether of Bisphenol A/epoxy Novolac Adhesives with Nanoreinforcing
Fillers. J. Appl. Polym. Sci. 2014, 131, 1–8. 40017. DOI: 10.1002/app.40017.
[3] Zhai, L. L.; Ling, G. P.; Wang, Y. W. Effect of nano-Al2O3 on Adhesion Strength of
Epoxy Adhesive and Steel. Int. J. Adhes. Adhes. 2007, 28, 23–28. DOI: 10.1016/j.
ijadhadh.2007.03.005.
[4] Bhowmik, S.; Benedictus, R.; Poulis, J. A.; Bonin, H. W.; Bui, V. T. High-performance
Nano Adhesive Bonding of Titanium for Aerospace and Space Applications. Int.
J. Adhes. Adhes. 2009, 29, 259–267. DOI: 10.1016/j.ijadhadh.2008.07.002.
[5] May, M.; Wang, H. M.; Akid, R. Effects of the Addition of Inorganic Nanoparticles on
the Adhesive Strength of a Hybrid Sol–Gel Epoxy System. Int. J. Adhes. Adhes. 2010,
30, 505–512. DOI: 10.1016/j.ijadhadh.2010.05.002.
[6] Tutunchi, A.; Kamali, R.; Kianvash, A. Effect of Al2O3 Nanoparticles on the Steel Glass
Epoxy Composite Joint Bonded by a Two Component Structural Acrylic Adhesive.
Soft. Mater. 2016, 14, 1–8. DOI: 10.1080/1539445X.2014.1003269.
[7] Turan, K.; Pekbey, Y. Progressive Failure Analysis of Reinforced-adhesively Single-lap
Joint. J. Adhes. 2015, 91, 962–977. DOI: 10.1080/00218464.2014.985379.
[8] Nassar, S. A.; Wu, Z.; Moustafa, K.; Tzelepis, D. Effect of Adhesive Nanoparticle
Enrichment on Static Load Transfer Capacity and Failure Mode of Bonded Steel–
THE JOURNAL OF ADHESION 23

Magnesium Single Lap Joints. ASME J. Manuf. Sci. Eng. 2015, 137, 051024. DOI:
10.1115/1.4030081.
[9] Imanaka, M.; Takeuchi, Y.; Nakamura, Y.; Nishimura, A.; Iida, T. Fracture Toughness
of Spherical Silica-filled Epoxy Adhesives. Int. J. Adhes. Adhes. 2001, 21, 389–396. DOI:
10.1016/S0143-7496(01)00016-1.
[10] Buchman, A.; Kenig, H. D.; Dotan, A.; Tenne, R.; Kenig, S. Toughening of Epoxy
Adhesives by Nanoparticles. J. Adhes. Sci. Technol. 2009, 23, 753–768. DOI: 10.1163/
156856108X379209.
[11] Meng, Q.; Wang, C. H.; Saber, N.; Kuan, H. C.; Dai, J.; Friedrich, K.; Ma, J. Nanosilica-
toughened Polymer Adhesives. Mater. Des. 2014, 61, 75–86. DOI: 10.1016/j.
matdes.2014.04.042.
[12] Abenojar, J.; Martínez, M. A.; Pantoja, M.; Velasco, F.; Del Real, J. C. Epoxy Composite
Reinforced with Nano and Micro SiC Particles: Curing Kinetics and Mechanical
Properties. J. Adhes. 2012, 88, 418–434. DOI: 10.1080/00218464.2012.660396.
[13] Khoramishad, H.; Khakzad, M. Toughening Epoxy Adhesives with Multi-walled
Carbon Nanotubes. J. Adhes. 2018, 94, 15–29. DOI: 10.1080/00218464.2016.1224184.
[14] Nemati Giv, A.; Ayatollahi, M. R.; Ghaffari, S. H.; Da Silva, L. F. M. Effect of
Reinforcements at Different Scales on Mechanical Properties of Epoxy Adhesives and
Adhesive Joints: A Review. J. Adhes. 2018, 94, 1082–1121. DOI: 10.1080/
00218464.2018.1452736.
[15] Ghosh, P. K.; Patel, A.; Kumar, K. Adhesive Joining of Copper Using Nano-filler
Composite Adhesive. Polymer. 2016, 87, 159–169. DOI: 10.1016/j.polymer.2016.02.006.
[16] Gude, M. R.; Prolongo, S. G.; Río, T. G. D.; Ureña, A. Mode-I Adhesive Fracture
Energy of Carbon Fibre Composite Joints with Nanoreinforced Epoxy Adhesives. Int.
J. Adhes. Adhes. 2011, 31, 695–703. DOI: 10.1016/j.ijadhadh.2011.06.016.
[17] ASTM D1002. Standard Test Method for Apparent Shear Strength of Single Lap Joint
Adhesively Bonded Metal Specimens by Tension Loading (metal-to-metal); ASTM
International: West Conshohocken, PA, 2019. DOI:10.1520/D1002-10R19.
[18] ASTM D3433. Fracture Strength in Cleavage of Adhesives in Bonded Metal Joints;
ASTM International: West Conshohocken, PA, 2012. DOI:10.1520/D3433-99R12.
[19] Hiremath, V.; Synthesis and Characterization of Epoxy-Alumina Nanocomposites:
Effect of Particle Morphology and Post Curing Temperature. Ph.D. Thesis, Motilal
Nehru National Institute of Technology Allahabad, Prayagraj, U.P./India, 2016.
[20] Arenas, J. M.; Narbo´n, J. J.; Alı´a, C. Optimum Adhesive Thickness in Structural
Adhesives Joints Using Statistical Techniques Based on Weibull Distribution. Int.
J. Adhes. Adhes. 2010, 30, 160–165. DOI: 10.1016/j.ijadhadh.2009.12.003.
[21] Hiremath, V.; Shukla, D. K. Effect of Particle Morphology on Viscoelastic and Flexural
Properties of Epoxy Alumina Polymer Nanocomposites. Plast. Rubber. Compos. 2016,
45, 199–206. DOI: 10.1080/14658011.2016.1159778.
[22] Lopes, R. M.; Campilho, R. D. S. G.; DaSilva, F. J. G.; Faneco, T. M. S. Comparative
Evaluation of the Double-Cantilever Beam and Tapered Double-Cantilever Beam Tests
for Estimation of the Tensile Fracture Toughness of Adhesive Joints. Int. J. Adhes.
Adhes. 2016, 67, 103–111. DOI: 10.1016/j.ijadhadh.2015.12.032.

S-ar putea să vă placă și